You are on page 1of 358

PUR E AND I\ I'PI.

I ED MATHEMAT I CS
A Series of and Textbooks
BANACH
ALGEBRAS
an introduction
Ronald Larsen
PURE AND APPLIED MATHEMATICS
A Series of Monogaphs and Textbooks
COORDIATOR OF THE EDITORIAL BOARD
S. Kobayashi
UNIVERSITY OF CAI.IFORNIA AT BERKELEY
1. K. YANO. Integral Formulas in Riemannian Geometry (1970)
2. S. KOBAYASHI. Hyprbolic Manifolds and Holomorphic Mappings (1970)
3. V. S. VLADIMIOV. Equations of Mathematical Physics (A. Jefrey, editor: A.
Littlewood. translator) (1970)
4. B. N. PSHLNICIINYI. Necessarv Conditions for an Extremum (L. Neustadt, trans
lation editor; K. Makowski, transiator) (1971)
S. L. NARICI, E. BECKI.NSI tIN, and G. 8oCHMAN. Functional Analysis and Valua
tion Theory (1971)
6. D. S. PASS'tAN. Infnite Group Rings (1971)
7. L. DORNIIOFF. Group Representation Theory (in two pas). Part A: Ordinary
Represntation Theory. Par B: Modular Represntation Theor
y
(1971. 1972)
8. W. BOOTHY and G. L. WEISS (ed. ). S}mmetric Space: Shor Courses Presnted
at Washington University (1972)
9.
Y. MA'I SlSHIMA. DifTrentiabJe Manifolds (E. T. Kobayahi. translator ) (1972)
10. L. E. WAR, JR. Topology: An Outlne for a First Course (1972)
11.
A. BA\KIfNIAN. Cohomological Mthods in Group Theory (1972)
12. R. GILM[R. Multiplicative Ideal Theory ( 1972)
13.
J. YEH. Stochastic Processes and the Wiener Integra) (1973)
14.
J. B,\KROS-NETO. Introduction to the Theory of Distrihutions (1973)
IS. R. LARSIN. Functional Analysis: An Introduction (1973)
16. K. Y,\NO and S. ISIIIIHRA. Tangent and Cotangent Hundles: Diferential Geometry
(1973 )
17.
C. PROCISI. Rings with Polynomial Idenlitiee ( 1973)
18. R. HIRMANN. Geometry, Physics. and Systems (1973)
19.
N. R. W-LI.CH. Harmonic Analysis on Homogeneous Spaces (1973)
20. J. DU.II))lNE. Introduction to the Theor) of Formal Groupe (1973)
21. I. VAISM.-N. Cohomoloy and Di ferent i al Forms ( 1 973)
22. B.-Y. CHEN. Geometry of Submanifolds (1973)
23. M. MARCUS. Finite Dimensional Multilinear Aigehra (in tWO paris, (1973)
24. R. LARS[N. Banach Algebras: An Introduction ( 1973 )
In Preparatioll:
K. B. STOLARSKY. Algebraic Numbr and Diophantine Approxmation
BANACH
ALGEBRAS
an introduction
RONALD LARSEN
DEPAR r M ~ T OF M4TIII M\TJ('S
WI SLEY.\N UNIVERSITY
M IOIlI FTO\\ N. CONN I C I ICl T
MARCEL DEKKER, INC. New York 1973
COPYRIGHT 1973 by MARCEL DEKKER, INC.
ALL RIGHTS RESERVED
Neither this book nor any part may be reproduced or transmitted
in any form or by any means, electronic or mechanical, including
photocopying, microfilming, and recording, or by any information
storage and retrieval system, without permission in w r t n ~ from
the publisher.
MARCEL DEKKER, INC.
95 Madison Avenue, New York, New York 10016
LIBRARY OF CONGRESS CATALOG CARD NUMBER: 73-84868
ISBN: 0-8247-6078-6
PRINTED IN THE UNITED STATES OF AMERICA
To
Joan
and
Nils Erik
PREFACE
The exposition in the following pages is an elaboration and
expansion of lectures I gave to second year mathematics graduate
students at Wesleyan University during the academic years 1970-71
and 1971-72. The aim of the exposition is to provide a compact intro-
duction to the theory of Banach algebras that not only acquaints the
reader with fundamental portions of the abstract theory but also illu-
strates the usefulness of Banach algebras in the study of harmonic
analysis and function algebras and gives the reader the basic tools
necessary for further work in these areas. The first half of the
book is devoted primarily to the general theory of Banach algebras,
while in the second half the emphasis is on various more specialized
topics related to harmonic analysis and function algebras - among
which are: Wiener's Tauberian Theorem, the problem of spectral
synthesis, the Bishop, Choquet, and ilov boundaries, representing
measures, Wermer's Maximality Theorem, the Commutative Gel'fand-
-Naimark Theorem, Plancherel's Theorem, the Pontryagin Duality Theorem,
almost periodic functions, and the Bohr compactification.
An intelligent reading of the book presupposes the usual mathe-
matical equipment possessed by second year mathematics graduate
students with regard to topology, algebra, and real and complex
analysis, as well as a reasonably good knowledge of basic functional
analysis. That portion of functional analysis which is necessary
forms a subset of my earlier book in this series "Functional Analysis:
An Introduction" [L], and I have retained, for the most part, the
same notation in this volume as in the previous one. Results with
which the reader is assumed to be familiar are frequently cited
without comment. However, in almost all such instances an appropriate
reference is given.
v
vi
Preface
Unfortunately, due to a lack of time and endurance, there are
no problem sets per!! in this volume. However, I have left unproved
results scattered throughout the exposition and the reader is strongly
urged to fill in these lacunae in order to test and strengthen his
or her understanding of the subject.
The conclusion of a proof is indicated by the symbol 0 at the
right hand margin.
I would like to thank all the graduate students at Wesleyan who
passed through my course while this book was evolving for their
comments and suggestions. In particular, I wish to thank Hans Engenes
and Michael Paul for their often perspicacious observations and
questions, and Polly Moore Hemstead for her valuable editorial
assistance.
I would also like to thank Helen Diehl who typed the original
manuscript for the book.
Finally, thanks are due to the editors and staff of Marcel Dekker
for their cheerful and expert cooperation during the production of
the book.
Middletown, Connecticut
May, 1973
C01\iENTS
PREFACE v
CHAPTER 1: FUNDAMENTALS OF BANACH ALGEBRAS 1
1.0. Introduction 1
1.1. Basic Definitions and Some Algebraic Preliminaries 3
1.2. Examples of Normed Algebras 16
1.3. A Theorem of Gel'fand 23
1.4. Regularity and Quasi-regularity 26
1.5. The Gel'fand-Mazur Theorem 35
1.6. Topological Zero Divisors 40
CHAPTER 2: SPECTRA 53
2.0. Introduction 53
2.1. Definitions and Basic Results S3
2.2. The Polynomial Spectral Mapping Theorem and the
Spectral Radius Formula S7
2.3. A Theorem of ilov on Spectra 60
CHAPTER 3: TIlE GEL' FAND REPRESENTATION THEORY 64
3.0. Introduction 64
3.1. Maximal Regular Ideals and Complex Homomorphisms 6S
3.2. The Maximal Ideal Space 68
3.3. The Gel'fand Representation 74
3.4. The Beurling-Gel'fand Theorem 79
3.5. Semisimplicity 81
vii
viii
CTER 4:
T GEL'FAD REPRESENATION OF SOE SPECI FIC
ALGEBRS
4.0.
4.1.
4.2.
4.3.
4.4.
4.5.
4.6.
4.7.
4.8.
CHPTER 5:
5.0.
5.1.
5.2.
5.3.
5.4.
Introduction
C(X) and C
o
(X)
C
n
( fa, b] )
L (X,S,)
C
A(D)
Finitely Generated Baac Algebras
AC(r)
LI (G)
M(G)
SEISIMPLE COMTATIVE BAACH ALGEBRS
Introduction
Te Gel'fand Representation of Semisiple
Comutative Banach Algebras
-
A -as a Baach Algebra
Homomorhism ad Isomorhism of Comutative
Baach Algebras
Characterization of Singular Elements in
Contents
85
85
85
92
92
95
98
101
104
127
129
129
130
133
135
Self-adjoint Semisimp1e Comutative Banach Algebras 139
CHTER 6: AALYIC FUNCTIONS AD BAACH ALGEBRS
6.0.
6.1.
6.2.
6.3.
6.4.
CHTER 7:
7.0.
Introduction
Aalytic Functions of Baach Algebra Elements
Some Consequences of the Preceding Section
Zeros of Entire Functions
-1
Te Connected Component of the Identity in A
RGUL CO ATIVE BAACH ALGEBR
Introduction
7.1. Te Hll-Kerel Topology and Reglar Comutative
Banach Algebras
142
142
142
151
154
155
159
159
160
Contents
7.2.
7.3.
CHAPTER 8:
8.0.
8.l.
8.2.
8.3.
8.4.
8.5.
8.6.
8.7.
CHAPTER 9:
9.0.
9.l.
9.2.
9.3.
9.4.
Some Examples
Normal Commutative Banach Algebras
IDEAL THEORY
Introduction
Tauberian Commutative Banach Algebras
Two Tauberian Theorems
The Problem of Spectral Synthesis
Local Membership in Ideals
Ditkin's Theorem
Ll(G) Satisfies Ditkints Condition
Some Further Remarks on Ideals
BOUNDARIES
Introduction
Boundaries
The Silov Boundary
Some Examples of Boundaries
Extreme Points, the ~ i o v Boundary, and the
Choquet Boundary
ix
167
174
180
180
181
189
192
200
204
209
21S
218
218
219
222
227
232
9.S. Some Applications of Boundaries 243
9.6. Representing Measures and the Choquet Boundary 249
9.7. Characterizations of the Choquet Boundary 259
9.8. Representing Measures and the ~ i o v Boundary 269
CHAPTER 10: B*-ALGEBRAS 272
10.0. Introduction 272
10.l. B*-Algebras 273
10.2. The Gel'fand-Naimark Theorem 277
10.3. The B*-Algebra A (G) 280
0
10.4. P1ancherel's Theorem 297
10.5. The Pontryagin Duality Theorem 308
x
Contents
1
0
.6.
A Spectral D
e
composition Te
o
r
e
m for Self-adjoint
contiuous Linear Transf
o
rmat
ions
1
0
.7. Almst Periodic Functions
REFERENCES
INDEX
316
323
333
337
"Those as hunts treasure must go alone, at night, and
when they find it they have to leave a little of their
blood behind them."
Unknown woman from Bimini
CHAPTER I
FUNDAMENTALS OF BANACH ALGEBRAS
1.0. Introduction. One of the central objects of study in
functional analysis is the normed linear space, that is, a linear
space A over a scalar field, usually the real or complex numbers,
together with a function from A to the real numbers, called the
norm on A, which is positive definite, homogeneous, and subadditive.
In elementary functional analysis the focus of attention is generally
restricted to the linear structure of normed linear spaces. However,
a goodly number of the specific spaces that occur in functional
analysis come equipped, in a more or less natural way, with additional
algebraic structure beyond that of a linear space. In particular,
many of these spaces are algebras, and the multiplication in these
algebras is continuous with respect to the given norm. The goal of
and the succeeding chapters is to introduce the reader to the
study of such algebras, generally called normed algebras, and, in
particular, to examine what additional information the algebraic
structure of certain specific spaces can reveal about them.
The development we shall present is intended to be neither the
most general possible nor the most exhaustive of the theory discussed.
Rather it is designed to expose the reader to the study of normed
algebras as abstract mathematical entities and to various questions
that arise in such a general and at the same time to inves-
tigate in some detail a limited selection of quite specific normed
algebras. In order to accomplish such a program within a reasonable
amount of space we have been forced to be rather selective in the
scope of our development. Thus, for example, we shall only consider
algebras over the complex numbers and shall focus most of our
1
2
1.
F
u
ndaental
S of Baach Algebras
attent
i
o
n on thos
e no
r
ed a
lgebras where
the
u
d
er
ly
i
ng nored linear
space is a Banach space, that is, on Banac
h
al
g
ebras. Moreover,
al
though no ass
u
ption of comutativit
y is gene
rally made in the
first twO ch
apte
r
s, almost all
o
f
t
he subs
equ
e
nt
m
aterial is concerned
o
nly with comutative Banach algebras.
Simi
l
arly, with regard to
s
pe
cific Banach algebras, we
h
ave placed
p
r
im
a
r emphasis on algebras
of continuous functions and
al
g
e
b
ras that
occu
r in haronic analysis,
and given secondary status to algebras of continuous linear trans
forations and of functions of several comlex variables. More
compendious treatments of the general th
ery can be found, for
instance, in [N, Ri].
The present chapter is concered with developing some of the
fundaental theory of normed algebras. We begin with the basic
definitions and a collection of essentially algebraic results. Many
of the latter may be already fa
i
liar to the reader, whereas others
may seem strange. Te utility of some of the concepts and theorems
discusse may not be apparent util subsequent chapters. This is
true also of some of the other results to be considered in this
chapter. Although some of these results are considered only to
round out the topic under discussion, the majority will find appli
cation at later stages of the development.
After this section on fund
aentals we shall present a nuber of
concrete exaples of nored algebras, followed by a discussion of a
theorem due to Gel'fand. Te next two sections contain a treatment
of the important topic of inversion in Banach algebras with identity
and its couterart, quas
i
-inversion, in Banach algebras without
identity. We shall see, for exaple, that in the forer case the
invertible eleents for a open subgroup of the algebra and that
the operation of inversion is continuous. Tese results will then
be applied to prove the Gel'fand-Mazur Theorem, which asserts that
every Baach algebra with identity in which each nonzero element is
inverible is isometrically isomorphic to the complex nubers. Or
stadi assuption that all algebras are over the complex nubers
1.1. Definitions and Preliminaries 3
is crucial for the validity of this result. The chapter concludes
with a rather extensive discussion of the notion of topological
zero divisors, including a proof of Arens' Theorem, which asserts
that an element in a commutative Banach algebra A with identity
is a topological zero divisor if and only if it is not invertible
in any Banach algebra B with identity in which A can be embedded.
We note that throughout this volume the letters C, lR, and Z
_ shall stand for the complex numbers, the real numbers, and the
integers, respectively. The basic notation and terminology from
functional analysis is as in [L].
1.1. Basic Definitions and Some Algebraic Preliminaries. As the
heading of this section indicates, we shall concern ourselves here
with introducing some of the basic definitions needed in the suc-
ceeding development and in proving several essentially algebraic
results. Some of the definitions and results are probably already
familiar to the reader, and we include them primarily in the interests
of completeness.
Definition 1.1.1. A linear space A over C is said to be
an algebra if it is equipped with a binary operation, referred to
as multiplication and denoted by juxtaposition, from A X A to A
such that
(i) x(yz) = (xy)z,
(ii) x(y+ z) = xy + xz; (y+ z)x = yx+ zx,
(iii) a(xy) = (ax)y = xray). (x,y,z E A; a E C).
It is said that A is a commutative algebra if A is an algebra
and
(iv) xy = yx (x,y E A),
whereas A is an algebra with identity if A is an algebra and
there exists some element e E A such that
(v) ex = xe = x
(x E A).
4 1. Fundamentals of Banach Algebras
It is evident that, if A is an algebra with identity, then
the identity element e is unique. To avoid triviality we shall
always assume that our algebras !!!. just !!!!. element.
Definition 1.1.2. A nonned linear space (A,II11l over C is
said to be a nonned algebra if A is an algebra and
\IxYIl IIxllllYIl (x, yEA)
A nonned algebra A is said to be a Banach algebra if the normed
linear space (A,IIII) is a Banach space.
A number of examples of normed algebras will be discussed in
the next section.
Definition 1.1.3. Let A be an algebra. Then I C A is said
to be a left (right) ideal in A if I is a linear subspace of A
such that xl C I (Ix C I), x E A; I C A is said to be a two-sided
ideal if I is both a left and a right ideal in A. An ideal I C A
is proper if I A, and a proper left (right, two-sided) ideal I
is said to be maximal if, whenever J C A is a left (right, two-sided)
ideal in A such that I C J, either I = J or J = A. Further-
more, I C A is said to be a subalgebra if I is a linear subspace
such that x,y E I implies xy E I.
Clearly every ideal in an algebra A is a subalgebra, and in a
commutative algebra all ideals are two-sided. In the latter case we
shall generally speak only of ideals, and drop the adjective "two-
sided." Note also that, by definition, a maximal ideal is always
proper.
The following result concerning normed algebras is easily proved.
The details are left to the reader.
Theorem 1.1.1. Suppose A is a normed algebra.
(i) The completion of A is a Banach algebra.
1.1. Definitions and Preliminaries
5
(ii) If I C A is a subalgebra, then
whereas if A is a Banach algebra and I
then I is a Banach algebra.
I is a normed algebra,
is a closed subalgebra,
(iii) If I C A is a closed two-sided ideal, then the quotient
space A/I is a normed algebra with the usual quotient norm
llIx + I III = inf IIx + yll
yEI
and with the multiplication
(x E A)
(x + I) (y + I) = xy + I (x,y E A).
(iv) If A is a Banach algebra and I C A is a closed
two-sided ideal, then A/I is a Banach algebra.
If an algebra A has an identity, then it obviously is meaningful
to discuss the algebraic notion of inversion. Moreover, as we shall
see, it is always possible to consider an algebra without identity
as a subalgebra of an algebra with identity, thereby allowing a
discussion of inversion even in the case of algebras without identity.
However, although this embedding process is very useful, it is also
desirable to be able to discuss some sort of concept of inversion in
algebras without identity without recourse to this process of embed-
ding. This fact will lead us to the notion of quasi-inversion. We
shall discuss inversion and quasi-inversion for normed linear algebras
in greater detail in Section 1.4. We now wish to concentrate mainly
on the algebraic aspects of these ideas.
Definition 1.1.4. Let A be an algebra with identity e. An
element x E A is said to have a left (right) inverse if there
exists some yEA such that yx = e (xy = e), whereas x is said
to have an inverse if there exists some yEA such that xy = yx = e.
If x E A has an inverse, then x is said to be regular, or
invertible, and x is said to be singular if it is not regular.
It is easily verified that, if x E A has both a left inverse
y and a right inverse z, then y = z is an inverse
6
I. Fundamentals of Banach Algebras
and that inverses are unique. As usual, we shall denote the inverse
of x by x-I. Obviously, if the algebra A is commutative, then
left and right inverses are inverses.
There is a simple connection between the existence of left and
right inverses and membership in ideals, as evidenced by the next
theorem.
Theorem 1.1.2. Let A be an algebra with identity and suppose
x E A. Then the following are equivalent:
(i) x has a left (right) inverse.
(ii) x does not belong to any proper left (right) ideal in A.
Proof. Suppose x has a left inverse y. If I is any left
ideal such that x E I. then e = yx E I. and so A = Ae C AI C I
shows that I = A. Thus x belongs to no proper left ideal.
Conversely, if x has no left inverse, then it is easily seen that
I = (yx lyE AJ is a left ideal in A such that x E I. Further-
more, I is proper since e I.
The proof of the remainder of the theorem is similar.
Corollary 1.1.1. Let A be an algebra with identity. If
x E A is regular, then x does not belong to any proper left,
right, or two-sided ideal in A.
In general for results involving left, right, and two-sided
concepts we shall prove only one case, leaving the others for the
reader.
In particular, we note that the identity e belongs to no
o
proper ideal. This observation, combined with Zorn's Lemma [OSl' p. 6].
yields the following result, the details of which are left to the
reader:
1.1. Definitions and Preliminaries
7
Theorem 1.1.3. Let A be an algebra with identity. If I C A
is a proper left (right, two-sided) ideal, then there exists a
maximal left (right, two-sided) ideal MeA such that I C M.
Corollary 1.1.2. Let A be an algebra with identity and
suppose x E A. Then the following are equivalent:
(i) x has a left (right) inverse.
(ii) x does not belong to any maximal left (right) ideal in A.
Moreover, if x is regular, then x does not belong to any maximal
two-sided ideal in A.
In particular, for a commutative algebra A with identity the
preceding results say that x E A is regular if and only if x
does not belong to any proper ideal in A if and only if x does
not belong to any maximal ideal in A.
Theorem 1.1.3, though algebraically quite elementary, is a
crucial tool in studying normed algebras with identity. The
verbatim counterpart of the theorem is, however, not valid for algebras
without identity. The appropriate analog can nevertheless be obtained
if we restrict our attention to so-called regular or modular ideals.
Definition 1.1.5. Let A be an algebra. A left (right, two-
sided) ideal I in A is said to be regular if there exists some
u E A such that xu - x E I (ux - x E 1, xu - x E I and ux - x E I),
x in A.
The element u is called an identity modulo I. Clearly, if
A has an identity e, then every ideal is regular, and e is an
identity modulo any ideal. The reason for this terminology is
apparent from the third portion of the next proposition.
Proposition 1.1.1. Let A be an algebra.
(i) If I C A is a proper regular left (right, two-sided)
ideal and u is an identity modulo I, then u I.
8
1. Fundamentals of Banach Algebras
(ii) If I C A is a regular left (right, two-sided) ideal and
J C A is a left (right, two-sided) ideal such that J ~ I, then
J is regular. Moreover, if u is an identity modulo I. then u
is an identity modulo J.
(iii) If I C A is a proper regular two-sided ideal, then A/I
is an algebra with identity.
Proof. If I is a proper regular left ideal and u is an
identity modulo I, then the assumption that u E I entails that
xu E I, x E A, whence
x = xu - (xu - x)
E I, x E A, contradicting
the properness of 1. Thus
u ~ I .
For part (ii) of the proposition we note that, if u is an
identity modulo the regular left ideal I, then xu - x E I C J,
x E A, and so J is regular and u is an identity modulo J.
Finally, it is evident that A/I is an algebra, and we claim
that u + I is an identity for A/I, where u E A is an identity
modulo I. The latter assertion is immediate on noting that
(u + I) (x + I) = ux + I
= x + I
= xu + I
= (x+I)(u+I) (x E A)
since ux - x E I and xu - x E I, x E A.
o
We can now state and prove the indicated analog of Theorem 1.1.3.
Theorem 1.1.4. Let A be an algebra. If I C A is a proper
regular left (right, two-sided) ideal, then there exists a maximal
regular left (right, two-sided) ideal MeA such that I C M.
Proof. Suppose
identity modulo I.
ideals J C A such
I is a regular left ideal and
Denote by J the collection of
that J ~ I. Clearly J ~ ~
let u be an
all proper left
each J E ] is
1.1. Definitions and Preliminaries 9
a proper regular left ideal, and u is an identity modulo J.
Moreover, u ~ J, J E J. The last two observations are consequences
of Proposition 1.1.1 (i) and (ii). We introduce a partial ordering
in J by setting J
l
> J
2
if and only if J
I
~ J
2
, J
1
,J
2
E J. If
(J 1 is a linearly ordered subset of J, then it is easily verified
a
that J = U J is a proper regular left ideal in A that contains
a a
I. The properness follows from the fact that u l J.
Consequently we may apply Zorn's Lemma [DS
I
, p. 6] to deduce the
existence of a maximal element M E J. Evidently M is a maximal
regular left ideal such that M I.
D
Next we wish to examine in detail how to embed an algebra without
identity in an algebra with identity. Suppose that A is an algebra
without identity and denote by A[e] the set of all pairs
(x,a), x E A, a E C, that is, as a point set A[e] = A x . The
point set becomes an algebra if one defines the linear space opera-
tions and multiplication as follows:
(i) (x,a) + (y,b) = (x+y,a+b)
(ii) b(x,a) = (bx,ba)
(iii) (x,a) (y,b) = (xy + ay + bx,ab) (x,y E A; a,b E C).
The routine verification will be left to the reader. Furthermore,
the element e = (0,1) E A[e] is an identity for A[e]. Indeed,
(x,a)(O,I) = (xO+ lx,a)
= (x,a)
= (0,1) (x,a) (x E A; a E C).
Moreover, it is easily shown that the mapping L: A - A[e], defined
by ~ x ) = (x,O), x E A, is an algebra isomorphism of A onto the
maximal two-sided ideal L(A) = ((x,O) I x E AJ C A[e].
This discussion, plus some additional. argument, leads to the
following theorem. The details are omitted.
10 1. Fundamentals of Banach Algebras
Theorem 1.1.5. Let A be an algebra. Then A[e] is an algebra
with identity and the mapping A - A(e] is an algebra isomorphism
of A onto the maximal two-sided ideal = [(x,O) l x E A} C A[e].
If A is a normed algebra. then A[e1 is a normed algebra under
the norm
IICx.a)ll = llxll + lal
(x E Aj a E C),
and A - A[e] is an isometric algebra isomorphism onto the
closed maximal two-sided ideal A[e]. Moreover, in this case,
the quotient algebra is isometrically isomorphic to c,
and A[e] is a Banach algebra provided A is a Banach algebra.
Evidently A is a commutative algebra if and only if A[e] is
a commutative algebra.
As a rule we shall speak of A itself as a maximal two-sided
ideal in ACe]. Similarly we shall often write elements (x,a) E A[e]
as (x,a) = x + ae. This makes sense on identifying x with
= (x,D) and recalling that e = (0,1). Hopefully the simplifi-
cation in notation gained by these conventions will outweigh the loss
of precision.
In the succeeding chapters we shall always use the symbol Are]
to denote the algebra with identity obtained from an algebra without
identity by the previously developed construction. We shall occa-
sionally speak of this process as that of adjoining identity to A.
It should be noted that even when A has an identity we can
still construct A[e] and Theorem 1.1.S is valid. The identity for
A is, however I not the identity for Are]. This observation will
at times be useful, but the majority of our applications of A[e] will
be to algebras A without identity.
The general utility of algebra Are1 lies in the fact
that algebras with identity are often easier to deal with than
algebras without identity, and one can often deduce properties of A
1.1. Definitions and Preliminaries 11
by examining a related property in A[e]. a first example of
this let us examine the relationship between ideals in A
and ideals in Are].
Theorem 1.1.6. Let A be an algebra. let 1 denote the
e
family of all proper ieft (right. two-sided) ideals in Are] that
are not contained in A, and let I denote the family of all proper
regular left (right, two-sided) ideals in A. Then lEI if and
only if I = I n A for a unique I E 7
e e e
Proof. Suppose
I fI. A and let I =
e
I C A[e] is a proper left ideal such that
e
InA. Clearly I is a left ideal in A. and,
e
moreover, I is proper, because if I = A, then I A, and so
e
since Ie contradicting the properness of Ie. Also. Ie = A[e],
since I r:f. A,
e
there exist some x E A and some a E C, a 0,
such that x + ae E I. Furthermore, x # 0, as if x = 0, then
e
eEl
e
again contradicting the properness of t by Theorem 1.1.2.
e
Let u = -x/a. Then u - e = -x/a - e = (-l/a) (x + ae) E I. We
e
claim that u is an identity modulo I.
I
e
we see that
is a left ideal. But A
yu - y = y(u - e) E A.
Indeed, if yeA, then in A[e]
yu - y = yu - ye = y(u - e) E Ie' as
is a two-sided ideal in Are], so that
Consequently yu - y E I n A = I,
e
and we see that u is an identity
I is a proper regular left modulo the left ideal that is,
ideal.
Conversely. suppose I is a proper regular left ideal in A and
let u E A be an identity modulo I. Set Ie = {y , y E Are]. yu Ell.
We claim that I is a proper left ideal in Are] such that
e
I A and I = InA. Clearly I is a linear subspace of Are1.
e e e
Further, we note that I is also a left ideal in A[e] because if
x E I and y = z + ae E A[e], then yx = (z + ae)x = zx + ax E I,
as J is a left ideal in A. It then follows at once that I is
a left ideal in A(e]. Moreover.
I = A[e],
e
then eu = u E I,
e
Ie is proper, since if
contrary to the conclusion of
12
1. Fundamentals of Banach Algebras
Proposition l.l.l(i).
2
(u - e)u = u - u I,
I f/. A.
e
Next we note that u - eEl
e
as u is an identity modulo
because
I. Thus
Finally, if x E I, then xu - x E I, as u is an identity
modulo I, and so xu E I. Thus I C InA. Conversely, if
e
x E I n A, then xu E I and xu - x E I, from which we conclude
e
that x E I. Therefore I = InA. Similar arguments, of course,
e
establish the result for right and two-sided ideals.
To complete the proof of the theorem we must show that, if I C A
is a proper regular two-sided ideal, then there exists a unique proper
two-sided ideal I (A[e), I A, such that I = InA. So
e e e
suppose I and J are two such ideals in A[e]. Then, as above,
e e
we can deduce the existence of identities modulo I, call them u
and v, such that u - eEl, v - e J. Since u and v are
e e
identities modulo I, it follows that vu - v I and vu - u E I,
whence v - u E I.
Now suppose y = z + ae E I. Then since I and A are both
e e
two-sided ideals in A[e], we see that uy = uz + au E I n A = I.
e
Thus
z + au = (z - uz) + (uz + au) E I + I = I,
as z - uz E I. Consequently
y = z + ae = z + au + aCv - u) + aCe - v)
I+I+J
CJ,
e
e
as J n A = I. Hence r C J. A similar argument mutatis mutandis
e e e
demonstrates that J C I, and so I = J . 0
e e e e
Corollary 1.1.3. Let A be an algebra. Then MeA is a
maximal regular two-sided ideal if and only if there exists a unique
maximal two-sided ideal M C A[e] such that M A and
e e
M = M n A.
e
1.1. Definitions and Preliminaries
This corollary will be useful when we discuss the complex
homomorphisms of commutative Banach algebras in Chapter 3.
13
If A has an identity, then the adjective "regular" in the
statement of Theorem 1.1.6 and Corollary 1.1.3 is obviously redundant.
Next let us see how the algebra A[e] can be used to introduce
a concept of inversion in algebras without identity. Suppose A is
an algebra without identity and let x E A. One could clearly ask
whether or not x has a left (right) inverse in A[e], but it turns
out for our purposes that it is more useful to ask this question for
e - x, rather than for x. Thus we see that e - x has a left
inverse in A[e] if and only if there exist some yEA and a E
(ae - y)(e - x) = ae - ax - y + yx = e,
that is, if and only if
(a - 1 ) e = ax + y - }'x.
Since ax + y - yx E A, it follows at once that a = 1, as A is
without identity. lienee we see that e - x has a left inverse in
Are] if and only if there exists some yEA such that
x + y - yx = o. This element y A will be the substitute we seek
for the left inverse of x in an algebra A without identity. Note
also that the equation defining this element y is meaningful in
algebras with identity as well.
With these observations in mind we make the following definition:
Definition 1.1.6. Let A be an algebra. An element x E A is
said to have a left (right) guasi-inverse if there exists some yEA
such that y 0 x = y + x - yx = 0 (x 0 y = x + y - xy
=
0) , and x
is said to have a 9,uasi-inverse if there exists some yEA such
that y 0 x = x 0 y = O. If x E A has a quasi-inverse, then x
is said to be quasi-regular, or quasi-invertible, and x is said to
be quasi-singular if it is not quasi-regular.
14
1. Fundamentals of Banach Algebras
Evidently Y is a left (right) quas1-1nverse for x if and
only if x is a right (left) quasi-inverse for y. Some other ele-
mentary results concerning quasi-inverses are collected in the next
proposition.
Proposition 1.1.2.Ci). If A is an algebra with identity e,
then x E A has a left (right) quasi-inverse if and only if e - x
has a left (right) inverse, and x is quasi-regular if and only if
e - x is regular. Moreover, yEA is a left (right) quasi-inverse
for x if and only if e - y is a left (right) inverse for e - x,
and y is a quasi-inverse for x if and only if e - y is an inverse
for e - x.
(ii) If A is an algebra and x E A has a left quasi-inverse
y and a right quasi-inverse z, then y = z is a quasi-inverse for
x. In particular, quasi-inverses are unique.
Proof. The proof of part (i) of the proposition is a routine
computation and will be omitted. For part (ii), considering the
context of our computations to be A[e] in the case that A is
without identity, we see from part (i) that e - y and e - z are
left and right inverses for e - x, respectively, and so, by the
remarks following Definition 1.1.4, we conclude that e - y = e - z
is a two-sided inverse for e - x, whence y = z is a two-sided
quasi-inverse for x by part (i).
In general, if x E A is quasi-regular, we shall denote its
quasi-inverse by x_I'
There are analogs for quasi-inversion to Theorem 1.1.2 and
Corollary 1.1.2.
Theorem 1.1.7. Let A be an algebra and x EA. Then the
following are equivalent:
(i) x has a left (right) quasi-inverse.
(ii) (-z + zx I z E A) = A({-z + xz I z E AJ = A).
o
1.1. Definitions and Preliminaries 15
Proof. Suppose x has a left quas1-1nverse y. Then
x = -y + yx E {-z + zx I zEAl. Clearly (-z + zx I z E AJ is a
left ideal in A, and so for any w E A we have
w = (w - wx) + wx E {-z + zx I zEAl,
that is, {-z + zx zEAl = A.
Conversely, if (-z + zx I z E AJ = A, then there exists some
yEA for which x = -y + yx,
left quasi-inverse.
that is, y 0 x = o. Thus x has a
This result, combined with Proposition 1.1.2(i), immediately
yields the next corollary.
o
Corollary 1.1.4. Let A be a commutative algebra with identity
and suppose x E A. Then
(i) x is regular if and only if {zx I z E A} = A.
(ii) x is singular if and only if (zx I zEAl is a proper
ideal in A.
The analog of Corollary 1.1.2 is the next theorem. The proof
is left to the reader.
Theorem 1.1.8. Let A be an algebra and suppose x E A. Then
the following are equivalent:
(i) x has a left (right) quasi-inverse.
(ii) If MeA is a maximal regular left (right) ideal, then
there exists some yEA such that y 0 x E M (x 0 y EM).
Of course in commutative algebras left and right quasi-inverses
are quasi-inverses and the previous results assume a somewhat simpler
form.
Finally we note that some authors define, for example, the left
quaSi-inverse of x as any element y such that x + y + yx = O.
This is the case, for instance, in [HIP, p. 680; N, p. 158]. The
16
1.Fundamentals of Banach Algebras
definition of quasi-inverses given here, however, seems to be the
more common one (see, for example, [HR
l
, p. 471; Lo, p. 64; Ri, p. 16;
Wa, p. 19]). The relationship between the two definitions is that
y is a left (right) quasi-inverse for x in the sense used here if
and only if -y is a left (right) quasi-inverse for -x in the
sense used in [HIP, N].
1.2. Examples of Normed Algebras. In this section we wish to
list a number of standard examples of normed algebras occurring in
functional analysis. The examples can be divided, loosely speaking,
into three classes: function algebras, convolution algebras, and
algebras of linear transformations. We shall not prove any of the
assertions made about the following examples, but instead leave the,
generally routine, verifications to the reader.
Example 1.2.1. The complex numbers C ~ i t the usual algebraic
operations and with absolute value as the norm are a commutative
Banach algebra with identity.
Example 1.2.2. Let X be a locally compact Hausdorff topologi-
cal space. By C(X), C (X), and C (X) we denote, respectively,
o c
the algebras of all continuous complex-valued functions on X that
are bounded, vanish at infinity, or have compact support. The
algebra operations are the usual ones of pointwise addition, multi-
plication, and scalar multiplication. With the usual supremum norm
lIfH = sup If (t) I
CD t (X
(f E C(X,
the algebras C(X) and C (X) are commutative Banach algebras,
o
whereas C (X) is a commutative normed algebra. If X is noncompact,
c
then only C(X) is an algebra with an identity, whereas if X is
compact, then C(X) = Co(X) = Cc(X) is an algebra with identity.
The algebra Cc(X) is a Banach algebra only in the case that X is
compact.
1.2. Examples of Normed Algebras
17
1.2.3. Let 0 denote the closed unit disk in C, that
is, 0 = (, I , E C, 1'1 < 1), and let A(O} denote the family of
all f E CeO) such that f is analytic on
int(D) = {, , , E c, "I < 13. With the usual pointwise operations
and the supremum norm
IIfli = sup I f (C) ,
CD ,E 0
(f E A(D)),
it is easily verified that A(O) is a commutative Banach algebra
with identity.
Example 1.2.4. More generally, let K CC be any infinite
compact set. We define P(K) and RCK}, respectively, as all those
f E C(K) that can be approximated uniformly on K by polynomials or
by rational functions with no poles in K. In analogy to A(D), A(K)
denotes all those f E C(K) that are analytic on the interior of K,
that is, on int(K). With pointwise operations and the norm of C(K),
it is easily seen that peK), R(K), and A(K) are all commutative
Banach algebras with identity. Moreover, it is apparent that
P(K) C R(K) C A(K) C C(K). For K = 0 it is a classical fact that
P(O) = A(O), whereas for K = r = (, I , E C, I" = 1) the Stone-
Weierstrass Theorem implies that R(f) = C(f). In general, the
question of when various pairs of algebras in the chain
P(K) C R(K) C A(K) C C(K) are equal, as well as the analogous
problem with C replaced by is a very difficult one. Although
we shall not investigate this question in any great detail, we shall
make some additional comments on it in later chapters. For more
detailed discussions the reader is referred to [B, Ga, Lb, S, Wm
l
,
Wm
2
]
Example 1.2.5. Let a,b a < b, and for each nonnegative
integer n let Cne[a,b]) denote the family of all n-times contin-
uously differentiable complex-valued functions defined on the closed
interval [a,b], with the usual convention about one-sided deri-
vatives at the end points of the interval. With pointwise operations
18 1. Fundamentals of Banach Algebras
and the norm
n
sup [1:
a<t<b k=O
n
(f C ([a,b],
where f(k) denotes the kth derivative of f, the space Cn([a,b])
becomes a commutative Banach algebra with identity. The choice of
the norm in Cn([a,b]) is made precisely to ensure that
IIfgll
n
< IIfllnllglln,f,g E Cn([a,b]).
Example 1.2.6. Let G be a locally compact Abelian topological
group. We assume, for convenience, that the topology on such a group
is always Hausdorff and that the group operation is written additively.
If f is a complex-valued function on G, then for each s E G we
denote by T (f) the function T (f)(t) = f(t - s), t E G. Further-
s s
more, AP(G) will denote the subalgebra of all those f E C(G) such
that the set (T (f) I s E G) has compact closure in C(G). The
s .
functions in AP(G) are said to be almost periodic. It can be shown
that AP(G) is a commutative Banach algebra with identity_
Example 1.2.7. Let be a positive measure space and
let L = L denote the family of all equivalence
CD CD
classes of essentially bounded complex-valued functions
on X. With the operations of addition, multiplication, and scalar
multiplication of equivalence classes obtained via pointwise operations
on equivalence class representatives and with the usual essential
supremum norm
IIfli = ess sup If(t)1
CD t Ex
= inf {M , I If(t)1 > M) = 0)
M
(f E L
CD
it can be shown that L is a commutative Banach algebra with
CD
identity.
Example 1.2.8. Let G be a locally compact topological group
and suppose that A is a left Haar measure on G; that is, A is a
1.2. Examples of Normed Algebras 19
regular positive Borel measure on G such that ~ s + E) = ~ E )
for each 5 E G and each Borel set E C G. For I < P < e, let
Lp(G) denote the usual Banach space of equivalence classes of
Borel measurable complex-valued functions on G whose pth powers are
integrable with respect to ~ and with the norm
(f E L (G)).
p
Here L (G) will denote the Banach space L G , ~ ) as defined in the
m _
preceding example. For 1 < p < -, f E Ll(G), and gEL (G), we
- - p
define f * g for almost all s E G by
f * g(s) - IG f(s + t)g(-t) dA(t),
where again we have written the group operaticn additively. It can
be shown that the above definition defines an element f * g, called
the convolution of f and g, which belongs to L (G), and, more-
p
over, the following inequality is valid.
It is now apparent, with convolution as the operation of multi-
plication, that LI (G) is a Banach algebra. It will be a commutative
Banach algebra if and only if G is an Abelian group.
If G is a compact topological group and the Haar measure ~ is
normalized so that ~ G ) = 1, then the preceding inequality combined
with HHlder's inequality reveals that L (G), I < p < c, is a Banach
p --
algebra with convolution as multiplication. Again, these algebras
are commutative if and only if G is Abelian.
When G is an Abelian group, the defining relation for convolu-
tion can be written in the more familiar form
f * g(s) = IG f(s - t)g(t) dA(t).
In this case we shall subsequently prove that LICG) is a commutative
Banach algebra with identity if and only if G is a discrete Abelian
20
1. Fundamentals of Banach Algebras
group.
For example, suppose G = Z, the additive group of the integers
with the discrete topology. Then clearly the elements of LI(Z) can
be identified with the doubly infinite sequences {akl of complex
numbers such that Ik=-e lakl converges. We assume, as is usually
done, that the Haar measure on Z is normalized so that
A({n}) = 1, n E Z. The convolution of two sequences (akl and (b
k
)
in LI(Z) is evidently the sequence
e
* (b
k
) = {1: an_kb
k
}.
k=-e
The identity element in L1(Z) is the sequence (e
k
), where eO = 1,
e
k
= 0, k o. This is easily verified.
More generally, let M(G) denote the Banach space of all bounded
regular complex-valued Borel measures on G with the total varia-
tion norm
= (G)
E M(G.
Here denotes the nonnegative measure known as the total vari-
ation of Given E M(G), we define, for each Borel set
E C G,
* = IG - t)
It can be shown that this identity defines an element * of
M(G), called the convolution of and and that
II .. * 1l .. It is then easily verified that M(G) with con-
volution as multiplication is a Banach algebra with identity. The
identity in M(G) is the measure 6
0
with unit point mass concen-
trated at the identity element of the group G; that is, for any
Borel set E C G, 6
o
(E) = 1 if 0 E E and 6
o
(E) = 0 if 0 l E.
Note that in the preceding sentence we have used "0" to denote the
identity element in both and G. We shall continue to do this
in the following pages, where the interpretation of the symbol "0"
will be obvious from the context of its use. If G = Z, then it is
1.2. Examples of Normed Algebras 21
easily seen that M(Z) = Ll(Z), and this is also true for any
discrete group G. The algebra M(G) is commutative if and only
if G is Abelian.
For the reader who is not familiar with the fundamentals of the
theory of topological groups and the notion of convolution we suggest
the following references: [HR
I
, pp. 15-105, 184-195, 262, 274,
283-298; Lo, pp. 108-133; N, pp. 357-372; Nb, pp. 49-119; Po;
Ri, pp. 318-331].
Example 1.2.9. Let us n ~ w consider Ll(f) , where, as before,
r = (c ICE c, Ici = 1) = (e
1t
I -n < t < n). It is easily verified
that r is a compact Abelian topological group with the relative
topology on r as a subset of C and with the group operation of
multiplication of complex numbers. The normalized Haar measure on
r is given by
that is, ~ is just Lebesgue measure on (-n,n] divided by 2n if
one considers r as being identified with the interval (-n,n].
Given f E Ll(f) and k E we define, in the classical way,
the kth Fourier coefficient of f by
Evidently, with each f E LI(f) we can thus associate a bounded

doubly infinite sequence f = (f(k)]. Classical results in Fourier

analysis assert that the mapping f ~ f, generally called the
Fourier transformation, is a norm-decreasing algebra isomorphism
from the commutative Banach algebra LI(f) onto the subalgebra
FLl(f) of C(Z) consisting of all those doubly infinite sequences
(at] ~ C(Z) for which there exists some f E Ll(f) such that
~ = f(t), k E Z. It is then not difficult to show that FLI(r) is
a commutative Banach algebra under pointwise operations if one
22 I.Fundamentals of Banach Algebras
..
defines the norm of (a
k
) = (fCk) E FLl(f) by
lI(
a
k
lll = Uflll = ~ fn If(t)1 dt.
Note that as Banach algebras Ll(f) and FLl(f) are isometrically
isomorphic. However, FLl(f) considered as a subalgebra of C(Z) is
not closed, and so FLl(f) with the supremum norm is not a Banach
algebra.
We shall return to an analog of this example for general commu-
tative Banach algebras in Section 5.2.
Example 1.2.10. Denote by AC(f)
such that t=_oolf(k)I converges, that
is not difficult to verify that AC(f)
the set of all f E C(n C Ll (f)
..
is, such that f E LI(Z). It
is a commutative Banach algebra
with identity under pointwise operations and the norm
oo" "
II f li
AC
= E ,f(k) 1 = Uflll
k=-oo
(f E AC(I1).
This algebra is generally called the algebra of absolutely convergent
Fourier series. The algebra AC(f) is a proper subset of C(f).
Actually, if we now had at our disposal an appropriate Fourier
transform for the elements of Ll(Z) , then it could easily be seen
that AC(f) is really just FLl(Z) where the algebra FLl(Z) is
defined as in Example 1.2.9, mutatis mutandis.
Example 1. 2 .11. Let (V, n'lI) be a Banach space over C and
let L(V) denote the space of all bounded linear transformations
from V to itself. As is well known, L(V) is a Banach space with
the norm
11111
= sup IIT(x)1I
IIxll<l
xEV
(T E L(V.
With the usual notion of composition of linear transformations as
multiplication, it is easily seen that Lev) is a Banach algebra
1.3. A Theorem of Gel'fand 23
with identity. Except in the case that V is one-dimensional L(V)
is not commutative.
An important subalgebra of L(V) is C(V) which consists of
all the T E L(V) that are compact; that iS
J
T E lev) for which
the image T(E) of a norm bounded set E C V is a set with compact
closure in the norm topology. The algebra C(V) is a noncommutative
Banach algebra unless the dimension of V is one. and it is without
identity unless the dimension of V is finite.
1.3. A Theol"etll of Gel' fand. The theorem of this section assert!
that we can weaken one of the requirements in the definition of a
Banach algebra A and still essentially obtain the same sort of
mathematical object. In particular. if we assume only that multipli-
cation in A is separately continuous, that is, the mapping
(x)y) ~ xy, x,y E A, is continuous in y for each x and continuou!
in x for each y, then there exists a norm on A equivalent to
the original norm on A and under which A is a Banach algebra as
defined in Definition 1.1.2. In this sense we need not require the
validity of the nora inequality IIxyll < IIxlillyll, X,Y E A, in order
to obtain a B8nach algebra. This weaker bypothesis was the one
originally used in [GRkl]. We make this precise in the following
theorem:
Theorem 1.3.1 (Gel'fand). Let A be an algebra whose under-
lying linear space is a Banach space (A,UU). If the operation of
.ultiplication from A X A to A is separately continuous, then
there exists a noI'll IIIIU on A. that is equivalent to II II and
such that
111 xy III < III x III iIIYIII
(x,y E A),
that is, A is a Banach algebra with the norm 1ll1I1.
Proof. We shall give the proof in detail only for the case that
A. is without identity. The argument when A has an identity is
essentially the s e ~ but so.ewhat simpler.
24 I.Fundamentals of Banach Algebras
Consider Are], that is, A with an identity adjoined, as a
Banach space, under the norm Ily + aell = lIyll + lal, yEA, a \C.
For each x E A define the linear mapping T A[e] ~ A[e] by
x
T (y + ae) = xy + ax
x
(y + ae E A [ e]) .
Since multiplication in A is separately continuous, it follows at
once that T E L(A[e]). Moreover,
x
liT II = sup' liT (y + ae) II
x lIy+aell=1 x
> sup lIaxll
- lal=l
= IIxll
(x E A).
The last estimate, together with some elementary verifications,
reveals that the mapping ~ : A ~ L(A[e]), defined by ~ x ) = T ,
x
x E A, is an algebra isomorphism from A onto a subalgebra of the
Banach algebra L(A[e]). The norm in L(A[e]) is, naturally, the
usual norm for continuous linear transformations. We now renorm A
by setting III x 111 = liT II, x E A. It is easily checked that A is a
x
normed algebra under the norm 111111 and that ~ : A ~ L(A[eD is
an isometric algebra isomorphism from (A, III III) to ~ A ) C L(A[e]).
To see that (A, IU 111) is a Banach space it suffices to show that
~ A ) is a closed linear subspace of L(A[e]).
To this end, suppose {x} C A is a sequence such that {T }
n x
converges in L(A[e)) to T. Then {T 1 is a Cauchy sequence n
x
n
in L(A[eD, whence, since liT II> IIxU, x E A, we see that (x 1
x - n
is a Cauchy sequence in (A,IIII). Let x E A be such that
limnllxn - xII = O. This is possible because (A,IIII) is a Banach
space. But then it is an immediate consequence of the separate
continuity of multiplication and the following estimate
1.3. A Theorem of Gel'fand
lIT(y + ae) - Tx{y + ae)1I < IIT(y + ae) - T (y + ae)1I
Xn
+ liT (y + ae) - T (y + ae)1I
x x
n
< IIT(y + ae) - T (y + ae)lI
- Xn
+ IIx y - xyll
n
2S
+ lI ax
n
- axil (n = 1,2,3, )
that T(y + ae) = Tx(Y + ae)
and ~ A ) is closed.
for each yEA, a E ~ . Thus T = T ,
x
Finally, we note that the equivalence of 1111 and 111111 on A
follows from the Two-Norm Theorem [L, p. 184], as (A,IIII) and
{A, III 1lI) are both Banach spaces and IIxll ~ IlIxlll, x E A.
o
In the case that A is an algebra with identity e, we of
course identify A with the subalgebra of L{A) consisting of the
transformations T, x E A, where T (y) = xy, yEA. Having done
x x
this, we see that
llIelll = liT II = sup liT (y) II = 1.
e llyll=l e
More generally, if A is any normed algebra with identity e under
the norm 1111, then it is easily seen that
which shows that the norm IIlelll = 1. Thus we see that by renorming a
normed algebra with identity as shown here we can obtain a normed
algebra with an equivalent norm where the norm of the identity e is
equal to one. Consequently in the following chapters we shall always
assume, without loss of generality, that the identity in ~ normed
algebra with identity has norm one.
- -----
26 1. Fundamentals of Banach Algebras
It is perhaps worth noting that in any normed algebra with
identity e we have Uell 1, since lIel1 = lIeell lIelll1el1. We are
assuming, naturally, that we do not have to deal with the trivial
algebra A = {oj. From the construction of A[e] it is clear that
II e II = I, as e = (0, 1)
1.4. Regularity and Quasi-regularity. We now wish to study in
greater detail the notions of regularity and quasi-regularity in
Banach algebras. In particular we shall establish some sufficient
conditions for the regularity and quasi-regularity of elements in a
Banach algebra, show that the sets of regular and quasi-regular
elements are open. and show that the operations of inversion and
quasi-inversion are continuous. At the end of the section we shall
use some of these results to prove that the norm closure of a proper
regular ideal is again a proper regular ideal and that maximal regular
ideals are always closed.
To begin let us look at the question of inversion in the rather
simple Banach algebra C. A simple sufficient condition for x E C
to be regular is that II - xl < 1. Indeed, in this case, we see at
once that the geometric series Ik=O(1 - x)k converges absolutely
and that
Thus x
CD k 1
E (1 - x) = 1 _ (1 _ x)
k=O
1
= -
x
is regular, and its inverse
-1
x is given by
-1
x
CD
= 1: (1 - x)k.
k=O
This construction actually carries over mutatis mutandis to any
Banach algebra with identity. thereby giving us the next theorem.
The details are left to the reader.
Theorem 1.4.1. Let A be a Banach algebra with identity e.
If x E A and lie - xII < 1, then x is regular and
1.4. Regularity and Quasi-regularity
27
-1 k
x = e + E (e - x)
k=l
As noted in Proposition l.l.2(i), if A is a Banach algebra
with identity, then x E A is quasi-regular if and only if e - x
is regular, and Theorem 1.4.1 thus ensures that x will be quasi-
-regular provided llxll = lie - (e - x)1I < 1. Actually we can improve
this observation, as shown in the next result.
Theorem 1.4.2. Let A be a Banach algebra. If x E A and
limnllxnUl/n < I, then x is quasi-regular and
1c
X_I = - 1: x
k=l
Proof. Appealing to the Root Test [Ru
2
, p. 57] for series of
numbers, we see that ~ = 1 IIxkll converges absolutely, and hence
there exists some yEA such that
CD
Y = - 1:
k=l
k
x ~
as A is a Banach space. In particular, if we set Yn = ~ = l xk,
n = 1,2.3 . then limnllYn - yll = o.
Furthermore. we see that
n
= - 1:
k=l
k+l
x
n+1
= _ I: x'k
k=2
n+l k
=x-J: x
k=l
= x + y 1
n+
(n = 1,2.3, ).
Thus, since multiplication is separately continuous, we conclude that
xy = yx = x + y,
28 1. Fundamentals of Banach Algebras
that s ~ y 0 x = x 0 y = o.
Therefore x is quasi-regular and y = x_I'
o
Combining this theorem with Proposition 1.1.2(i), we obtain the
next corollary, of which Theorem 1.4.1 is a special case.
Corollary 1.4.1. Let A be a Banach algebra \\lith identity e.
If x E A and lim lI(e - x)nlli/n < 1, then x is regular, and
n
-1 ~ k
x = e + ~ (e - x) .
k=l
Utilizing the defining relation for quasi-inverses, Theorem
1.4.2, Corollary 1.4.1, and Proposition 1.I.2(i) again, we obtain the
following corollary:
Corollary 1.4.2. Let A be a Banach algebra.
(i) If x E A and lIxli < 1, then x is quasi-regular, and
IIxll II II IIxll
(1 + xII) < x_I < (1 -xiD .
(ii) If A has an identity e and x E A is such that
lie - xII < 1, then x is regular, and
lie - xII \I -111 lie - xII
(1 + lie - xlD < e - x <(1 - lie - xlD
Proof. To prove part (i) we note that
whence
1.4. Regularity and Quasi-regularity
from which we deduce at once that
On the other hand, we also have
IIxll = IIx_I - x_Ixll IIx_III + II
x
_
l
llll xll,
whence
Part (ii) follows from part (i) on applying Proposition
1.1.2(i).
29
We are now in a position to prove probably the most important
result about quasi-regularity: the quasi-regular elements form an
open set, and quasi-inversion is a continuous operation. First we
make the following definition:
Definition 1.4.1.
denote the set of all
-1
an identity, then A
are regular.
Let A be an algebra. By A_I we shall
x E A that are quasi-regular. If A has
shall denote the set of all x E A that
o
Theorem 1.4.3. Let A be a Banach algebra. If yEA is
quasi-regular and if x E A is such that IIxli < 1/ (1 + lIy -1"), then
(i) x + y is quasi-regular.
(ii)
Moreover, A_I is an open subset of A,
is a homeomorphism of A_I onto itself.
Proof. We note first that
and the mapping y-y
-1
30 I. Fundamentals of Banach Algebras
shows that x - Y_Ix E A_I' by Theorem 1.4.2. We claim that
is a left quasi-inverse for x + y. Indeed,
z 0 (x + y) = z + x + Y - z(x + y)
= [(x - y_lx) + (x + y_Ix)_1 - (x - y_Ix)_I(x - Y_Ix)]
+ (y + Y-
l
- y-lY) - (x - y_lx)_l(Y + Y-
l
- Y-lY)
= (x + y_
1
x)_1 0 (x + Y_lx) + Y-
1
0 Y
- ex - y_
1
x)_I(Y_
1
0 y)
= O.
Similarly it can be shown that x + Y has a right quasi-inverse.
Consequently from Proposition 1.1.2(ii) we conclude that x + Y E A_I
and that (x + Y)_I = z.
Next, appealing to Corollary 1.4.2(i), we see that
H(x + Y)-l - y_11I = II-(x - y_lx)_IY_l + (x - y_Ix)_lll
lI(x - y_lx)_lll (lIy_lll + I)
Ilx - y_lxll(lIy_l" +1)
< 1 - IIx - y xII
-1
2
IIxll(lIy_11I + I)
1 - IIxll(lIy_11I + I)
Thus parts (i) and (ii) of the theorem are proved.
Now, if yEA_I' then for any w E A such that
we see that w = (w - y) + y E A_I by part (i). Thus A_I is open.
Moreover, from Proposition 1.1.2 and the comments preceding it we see
that the mapping Y - Y-
1
, yEA_I' is a bijective mapping that is
its own inverse, If y E A_I and w E A is such that
1.4. Regularity and Quasi-regularity
lIw - yll < II CI + \ly _llil , then parts Ci) and Cii) reveal that
x E A_I and
lIw - yllClly III + 1)2
llw - y 11 < - ,
-I -I - I _ IIw - yllClIy_11I + I)
31
from which it follows immediately that the mapping y - Y-
I
tinuous and hence a homeomorphism.
is con-
o
Using much the same sort of reduction as before, we obtain the
following corollary for algebras with identity:
Corollary 1.4.3. Let A be a Banach algebra with identity e.
If yEA is regular and x E A is such that IIxll < lIlly-Ill, then
Ci) x + y is regular 1 2
Cii) IICx + y)-I _ y-III < jlxllcl/y- 1\ + 1)
- I - IIxllClly-
l
U + I)
-1 1
Moreover, A is an open s u s ~ t of A, and the mapping y - y-
is a homeomorphism of A-I onto itself.
It should be noted that, if A is only assumed to be a normed
. -1 h
algebra, then ne1ther A_I nor A need be an open set. Furt er-
more, if A is a Banach algebra with identity, then it is easily
seen that A-I is a topological group.
With the exception of the discussion in Sections 1.6 and 6.4,
we shall not investigate in any detail the topological properties of
A_lor A-I. Nevertheless a certain amount is known, and the
interested reader is referred to [Ga, pp. 90 and 91; Ri, pp. 13-15,
19-23, 280-283, 293-296; 5, pp. 7-90, 99-103].
It is also worth noting that, if A is a commutative Banach
algebra with identity, yEA-I, and lIxll < l/lly-lU, then
32
1. Fundamentals of Banach Algebras
and
CD
-1 -1 -1 -1 k
(y - x) = y + y E (y x)
k=l
This can be established via a direct computation.
Next let us return to an obvious question that we skimmed over
previously: If x is an element of a Banach algebra A, then does
lim IIx
n
ll
l/n
exist? The answer is in the affirmative, as shown by
n
the next theorem.
Theorem 1.4.4. Let A be a normed algebra. If x A. then
lim UxnU
I
/
n
exists. Moreover,
n
(i) limllxnU
1/n
= infilxnl1
1/n
.
n n
(ii) 1 imllxnll
i/n
< IIxli.
n
Proof. Let a = inf IIx
n
ll
1/n
Clearly a < lim infllxnll
1/n
.
n - n
Suppose that c > 0 and choose some positive integer m such that
UxmU
1/m
< a + c. Then for each positive integer n there exist
nonnegative integers a and b, 0 < b < m, such that
n n - n
n = a m + b. Moreover, it is easily verified that lim a mIn = I
n n n n
and lim b In = o.
n n
But then
a m+b
IIxnU
l/n
= Ux n nlilln
a In b In
UxmU n IlxU n
a mIn b In
(a + c) n IIxll n
from which it follows at once that
n
Cn = 1 J 2.3 . ) ,
Since e > 0 was arbitrary, we conclude that lim sUPnllxnlll/n < a,
and hence lim IIxnU
1/n
exists and
n
1.4. Regularity and Quasi-regularity
33
limllxnll
l/n
= infllxnll
l/n
.
n n
The fact that
lixnll < IIxlin.
lim /lxnll
l/n
< IIxll is immediate on observing that
n -
o
(
Ilx
nlll/n)
The sequence and its limit will recur repeatedly in
the succeeding chapters and play an important role in the study of
Banach algebras. We have already seen some uses of it in Theorem
1.4.2 and Corllary 1.4.1. At this juncture it also seems appropriate
to introduce a new definition.
Definition 1.4.2. Let A be a normed algebra. Then x E A is
said to be if there exists some nonnegative integer n
n
such that x = 0, and x E A is said to be topologically nilpotent
if lim IIx
n
ll
l/n
= o.
n
Evidently every nilpotent element is topologically nilpotent,
but the converse need not be so. Topological nilpotents shall also
make recurrent appearances in the sequel (e.g., in Section 1.6). Here
we wish to prove only one result about such elements.
Proposition 1.4.1. Let A be a Banach algebra with identity
e. If x E A is topologically milpotent. then x is singular.
Proof. Suppose x is regular. Then
1 = lIell = lI(x-
l
x)nlll/n
=i1(x-
l
)n(x)nIl
1
/n
< Iix-lllllxnlli/n
where we have used the fact that xx-
l
= x-Ix.
leads to a contradiction, as lim IIx
n
ll
1/n
= O.
n
Therefore x is singular.
en = 1,2,3, ... ),
But this clearly
o
34
1. Fundamentals of Banach Algebras
OUr last concern in this section will be to utilize the previous
discussion of regularity and quasi-regularity to establish a funda-
mental result about ideals in Banach algebras. It is easily seen
that. if 1 is a left (right, two-sided) ideal in a Banach algebra
A. then the norm closure of I, denoted by e1(I). is again an
ideal of the same sort and that, if I is regular, then so is cl(I).
It is, however, not as obvious, although it is true, that the closure
of a proper regular ideal is again a proper ideal.
Theorem. 1.4.5. Let A be a Banach algebra. If I C A is a
proper regular left (right, two-sided) ideal, then cl(I) is a
proper regular left (right, two-sided) ideal.
Proof. We need only prove that cl(I) is proper. Since I
is regular, there exists some u E A such that xu - x E I, x E A.
From Proposition l.l.l(i) we see that u I. We claim, moreover,
that u eICI). To show this it clearly suffices to prove that
lIu - xII l, x E I. So suppose the contrary--that is, suppose x E I
is such that lIu - xII < 1. Then. by Corollary 1.4.2.(i). u - x is
quasi-regular, and so there exists some yEA, namely,
y = (u - x)_l' such that
y 0 eu - x) = y + u - x - yeu - x)
= y - yu - (x - yx) + u
= O.
Hence u = yu - y + ex - yx}. But yu - y E I. as u is an identi-
ty modulo the regular left ideal I, and x - yx E I, as I is a
left ideal and x E I. Consequently, u E I, which contradicts the
fact that u I.
Therefore elCI) is proper.
D
An immediate and important corollary is the following resul t:
1.5. The Gel'fand-Mazur Theorem
3S
Corollary 1.4.4. Let A be a Banach algebra. If MeA is
a maximal regular left (right, two-sided) ideal, then M is a closed
maximal regular left (right, two-sided) ideal.
Thus maximal regular ideals are always closed.
It shoUld be noted that the closure of a proper ideal in a
Banach algebra need not be proper; that is, the assumption of
larity cannot in general be dropped. For example, if G is a com-
pact Abelian topological group, then L (G), I < p <., is a dense
p -
ideal in L
1
(G).
Another fundamental result in the study or Banach algebras. the
Gel'fand-Mazur Theorem, will be proved in the next section with the
aid of our knowledge of regularity and quasi-regularity.
1.5. The Gel'fand-Mazur Theorem. The Gel'rand-Mazur Theorem,
which says that every Banach algebra that is a division algebra is
isometrically isomorphic to C
t
is truly one of the most fundamental
theorems in the study of commutative Banach algebras. Its important
Tole in the investigation of such algebras will become apparent when
we develop the Gel'fand representation theory for commutative Banach
algebras in Chapter 3. The proof of the Gel'fand-Mazur Theorem that
we shall give is a standard one involving an application of Liou-
ville's Theorem. The use of Liouville's Theorem is the first example
of how the theory of functions of a complex variable enters into the
study of Banach algebras. We shall see many other examples of this
phenomenon in the succeeding chapters.
For the sake of completeness we make the following definition:
Definition 1.5.1. Let A be an algebra with identity. If each
x E A, x 1 O. is then A is said to be a division algebra.
Theorem 1.5.1 (Gel'fand-Mazur Theorem). Let A be a Banach
algebra with identity e. If A is a division algebra, then there
exists an isometric algebra isomorphism of A onto C.
36
1. Fundamentals of Banach Algebras
Proof. We shall show that A = tCe ICE el, and hence the
mapping Ce - C clearly defines a mapping from A onto C with the
desired properties. As before, we assume that lIell = 1. To prove
the above, suppose x E A and x - Ce 0, C E C. Since A is a
division algebra, it follows that (x - Ce)-l exists for each ,E
and, in particular, that x 0 and x-I exists. Let x* be a
continuous linear functional on A such that
-1
x* (x ) = 1. Such a
-1
functional exists, by the Hahn-Banach Theorem [L, p. 86], as x o.
-1
Now define the function g: C - e by gee) = x*[{x - 'e) ]" E C.
We claim that g is a bounded entire function.
Before we prove this we need to note
namely, if (x - 'e)-l and (x _ , e)-l
o
one preliminary fact;
exist, then
-1 -1 -1
(x - 'e) (x - 'oe) = (x - 'oe) (x
-1
- 'e) .
This, however, clearly holds if and only if
(x - , e)(x - 'e) = (x - 'e) (x - C e),
o 0
as inverses are unique. But elementary calculations reveal that both
sides of the last equation are equal to
thereby establishing the desired assertion.
Using this fact, we see at once that, for any',' E C,
o
ex - 'e)-l - ex - 'oe)-l = (x 'e)-l(x - 'oe)-l[(x - 'oe) - (x - 'e)]
= (' - 'o)ex - 'e)-lex - 'oe)-l.
Consequently, given
,
E C,
o
we see that
-1 -1
x*[(x - 'e) ] - x*[(x - 'oe) ]
=
C - '0
1.5. The Gel'fand-Mazur Theorem 37
-1 -1
x*[(, - 'o)(x - Ce) (x - Coe) ]
= ~ ~ ~ ~
C - ,
o
= x*[(x - Ce)-lex - Coe)-l] ec E c; C ; , ).
o
However, since the mapping C - x - ,e, C E C, is clearly continuous
from ~ to A and inversion in A is continuous, by Corollary
1.4.3, we conclude that
g(C) - g(,o)
lim
C -, C - Co
-1 -1
= x* [ex - C e) (x - C e) ] ;
o 0
o
that is, the derivative of g at , exists. Since C E C was
o 0
arbitrary, we see that g is an entire function.
To see that the entire function g is bounded it suffices to
show that lim", _cog(,) = O. But if C ~ 0, it is evident that
(x - Ce)-l = (x" - e)-I,C, whence from the continuity of inversion
in A we deduce that
-1
lim (x - Ce)
'C, -CD
'C e)-1
= lim ex -
'C, -CII C
= 0 .
Thus, since x* is continuous, we obtain
lim gCC) =
'C,-CD
-1
lim x*[Cx - 'e) ]
'C'-CD
= x* (0)
= o.
Hence g is a bounded entire function. and so, by Liouville's
Theorem [A, p. 122]. we see that g is a constant. Since
lim,C' _CIIgCC) = 0, we see, moreover, that gCC) = 0, , E C. However,
this contradicts the fact that g(O) = x*(x-
1
) = 1.
Therefore for each x A there exists some ,E C such that
x = Ce. The remainder of the proof is now apparent.
o
38 1. Fundamentals of Banach Algebras
In particular the theorem asserts that every Banach division
algebra is commutative. If we drop our standard assumption that
lieU = 1, then we can only conclude that the isomorphism from A to
C is continuous.
Proofs of the Gel'fand-Mazur Theorem which do not involve the
theory of functions of a complex variable are also available (see,
for example, [Ri, p. 40]). A somewhat different function theory
proof is to be found in [GRk, p. 128].
This is an opportune point at which to emphasize the advantages
of considering only algebras over the complex numbers. If A is a
normed division algebra over the real numbers, then it may be iso-
morphic to either the complex numbers, the real numbers, or the
quaternions (see, for example, [Ri, p. 40]).
There are some other conditons on a Banach algebra A which
ensure that it is isomorphic to C. We present two of them here.
Theorem 1.5.2 (R.E. Edwards). Let A be a Banach algebra with
identity e. If (Ix-III < 1/lIxll, x E A-I, then there exists an
isometric algebra isomorphism of A onto C.
Proof. In view of the Gel'fand-Mazur Theorem it suffices to
show that each x E A, x 0, is regular. Now evidently, if x E A,
x 0, then x belongs to the set A = {y , yEA, lIyll > p) for
p -
-1
some p > O. Thus we really need only prove that Ap C A for
each p > 0; that is, A-I = A-I n A = A for each p > o.
p p p
We note first that each Ap is connected. Indeed, suppose
x,y (Ap. If Y = -ax for some a > 0, then it is easily seen
that the circular arc from y = -ax to ax, defined by fl(t) =
aeintx, -1 t 0, followed by the straight-line arc f
2
(t) =
(1 - t)ax + tx, 0 < t < 1, determines a cantinuous arc lying com-
pletely in A whose end points are y = -ax and x. In the case
p
that y -ax for any a > 0, easy arguments show that
1.5. The Gel'fand-Mazur Theorem 39
f (t) = (1 - t) IIxll + tllyll [(1 _ t)x + ty]
11(1 - t)x + ty
(0 < t < 1)
defines a continuous arc lying in A whose end points are x and
p
y. Thus A is connected. Furthermore, A-I is an open subset of
p p
Ap in the relative topology since A-I is open by Corollary 1.4.3
and A-I as pe E A-I. Moreover, A-I is closed. Indeed,
p -1 P P
suppose (x
k
} c: Ap and x E A are such that - xII = O.
Since each x
k
E A-I nAp' we see that IIx;lll < 1/lIxkll < IIp,
k = 1,2,3, ... , and so for each nand k we have
_ = -
< IIx;lllllxn -
II
x
n
- xkll
< 2
P
Hence (Xkl) c: A is a Cauchy sequence, and so there exists some
yEA such that limkllx;l - YII = o.
But then
IIxy - ell < IIxy - xx;lll + lIxx;1 - xkx;lll
IIxlllly - x;lll + IIx - xkllUx;111
< IIxllllY - x;lll + IIx - xkll
p
(k = I, 2 , 3, ... ) ,
from which we conclude that xy = e. Similarly we see that yx = e,
-1
whence x EA. However, A is clearly a closed subset of A,
p
and so x E A-I = A-I n A .
p p
-1
Therefore Ap is a nonempty open and closed subset of the
-1 -1
connected set A, whence A = A, that is, A c: A. 0
P P P P
As a consequence of this theorem we have the following corOllary:
40 1. Fundamentals of Banach Algebras
Corollary 1.5.1 (Mazur). Let A be a Banach algebra with
identity e. If lIxyll = IIxllllylL x,y (. A, then there exists an iso-
metric algebra isomorphism of A onto ~
Proof. If x E A-I, then 1 = lIe\l = IIxx-
1
1i = IIxllllx-
l
ll,
whence lix-lll = l/lIxli.
o
1.6. Topological Zero Divisors. In this section we wish to
consider another standard algebraic concept, that of a zero divisor,
and extend it in a natural way to normed algebras. The objects we
thus obtain will be called topological zero divisors. After examin-
ing examples of topological zero divisors in some specific algebras,
we shall establish some necessary and sufficient conditions for an
element in a Banach algebra to be a topological zero divisor and
discuss some of the connections between topological zero divisors
and the notions of regularity and quasi-regularity. In particular,
we shall prove Arens' Theorem, which asserts that in a commutative
Banach algebra A with identity an element x is a topological
zero divisor if and only if it is a singular element in some Banach
algebra containing A.
We begin with a definition and some examples.
Definition 1.6.1. Let A be a normed algebra. Then x E A
is said to be a left (right) zero divisor if there exists some yEA,
Y , 0, such that xy = 0 (yx = 0), and x is said to be a two-sided
zero divisor if there exists some yEA, Y , 0, for which xy =
yx = O. Furthermore, x (A is said to be a left (right) topologi-
cal ~ divisor if there exists a sequence {Yk) C A such that
lIy
k
ll = 1, k = 1,2,3, ... , and limkllxYkll = 0 (limkllYkxll = 0), and
x is said to be a two-sided topological ~ divisor if there exists
a sequence {Y
k
) C A for which IIYk
ll
= 1, k = 1,2,3, ... , and
limkllxYkll = limkllYkx!l = O.
Clearly a zero divisor of any sort is a topological zero divisor
of the same sort. If the algebra A is commutative, then the notions
1.6. Topological Zero Divisors
41
of left, right, and two-sided topological zero divisors are identi-
cal, and, in this case, we shall speak only of topological zero divi-
sors.
As our first example of topological zero divisors lets us con-
sider the commutative Banach algebra C([O,l]). We claim that
f (C([O,l]) is a topological zero divisor if and only if there
exists some s, 0 < s < I, for which f(s) = O.
Indeed, suppose that f E C([O,l]) and f(s) = 0 for some
s, 0 < s < 1. Then we define the tent functions gk' k = 1,2,3, .. ,
by
gk{t) = k(t - s) + 1 for
1
$ - it ~ t < $,
gk(t) = -k(t - s) + 1
for
gk (t) = 0 for 1 t - s I > ~ .
Clearly each gk is a continuous function on lR, IIg
k
llCl) = 1, and
for k > max[l/s,l/(l - s)] we can consider gk as an element of
C([O,l]) since 0 < s < 1.
Moreover, we claim that limkllfgkllCl) = 0 on the following
grounds: if e > 0 and 6 > 0 are so chosen that If(t)1 < e
whenever It - sl < 6, 0 < t < 1, then for k > max[l/s,I/(1 - s),l/b]
we see that
lfgk(t)1 = 0 for It - 51 > ~
lfgk(t)1 < e for It - sl ~ ~ .
Thus f is a topological zero divisor. If s = 0 or 1, then a
similar argument using the half-tent functions
I
for 0 < t ~ k'
42
1. Fundamentals of Banach Algebras
when s = 0 and
t
gk(t) = 0 for 0 < t < 1 - k'
1
gk(t) = k(t - 1) + 1 for 1 - r< t ~ 1,
when s = 1 yields the same conclusion.
Conversely, suppose f E C([O,l]) and f(s) ~ 0, 0 ~ s < 1.
Since f is continuous, for each s, 0 ~ s ~ 1, there exists some
6
s
> 0 such that If(t)I > If(s)I/2, It - sl < 6
s
' 0 ~ t ~ 1. Obvi-
ously the open intervals (s - 6 ,s + 6 ), 0 < s < I, cover [0,1],
s s --
and hence there exists a finite set of points sl-s2, .. ,sn for
which [ ~ l ] cuP. 1(5. - 6 ,5. + 6 ). Thus, if
J= J Sj J Sj
6 = min If(sj)l
j=1,2, . ,n 2
we see easily that If(t)1 > 6 > 0 for all t, 0 ~ t ~ 1. But then
the function l/f is clearly defined and continuous on [0,1),
that is, f-
l
= l/f E C([O,l]).
Consequently, if (gk) C C([O,l]) is any sequence such that
limkllfgkllCD = 0, then one would have also limkllgklt., = 0 since
lIg
k
ll
CD
= IIf-IfgkllCD < IIf-IIlCDllfgkIlCD' Hence f cannot be a topological
zero divisor.
Furthermore, we note that, if f E C([O,l]) is regular, then
f(s) ~ 0, 0 < s < 1, because if f is regular, then f-
l
E C([O,l])
-1- -
and so (f f)(s) = 1, 0 ~ s < 1, whence f(s) ~ 0, 0 < s < 1. This
observation allows us to conclude that f E C{[O,l]) is a topologi-
cal divisor if and only if f is singular.
Moreover, the previous arguments, with the aid of Urysohn's
Lemma [W
2
, p. 55], can be carried over mutatis mutandis to C(X),
where X is any compact Hausdorff topological space. We summarize
this discussion in the next theorem.
1.6. Topological Zero Divisors
Theorem 1.6.1. Let X be a compact Hausdorff topological
space and suppose f C(X). Then the following are equivalent:
(i) f is a topological zero divisor.
(ii) f is singular.
(iii) There exists some sEX such that f(s) = O.
43
It is well to note that such a simple characterization of topo-
logical zero divisor is not valid in general.
For a second specific example we consider the commutative Banach
algebra ll(f), where convolution is the algebra multiplication.
r is, as it will be throughout the book, the compact Abelian group
under multiplication of complex numbers of absolute value one; that
is, r = {, I , 1'1 = 1) = {e
it
I -u < t < u). The convolution
of two elements f,g E Ll(f) naturally has the following form:
f * g(e
is
) = f(ei(s - t))g(e
it
) dt.
We claim that every element of LI(f) is a topological zero divisor.
S (
it) ikt k 1 2 3 Cl 1 { ) C L (f)
uppose gk e = e , = " ,.... ear y gk I
and IIg
k
il
l
= 1, k = 1,2,3, .... Moreover, for any f E Ll(f),
IIg
k
* fill = eik(s - t)f(e
it
) dtl ds
1
iks .. k
= 2n le
2u
f(e1t)e-
1
t dtl ds

= I f(k) I
(k = 1, 2 , 3, . )
However, from the Riemann-Lebesgue Lemma [E
2
, p. 36], we know that
lim
k
_ +ClJ f (k) I = 0, whence we conclude that f is a topological
zero divisor in LI{f).
44 1. Fundamentals of Banach Algebras
If one wishes to avoid the use of the Riemann-Lebesgue Lemma
in this example, one can instead consider the commutative Banach al-
gebra L
2
(f) in place of LICf). The same argument then shows that
every element of L
2
Cf) is a topological zero divisor. The assertion

that lim
k
_ If(k)1 = 0, f E. L.,Cf), is now a consequence of the
+00 .... kt
complete orthonormality of the set eel I k ] in L
2
Cf). (See,
for example, fL, p. 408].)
As a final, and more abstract, example of topological zero divi-
sors we have the following proposition:
Proposition 1.6.1. Let A be a Banach algebra with identity
e. If x ~ A is topologically nilpotent, then x is a two-sided
topological zero divisor.
Proof. Let (akl be any sequence of distinct nonzero complex
numbers such that limklakl = 0 and consider the sequence x / ~ J
contained in A. Since x is topologically nilpotent, it follews
at once that
~
l / n
limll ~ )nlli/n = Ii x = 0
n a
k
n a
k
Ck = I. 2 , 3 , .. ) ,
and so, by Theorem 1.4.2. we deduce that each x/a
k
is quasi-regular.
Thus, by Proposition I.I.2(i), e - x/a
k
= Cake - x)/a
k
is regular,
-1
whence ake - x is regular, k = 1,2.3, .... Let Yk = Cake - x) ,
k = 1,2,3, .
We note next that
xYk
=
akYk
(ake
- x)y
k
=
akYk
- e
=
akYk
- YkCake
- x)
=
Yk
x
(k = 1,2,3, ... ).
Furthermore,
since x is topologically nilpotent, it is singular.
1.6. Topological Zero Divisors 45
by Proposition 1.4.1, and so it is easily seen that xYk = Ykx is
also k = .... But this implies that !Ia
k
y
k
ll > I,
k = 1,2,3,.... Indeed, if lIakykll < 1, then, by Corollary 1.4.2,
akYk is quasi-regular, whence xY
k
= Ykx = a
k
Y
k
- e is regular,
contrary to the previous observation. Hence "akyk" > 1, for all
positive integers k, from which it follows at once that
limkllYkll = GI, since limklakl = O.
Consequently the estimates
- ell
lIy
k
ll
(k = 1,2,3, .. )
reveal that
that is, x is a two-sided topological zero divisor.
o
Now let us turn to some general results about topological zero
divisors. We first state a simple proposition whose proof we leave
to the reader.
Proposition 1.6.2. Let A be a Banach algebra.
(i) The set of left (right, two-sided) topological zero divi-
sors in A is closed.
(ii) If A has an identity and x E A is a left (right, two-
sided) topological zero divisor, then x is singular.
The converse of Proposition 1.6.2(ii) is not generally valid.
For example, consider the commutative Banach algebra with identity
A(D) introduced in Example 1.2.3. Evidently the element fez) = z,
z E 0, is singular. However, f is not a topological zero divisor
as can be seen from the following: suppose (gk] C A(D) is such
46
I. Fundamentals of Banach Algebras
that li,\lIgkfllCD = o. Since If(z)l = 1 when lzl = 1, we see at
once that limk(suPlzl = llgk(z)l) = O. whence by the Maximum Modulus
Theorem of complex function theory [A, p. 134] we conclude that
limkllgkllCD = o.
The next theorem will provide us with some necessary and suffi-
cient conditions for an element to be either a left or right topolo-
gical zero divisor. We need one definition and some notation before
we can state the result.
Definition 1.6.2. Let A be a normed algebra and x E A. The
left (right) modulus of integrity of x is defined by
~ (x) = inf IIxyll ell (x) = inf ~ ) .
y ~ 0 llYlr y ~ 0 llYlr
Given a Banach algebra A, for each x E A we shall denote by
T and T
X
the elements of L(A) defined by T (y) = xy and
x x
~ y ) = yx, yEA.
Theorem 1.6.2. Let A be a Banach algebra and x ~ A. Then
the following are equivalent:
(i) x is not a left (right) topological zero divisor.
(ii) ~ x ) > 0 ~ x ) > 0).
(iii) There exists some constant K > 0 such that lIxyll > KlIYll
cllyxll > Kllyll) , yEA.
(iv) x is not a left (right) zero divisor and {xy lyE A)
({yx lyE Al) is a closed right (left) ideal in A.
(v) T ~ ) has a continuous inverse when considered as a
x
continuous linear transformation from A to T (A) (Tx(A.
x
Proof. It is apparent from a standard result concerning con-
tinuous linear transformations [L, p. 65] that parts (iii) and (v)
of the theorem are equivalent, and the equivalence of parts (ii) and
1.6. Topological Zero Divisors 47
(iii) follows at once from the definition of the moduli of integrity.
If x were a left topological zero divisor, then there would
exist some sequence (Yk} C A such that IIYkll = 1, k = 1,2,3, ... ,
and limkllxYkll = o. Clearly, in this case, we would have ~ x ) = 0,
and so part (ii) implies part (i). The converse assertion is equally
easy and is left to the reader.
It is obvious that, if part (iii) holds, then x is not a left
(right) zero divisor. Moreover, for instance, the set (xy lyE A)
is clearly a right ideal in A.
converges to z (A. Then, since
Suppose (XYk) is a sequence that
IIx (y,. y. ) II > KIIY
k
- YII,
~ J - J
k,j = 1,2,3, .. , we deduce that (Yk) is a Cauchy sequence in A,
and hence it converges to some w E A. It is then apparent that
xw = z E (xy lyE Al, and so (xy lyE A) is a closed right
ideal. Thus part (iii) implies part (iv).
Finally, suppose x is not a left zero divisor and (xy lyE A)
= Tx(A) is a closed right ideal. Then we note first that Tx is
injective, since Tx(Yl) = T
x
(Y2) implies x(Yl - Y2) = 0 implies
Yl = Y
2
' as x is not a left zero divisor. Thus Tx is an lnJec-
tive continuous linear transformation of A onto the Banach space
T (A). Hence, by a consequence of the Open Mapping Theorem
x
[L, p. 187], we conclude that T has a continuous inverse on its
x
range. Consequently, part (iv) implies part (v).
Therefore parts (i) through (v) are equivalent.
Note, in particular, that the theorem asserts that
a left (right) topological zero divisor if and only if
(l1(x) = 0).
o
x E A is
~ x ) = 0
The second portion of Proposition 1.6.2 asserts that topologi-
cal zero divisors in Banach algebras with identity are always singu-
lar. The converse of this result, as we noted, need not be valid.
However, if A is a commutative Banach algebra with identity and
48 1. Fundamentals of Banach Algebras
x is singular in every superalgebra of A, then it is a topologi-
cal zero divisor. This is the important part of the next theorem.
We first need to define "superalgebra" precisely.
Definition 1.6.3. Let A be a Banach algebra with identity.
A Banach algebra B with identity is said to be a superalgebra of
A if there exists an isometric algebra isomorphism of A into B.
Theorem 1.6.3 (Arens). Let A be a commutative Banach algebra
with identity e and let x A. Then the following are equivalent:
(i) x is a topological zero divisor in A.
(ii) x is singular in every superalgebra B of A.
Proof. If x is a topological zero divisor in A, then x
is clearly a two-sided topological zero divisor in every superalge-
bra B of A, whence, by Proposition 1.6.2(ii), x is singular
in B. Thus part (i) implies part (ii).
Conversely, suppose x is not a topological zero divisor in
A. Then we shall construct a superalgebra B of A in which x
is regular. First, we note, by Theorem 1.6.2, that since x is
not a topological zero divisor, we have > O. Let p >
and consider the commutative algebra Bl consisting of all formal
power series in t, yet) = Ik=oYkt
k
, Yk E A, k = 0,1,2, . , such
that lIy(t)1I = is finite. :xample, if y (A is
such that IIYII < lip, then yet) = t belongs to B
l
, where,
of course, yO = e. The algebra operations in Bl are the usual
formal operations of addition, multiplication, and scalar multipli-
cation applied to power series. Moreover, it is not difficult to
that 1:11 defined above is a norm on 8
1
under which BI
is a commutative normed algebra. By Theorem l.l.l(i), the comple-
tion of B
1
, denoted is a commutative Banach algebra.
Let I be the closed ideal in generated by the element
e - xt; that is, I is the closure in of the ideal
1.6. Topological Zero Divisors 49
{(e - xt)w I w E Then 8 is defined to be the quotient algebra
B = with the usual quotient norm lIIw + IIlI = infvEIllw + vU.
By Theorem l.l.l{iii), B is a commutative Banach algebra. We
claim that B is a superalgebra of A.
Indeed, it is evident that the mapping A - B, defined by
CD k
q:(z) = 1: <P{z)kt + I (z E A),
k=O
where = z, = 0, k = 1,2,3, ... , is a homomorphism of
A into B. Furthermore, given z E A and (e - xt)y(t) I,
where yet) = Lk=oykt
k
, we see that
CD k CD k
II E <P(z)k
t
+ (e - xt)y(t)1l = liz + YO + 1: (Yk - XYk_l)t 11
k=O k=l
= IIzll + - IJlIy(t)1l
> IIzlI
The final inequality is valid since p > and the penulti-
mate inequality utilizes the fact that IIxyli y A. It
is then apparent from the previous inequality and the fact that
{(e - xt)y(t) I yet) E Bil is dense in I that
CD k
IIlcp(z) III = III 1: + I III > IIzlI (z E A).
k=O
The inequality in the opposite direction is trivial, so we conclude
so 1. Fundamentals of Banach Algebras
that 11Iep(z) III = IIzll, z E A; that is, cp is an isometric algebra
isomorphism of A into B. Furthermore, an elementary argument
reveals that epee) = e + I is an identity for B. Thus B is a
commutative superalgebra of A.
Finally, we claim that x is regular in B; that is,
cp(x) = x + I is regular in B.
(x + I)(et + I) = xt + I = e + I,
Indeed, since e - xt E I, we have
that is, (x + 1)-1 = et + I.
Therefore part (ii) sf the theorem implies part e i ~ and the
proof is complete.
o
Returning again to Proposition 1.6.2(ii), we can rephrase the
result there to say that in a Banach algebra with identity e, if
e - x is a left (right, two-sided) topological zero divisor, then
x is quasi-singular. A partial converse of this observation is
contained in the next proposition.
Proposition 1.6.3. Let A be a Banach algebra with identity
e. If x E A is the limit of a sequence of quasi-regular elements
in A, then either x is quasi-regular or e - x is a two-sided
topological zero divisor.
Proof. Suppose {xkJ C A_I is such that limkllxk - xII = o.
If x is quasi-singular, we must show that e - x is a two-sided
topological zero divisor.
We claim first that limkllexk)_lll = CD. Indeed, suppose the
sequence {II Cx
k
)_lIlJ is bounded. Now x
k
0 eXk)_l = 0 implies
that (xk)-l = -x
k
+ x
k
(x
k
)_I' whence
Ck = 1,2,3, ... ).
1.6. Topological Zero Divisors
51
Since limkUx
k
- xU = 0 and (lie - (xk)_IIlJ is bounded. it follows
that limkllYkll = O. In particular, there exists some ko such that,
for k::: ko. IIYkll < 1, and so, by C.,rollary 1.4.2, Yk is quasi-
regular for k > k. Thus, by Proposition 1.1.2{i), e - Yk is
- 0
regular, k > ko. However,
Consequently, since e - Y
k
and e - {xk)_l are regular for k > k ,
-1 - 0
and A is a group, we deduce that e - x is regular, that is,
x is quasi-regular, contrary to the hypothesis that x is quasi-
singular. Hence {1I{x
k
)_IIlJ is unbounded. The same argument,
mutatis mutandis, shows that no subsequence of (II {Xk)_ll!l can be
bounded. and so we conclude that limkUCxkJ_III = CD.
Thus we see that for each k = 1,2,3, ...
liCe - x)(xk)_lll
U{xkJ_llI
=
IICXk)_1 - x{xkJ_Ill
IICx
k
J_
I
II
IIYk - xII
= ~ ~ ~ '
II {xk)_lll
=
lIex - xk)[e - (xk)_l] - xII
II (xkl_lll
from which it follows at once that limkl\Ce - xJexk)_lIl/IlCxk)_11l = o.
Similarly limkl\cxkJ_ICe - xlll/llCxkJ_11I = o.
Therefore e - x is a two-sided topological zero divisor.
[j
The proposition has the following simple corollary. We denote
the topological boundary of a set E by bdy(E).
S2 1. Fundamentals of Banach Algebras
Corollary 1.6.1. Let A be a Banach algebra with identity e.
(i) If x (bdy(A_
I
), then e - x is a two-sided topological
zero divisor.
(ii) If x (bdy(A-
1
), then x is a two-sided topological
zero divisor.
Proof. Part (i) follows immediately from Proposition 1.6.3,
and part (ii) is apparent on noting that x E bdy(A-
l
) if and only
if e - x E bdy(A_
l
) 0
...
The converse of part (ii) may fail; that is, there exist Banach
algebras A with identity which have topological zero divisors not
in bdY(A-
1
).
We shall return to the notion of a topological zero divisor at
various points in the succeeding chapters.
CHAPTER 2
SPECTRA
2.0. Introduction. This chapter is devoted to introducing the
concept of the spectrum of an element of a Banach algebra and to
proving various results connected with this concept. The concept
will be seen to be precisely the extension to the context of Banach
algebras of the notion of the spectrunl of a continuous linear trans-
formation on a Hilbert space, and the reader acquainted with the
Hilbert space theory will find many of the following results and
proofs familiar. As with many of the topics in the preceding chapter,
the contents of the following sections will appear repeatedly in the
sequel.
We begin with the definition of spectrum and then prove some
fundamental theorems, the most important of these being that the
spectrum of any element of a Banach algebra is a nonempty compact
subset of C. The third section cor.tains proofs of the Polynomial
Spectral Mapping Theorem and the Spectral Radius Formula. The for-
mer theorem asserts that polynomials map spectra onto spectra,
whereas the latter provides us with a formula for computing
limnllxnlli/n in terms of the spectrum of x. The final section
discusses the relationship between the spectra of an element when
computed in different algebras.
Once again the reader should observe the role played by the
theory of functions of a complex variable in the study of Banach
algebras.
2.1. Definitions and Basic Results. We have already noted
that, if V is a Hilbert space over ~ then Lev), the space of
S3
54 2. Spectra
continuous linear transformations from V to itself, is a Banach
algebra with identity. In studying such transformations an important
role is played by the notion of the spectrum of an element T E LeV),
that is, by the set oCT) of all C E for which T - ,I is
singular. Here, of course, I denotes the identity transformation
on V. It is evident that this definition of spectrum can be carried
over verbatim to the context of any Banach algebra with identity.
However, we also wish to define the spectrum of an element in a Banach
algebra without identity. The motivation for the definition of the
spectrum in this case comes from the fact that in a Banach algebra
with identity e, x - Ce, , # 0, is singular if and only if xl'
is quasi-singular.
Definition 2.1.1. Let A be a Banach algebra and let x E A.
If A has an identity e, then the sEectrum of x, denoted by a(x) ,
is the set of all
E ~
such that x - Ce is singular; if A is
without identity, then a(x) is the set of all C E ~ , , ~ 0, such
that
xl'
is quasi-singUlar, together with
, =
o.
Note that, if A is a Banach algebra without identity, then
o E o(x), x E A. If A has an identity, then 0 E o(x) if and
only if x is singular.
In discussing the spectrum of an element x 1n a Banach algebra
A it is often important to emphasize that the spectrum is being
computed with respect to a particular algebra. When this is the case,
we shall write o(x) = 0A(x) to highlight this point. Such a dis-
tinction is important, for example, in the next theorem.
Theorem 2.1.1. Let A be a Banach algebra without identity.
If x E A, then 0A(x) = 0A[e] (x).
Proof.
regular in
such that
sible.
We note first that 0 E 0A[e] (x),
A[e], then there would exist some
(x,O)(y,a) = (xy + ax,O) = (0,1),
because if x were
yEA and a E
which is clearly impos-
2.1. Definitions and Basic Results ss
Now suppose ,E aA(x). If ,= 0, then, from the preceding
paragraph, we see that ,E aA[e] (x). If , ~ o . then xl' is
quasi-singular in A and hence in A[e]. The latter follows on
observing that, if xl' were quasi-regular in A[e], then there
would exist yEA and a ~ for which
x xy ax )
= (, + y - c - C,a
= (0,0).
Consequently a = 0 and xl' 0 y = o. Similarly y 0 xl' = 0, and
so xl' is quasi-regular in A. a contradiction. Thus xl' is
quasi-singular in Are], whence x - 'e is singular in A[e].
Hence aA(x) C aA[e] (x).
Conversely, the preceding argument shows that, if , ~ O and
x - 'e is singular in A[e], then xl' is quasi-singular in A,
and 0 E aA[e] (x) from the first paragraph of the proof.
Therefore aA(x) = aA[e] (x).
o
Thus we see that in a Banach algebra A without identity we
may compute a(x) with respect to A or A[e], whichever is most
convenient, and obtain the same result.
Note that the theorem fails if A has an identity. Indeed,
as seen above, 0 E aA[e] (x), x E A, whereas 0 l aA(x) whenever
x E A is regular. The most that can be said in this case is that
aA(x) c aA[e] (x).
Suppose X is a compact Hausdorff topological space and f E C(X).
Then, by Theorem 1.6.1. ,E a(f) if and only if there exists some
t E X such that f(t) = ,. Thus a{f) is precisely R{f), the
range of f. If X is a locally compact noncompact Hausdorff topo-
logical space, then it is easily verified that a(f) for fEe (X)
o
56 2. Spectra
is just R(f) U (O}. We shall see in Section 3.4 that an analog of
these observations is valid for any commutative Banach algebra.
If V is a Hilbert space over ~ and T E L(V), then a funda-
mental theorem from the study of Hilbert spaces asserts that aCT) ~ ~ .
This result is also valid for arbitrary Banach algebras.
Theorem 2.1.2. Let A be a Banach algebra. If x E A, then
a(x) is a nonempty compact subset of {C, e E \l;, Ie' < IIxli}.
Proof. Since, by Theorem 2.1.1, aA(x) = aA(e] (x) when A
is without identity, we may assume, without loss of generality, that
A has an identity e. If a(x) = ~ then x - Ce is regular for
each C E ~ . Then, as in the proof of the Gel'fand-Mazur Theorem
(Theorem 1.5.1), let x* be a continuous linear functional on A
such that x*(x-
l
) = I and consider the function g: ~ - ~ defined
(
-1
by g(C) = x* (x - Ce) ], e E ~ . Precisely the same arguments as
used before show that g is a bounded entire function such that
lim'C,_eog(C) = O. Consequently, again applying Liouville's Theorem
(A, p. 122], we deduce that gee) = 0, C E ~ contradicting the fact
that g(O) = 1. Thus a(x) ~ ~ .
Furthermore, if e EIC is such that 'el > IIxll, then IIx/CII<l,
whence, by Corollary 1.4.2, x/e is quasi-regular. Hence
a(x) c: (e , , E C, 'e' < IIxll). Finally, since by Corollary 1.4.3
the set of regular elements in a Banach algebra with identity is
open, we see at once that a(x) is a closed subset of
{e ICE \C, Ie' < IIxllL and so a(x) is compact.
o
It should be apparent to the reader that we could have proved
this result before discussing the Gel'fand-Mazur Theorem (Theorem
1.5.1). If we had done this, the proof of the Gel'fand-Mazur Theorem
could have been considerably shortened. Indeed, suppose A is a
Banach algebra with identity e which is a division algebra. If
x E A, then from Theorem 2.1.2 we see that a(x) ~ ~ ; that is,
there exists some ,E C for which x - Ce is singular. Hence
2.2. Polynomial Spectral Mapping Theorem 57
x = Ce, as in a division algebra every nonzero element is regular.
The mapping Ce -, then provides us with an isometric algebra iso-
morphism of A onto ~
2.2. The Polynomial Spectral Mapping Theorem and the Spectral
Radius Formula. In Theorem 1.4.4 we saw that, if x is an element
of a/Banach algebra, then lim IIxnU
l
/
n
exists and is no larger than
n
!lxll. Our main goal in this section is to prove that lim IIxnll
l
/
n
n
is precisely the supremum of the absolute values of the numbers in
a(x). This result is generally called the Spectral Radius Formula.
In order to establish this result we first need to prove another
theorem, which is of considerable interest in itself. Note that, if
p(C) = Ik=Oakc
k
, a
k
(C, k = O,I,2, ,n, defines a polynomial on
C, then p(x) = = O a k x k is an element of the algebra A whenever
x E A.
Theorem 2.2.1 (Polynomial Spectral Mapping Theorem). Let A
be a Banach algebra with identity e and let p be a polynomial
on C. If x (A, then a(p(x)) = p[a(x)]; that is, w E a(p(x))
if and only if there exists some C E aCx) such that p(C) = w.
Proof. Suppose ,E a(x) and set q(t) = pet) - pCC), t (C.
Clearly. if P is a polynomial of degree
polynomial of degree n, and q(C) = O.
the other n - I roots of q and write
n. then q is also a
Let '1"2"",C
n
-
1
be
q(t) = aCt - C)(t - , )"'Ct - C )
1 n-l
(t E C),
where a E C is suitably chosen. It is then evident that
q(x) = a(x - Ce)(x - C e)(x -, e).
I n-l
However, since x - 'e is singular, it follows at once that
q(x) = p(x) - pCC)e is also singular. Thus pCC) E a(p(x)).
Conversely, suppose w E a(p(x)) and set q(t) = pet) - w;
58
2. Spectra
t ~ C. Then, factoring the polynomial q, we see that there exist
'O"l' .. "n-l in C and some a E C such that
q (t) = aCt - , ) (t - , ) .. (t -, )
o I n-l
(t E CC),
from which it follows that
q(x) = p(x) - we = a(x - , e) (x - , e)(x -, Ie).
o 1 n-
But since q(x) is singular, we conclude that x - eke must be
singular for at least one k, k = 0,1,2, . ,n - 1. Denote anyone
such 'k by C. Then, clearly, ,E a(x) and pee) = w.
Therefore a(p(x = p[a(x)].
o
Before proving the next result we require one further definition.
Definition 2.2.1. Let A be a Banach algebra. If x E A,
then we set lIxlia = sup, E a(x) 'e', II
x
li
a
being called the spectral
radius of x.
Note that IIxlla ~ lIxU since a(x) c {, , , E C, ", < Uxlll,
by Theorem 2.1.2.
Theorem 2.2.2 (Spectral Radius Formula). Let A be a Banach
algebra. If x E A, then IIxlla = 1im
n
UX
n
U
l
/
n
.
Proof. Since for algebras A without identity we have aA(x)
= aA[e] (x), we may once again assume, without loss of generality,
that A has an identity e. Furthermore, we may assume that x ~ o.
Appealing to Theorems 2.2.1 and 2.1.2, we see that, if ,E q(x) ,
then ,n (a(x
n
) and len, = ~ I n < IIxnU, n = 1,2,3, .. Hence
I ~ < IIx
n
U
l
/
n
, n = 1,2,3, , whence I ~ < limnUxnlll/n. Conse-
quently IIxlia < limnllxnUl/n.
On the other hand, suppose ,E C is such that 0 < I ~ < l/llxll
a
.
Then 1/' ~ q(x), and so x - Ce is regular. Thus, using the same
arguments mutatis mutandis as used in the proof of the Gel'fand-Mazur
2.2. Polynomial Spectral Mapping Theorem
59
Theorem (Theorem 1.5.1) and Theorem 2.1.2, we see that
-1
gx*C,) = x*[Cex - e) ]
defines an analytic function on {, leE ;, I, I < l/lIxliol for any
continuous linear functional x* on A. Moreover, some direct
computations reveal that, if 1'1 < 1/lIxli < l/Uxll, then
- 0
1 CD k CD k k
(ex - e)- = - E (ex) = - E ex.
k=O k=O
The series converges in A since ll,xll < 1. Thus
-1
g .C') = x*[c,x - e) ]
x
CD k k
= - E x*(, x )
k=O
CD k k
= - E x*(x )'
k=O
(,,1 < l/lIxll)
But since g
x*
is analytic in (e' e E C, ,,' < l/lIxliol and
l/lIxll < 1/lIxllo'
sis [A, p. 177],
we deduce, via a classical theorem of complex analy-
that g * is represented in all of
x
by the power series indicated above.
In particular, for each " 'el < l/llxllaJ we can conclude that
limkx*cekxk) = 0 for each continuous linear functional x on Ai
that is, for each " "I < l/llxlla' the sequence (ekxk} converges
to zero in the weak topology [L, p. 239] on A. Consequently, for
each e, I" < l/llxll, there exists some constant K, > 0 such
k. k 0
that 'e' fix II <K" k = 0,1,2, ... , as weakly convergent sequences
are bounded [L, p. 246].
Hence we see that for each e, 'el < l/llxll
o
'
II krl/k (Ke>l/k
x I ( , , (k = 0, 1 , 2 , ) ,
60
2. Spectra
k 11k
whence we have I imkllx II < 111, I,
since limk(K,)l/k = 1. But
, such that 0 < I' I < l/llxlla'
the last inequality holds for every
that is, such that 1/1,1 > IIxlla.
Therefore we conclude that limkllxkUl/k IIxlla' which completes
the proof.
An examination of the proof reveals that, if one argues using
lim inf and lim sup in place of lim in the first and second portions
of the proof, respectively, then the proof shows not only that
lim IIx
n
ll
l/n
= IIxli but also that lim IIxnU
l/n
exists.
nan
If X is a compact Hausdorff topological space, then it is
apparent that Ufll = lim IIf
n
U
l/n
= UfU , f E C(X). But for arbi-
CD n CD a
trary Banach algebras A it is not generally the case that
IIxli = IIxlla' x A. However, in Section 3.4 we shall obtain another
expression for II xli in commutative Banach algebras which is an
o
appropriate counterpart of the situation in C(X).
An obvious consequence of Theorem 2.2.2 is the next corollary.
Corollary 2.2.1. Let B be a Banach algebra and let A be a
closed subalgebra of B. If x E A, then
Proof.
o
2.3. A Theorem of on Spectra. Suppose that B is a
Banach algebra with identity e and A is a closed subalgebra of
B that contains e. Given an x E A, what can be said about the
relationship between aA(x) and aB(x)? One such result was given
at the close of the preceding section. The theorem indicated in the
heading of this section, together with its con-
siderable additional information on this relationship. The proof
2.3. A Theorem of
61
of the theorem will utilize our knowledge of topological zero divi-
sors; in particular, we shall need Corollary 1.6.1 and Proposition
1.6.2.
Theorem 2.3.1 (ilov). Let B be a Banach algebra with iden-
tity e and let A be a closed subalgebra of B that contains e.
If x E A, then oBex) c 0A(x) and bdy[oA(x)] c bdy[aB(x)].
Proof. Clearly, if x - Ce is singular in B, then it is
singular in A, and so 0B(x) c aA(x).
On the other hand, suppose' E bdy[aA(x)]. Then there exists
a sequence {C
k
) C C such that =, and x - eke is regular,
k = 1,2,3, . Thus limkllex - Ce) - ex - Cke)1I :: 0, whence x - Ce
is a limit of a sequence of elements that are regular in A. However,
since aA(x) is closed, bdy[aA(x)] c aACx), and so x - 'e is
-1
singular. Hence x - 'e (bdy(A ). Consequently from Corollary
1.6.1 we see that x - 'e is a two-sided topological zero divisor
in A and hence in B. Thus, by Proposition 1.6.2(ii), x - 'e
is singular in B; that is, ,E 0B(x).
But the sequence (x - eke) c B-
1
, and
liml'k - ,I = 1imllCx - 'e) - (x - 'ke)1I = o.
k k
Therefore '( bdY[OB(x)], and so bdy[aA(x)] c bdy[aB(x)]. 0
Theorem 2.1.1 allows us to partially extend this result to alge-
bras without identity.
Corollary 2.3.1. Let B be a Banach algebra without identity
and let A be a closed subalgebra of B without identity. If
x (A, then aB(x) c aA(x).
Proof. It is easily verified that A[e] is a closed subalgebra
of B[e]. Since 0A(x) = 0A[e]Cx) and aB(x) = aB[e] (x) by Theorem
2.1.1, we obtain the desired conclusion by applying Theorem 2.3.1. 0
62
2. Spectra
Note that the situation when B is without identity but A
has an identity seems more complicated. The complication arises
since, in this case, given x E A, 0 E 0B[e] (x), but 0 E 0A(x)
and only if x is singular in A.
if
Corollary 2.3.2. Let B be a Banach algebra with identity e,
let A be a closed subalgebra that contains e, and suppose x E A.
Ci)
If
C - 0B(x)
is connected, then
0Aex) = oBex).
(ii) If int(oA(x)] - ~ - ,
then
0A(x) = 0B(x).
(iii) If
~ ~ C ~
then
0A(x) = 0B(x).
Proof. From Theorem 2.3.1 we see that
and that the union is disjoint. Clearly ~ - 0Aex) is open and
nonempty, as 0A(x) is compact, by Theorem 2.1.2. Moreover, we
claim that 0A(x) - 0B(x) is open. Indeed, if ,E 0A(x) - B(X) ,
then ,E int[oA(x)], as otherwise ,E 0A(x) - int[oA(x)] =
bdy[oA(x)] C bdy[oB(x)] C B(X) , by Theorem 2.3.1 and the fact that
0B(x) is closed. But then, if U is an open neighborhood of ,
such that U C int[oA(x)], it follows at once that
is an open neighborhood of , such that We 0A(x) - 0B(x). Thus
0A(x) - 0B(x) is an open subset of the open connected set ~ - B(X) ,
and we conclude that 0A(x) - 0B(x) = ~ , which proves part (i).
Part (ii) of the corollary is immediate on noting that
0B(x) C 0A(x) = bdy[oA(x)] C bdy[oB(x)] C 0B(x), as int[oA(x)]
and part (iii) follows at once from part (ii).
= ~ ,
o
Corollary 2.3.3. Let B be a Banach algebra with identity e
and let x E B be such that 0B(x) c ~
2.3. A Theorem of ilov 63
(i) If A is any closed subalgebra of B that contains x
and e, then aA(x) = aB(x).
(ii) If A is any superalgebra of B such that e is the
identity in A, then aA(x) = aB(x).
Proof. Part (i) follows from Corollary 2.3.2(i) on noting that
- aB(x) is connected, and part (ii) follows from Corollary 2.3.2(iii)O
In closing, it is perhaps worth mentioning that a translation of
the results in this chapter for the specific Banach algebras of con-
tinuous linear transformations on a Banach or a Hilbert space V,
that is, for A = L(V), immediately yields a number of the funda-
mental results in the spectral theory of such transformations. This
is true, for example, of Theorems 2.1.2,2.2.1, and 2.2.2.
CHAPTER 3
THE GEL'FAND THEORY
3.0. Introduction. Beginning with this chapter, we shall
concentrate our attention almost entirely on commutative Banach alge-
bras. This restriction is imposed because the main tool we shall
utilize in the further study of Banach algebras is the Gel'fand
representation theory, which is valid only for commutative algebras.
The reason for this is not difficult to understand.
As will be seen shortly, the Gel'fand representation theory is
based on the observation that the quotient algebra of a commutative
Banach algebra modulo a maximal regular ideal is a division algebra,
and hence, from the Gel'fand-Mazur Theorem (Theorem 1.5.1), the
quotient algebra is isometrically isomorphic to Thus the maxi-
mal regular ideals in a commutative Banach algebra A determine
homomorphisms of A onto and using these homomorphisms we can
construct an algebra of continuous functions on a certain locally
compact Hausdorff topological space which is a continuous homomor-
phic image of A. Such a representation of a commutative Banach
algebra will be seen to be very useful in the investigation of these
algebras. The development fails for noncommutative algebras because
the quotient of such an algebra by a maximal regular two-sided ideal
need not be a division algebra.
For instance, consider the noncommutative algebra A with iden-
tity consisting of all 2 x 2 matrices with complex entries. Theft
it is not difficult to verify that the only two-sided ideals I in
A are either I = ([g gJl or I = A. Consequently the only maxi-
mal two-sided ideal is I = {[g gJ), and A/I = A, which is clearly
not a division algebra.
64
3.1. Maximal Regular Ideals
6S
For this reason in the following pages we shall deal substan-
tively only with commutative algebras.
Proceeding as indicated above, we shall develop the Gel'fand
representation theory, which asserts that, given a commutative Banach
algebra A, there exists a norm-decreasing homomorphism of A onto
a subalgebra of C ~ A ) ) , where ~ A ) is a certain locally compact
o
Hausdorff topological space associated with A. The points in this
space A(A) will be the maximal regular ideals in A, or, equival-
ently, the homomorphisms of A onto C. This development will be
carried out in the next three sections.
Afterward we shall briefly discuss some connections between the
Gel'fand representation theory and spectra, and the question of when
the Gel'fand representation of A is an isomorphic image of A. In
particular we shall show that A is isomorphic to a subalgebra of
C o ~ A ) ) precisely when the only topologically nilpotent element
of A is the zero element.
3.1. Maximal Regular Ideals and Complex Homomerphisms. We
begin this section by proving the algebraic result alluded to in
the introduction.
Theorem 3.1.1. Let A be a commutative algebra. If MeA
is a maximal regular ideal, then A/M is a division algebra.
Proof. Let u ~ A be an identity modulo M. By Proposition
1.1.1, the element u + M is an identity for the commutative algebra
A/M. To prove the theorem it clearly suffices to show that, if
x E A, x l N, then x + N is regular in A/M.
Let J = {y + zx I y ~ N, z ~ A). Then J is obviously a
regular ideal in A that contains M. Moreover, J ~ M. Indeed,
since u is an identity modulo N, we have ux - x E M, whence
x = -(ux - x) + ux E J. Thus J ~ M, as x l M. Consequently,
since M is a maximal regular ideal, we conclude that J = A.-
66
3. The Gel'fand Representation Theory
In particular, then, there exist y M and z A such that
y + zx = u, from which it follows at once that (z + M)(x + M) =
u + M; that is, z + M is an inverse for x + M.
o
Now suppose that A is a commutative Banach algebra. If
MeA is a maximal regular ideal, then from Theorems 1.1.1 and
3.1.1 we see that AIM is a Banach division algebra and hence iso-
metrically isomorphic to the complex numbers ~ by the Gel'fand-
Mazur Theorem (Theorem 1.5.1). If a denotes this isomorphism of
AIM onto C and ~ denotes the canonical homomorphism of A onto
AIM, then 1" = a 0 ~ is clearly an algebra homomorphism of A onto
C such that ,.-1(0) = {x I x E A, 1"(x) = oj = M. Since 1" is linear
and M is a closed linear subspace of the Banach space A, by
Corollary 1.4.4, it follows from a standard theorem of functional
analysis [L, p. 69] that 1" is continuous. Thus 1" is a continu-
ous linear functional on A that is not identically zero and such
that T(XY) = T(X)T(y), x,y E A. Moreover, since ~ is norm de-
creasing and a is an isometry, we see at once that
the case that A has an identity e and lIell = 1,
assume without loss of generality, T(e) = 1"(e)1"(e)
T (e) = I, and so 111"11 = 1.
111"11 < 1. In
which we may
implies that
So far we have seen that each maximal regular ideal M in a
commutative Banach algebra A determines a continuous homomorphism
of A onto C with norm at most one. Furthermore, the homomorphism
so determined is unique. Suppose 1" and ware homomorphisms of
-1 -1
A onto C such that T (0) = w (0) = M. Now M, being the
kernel of a nonzero linear functional on A, is a maximal linear
subspace [L, p. 68]. From this observation it follows easily that
there exists some b ~ b ~ 0, for which w = b1". We claim that
b = 1. Indeed, if x E A is such that 1"(x) ~ 0, then
bT(X)2 = b1"(xx) = w(xx) = wex)w(x) = b
2
1"ex)2,
h b2 = b d b I Th d h . I
w ence , an so =. us 1" = w, an eac maXIma
regular ideal M determines a unique homomorphism.
3.1. Maximal Regular Ideals 67
We wish to give the homomorphisas T a special n8JIe. As the
abstract notion of such a homomorphism obviously makes sense even
for noncommutative algebras. we frame the next definition in this
context.
Definition 3.1.1. Let A be a nOrlDed algebra. A homollorphiSll

T of A onto C is said to be a cOMplex homomorphism.
Complex homomorphisms are also often called multiplicative
linear functionals.
Although a maximal regular twosided ideal in a noncommutative
Banach algebra need not determine a complex homomorphis. of the alge-
bra. as indicated in the introduction, the converse assertion is
valid. Indeed, suppose A is an arbitrary Banach algebra and ,.
is a complex homomorphism of A.
linear functional on A so that
linear subspace of A [L, p. 68].
~ x y ) = T(X)T(Y), x,y E A, that
Then evidently T is a nonzero
M T-I(O) is a proper maximal
It follows at once from
M is a two-sided ideal in A.
Moreover, M is a regular ideal as can be seen fTom the following:
Suppose u E A - M is such that T(U). 1. Then we see that, ~ o r
any x E A,
T(UX - x) T(U)T(X) - T(X) = 0 = T(X)T(U) - T(X) T(XU - x),
whence ux - x E M and xu - x ~ M. Thus u is an identity modulo
M, and M is regular. Consequently, since M is a proper maximal
linear subspace of A and M is a proper regular two-sided ideal
in A, we conclude, by Theorem 1.4.5 and Corollary 1.4.4. that M
is a maximal regular two-sided ideal. In particular, T is continu-
ous, as M = T-I(O) is a closed linear subspace [L, p. 69]. Further-
aore we see once again that iI'fll < 1.
Indeed, suppose x E A, x ~ 0, and assume r(x), is such
that lei> IIxll. Then, by Corollary 1.4.2(i), xl' is quasi-regular,
as lIx1clI < I, and so there exists some yEA such that:
68 3. The Cel'fand Representation Theory
0 y = + y - f = O.
Hence, since T is a homomorphism, we have
x xy
o = T(- + y - -)
, ,
= T(X) + _ T(X)T(y)
, ,
= 1 + T(Y) - T(y)
= 1,
which is an absurdity. Thus J T (x) I < IIxU, x . A, and IITII < 1.
As before .. we see at once that UT/I .. 1 in the case that A
bas an identity.
We can summarize the preceding discussion.. in the case of commu-
tative Banach in the following theorem:
Theorem 3.1.2. Let A be a commutative Banach algebra.
(i) If T is a complex homomorphism of A, then T is con-
tinuous and UTI! < 1. Moreover, if A has an identity e and
lieU = 1 JI then IiTIl = 1.
(ii) If T is a complex homomorphism of A, then M = T-
1
(O)
is a maximal regular ideal in A.
(iii) If MeA is a maximal regular ideal, then there exists
a unique complex homomorphism T of A such that T-1(O) = M.
(iv) The correspondence between the complex homomorphisms of A
and the maximal regular ideals in A determined by parts (ii) and
(iii) is bijective.
3.2. The Maximal Ideal Space. In this section we wish to look
more closely at the collection of maximal regular ideals in a commu-
3.2. The Maximal Ideal Space
69
tative Banach algebra, or, equivalently, at the collection of complex
homomorphisms. First, we shall consider the relationship between
the complex homomorphisms of a commutative Banach algebra A with-
out identity and the complex homomorphisms of A[e]. Using this
knowledie we shall introduce a topology on the collection of complex
homomorphisms -- namely, the weak* topology -- and show that the
topological space so obtained is a locally compact Hausdorff topo-
logical space.
We begin by setting terminology.
Definition 3.2.1. Let A be a commutative Banach nlgebra and
let denote the collection of all the maximal regular ideals
M in Aj A (A) will be called the maximal ideal space of A.
Other common names given to ntA) are the structure space, the
homomo!phism space, or the spectrum of A.
In view of Theorem 3.1.2, we can and will. generally without
explicit cemment, identify the maximal regular ideals in A (A) with
the complex homomorphisms they det.rmine. The choice of whether to
think of the points in as ideals or as complex homomorphisms
usually depends on the question at hand. For example, in this sec-
tion we shall introduce a topology into A (A) where we shall consi-
der the points of A (A) as complex whereas in Section
7.1 we shall introduce a different topology into A(A), described
in terms of ideals.
In the next theorem we shall think of the elements of A (A) as
complex homomorphisms.
Theorem 3.2.1. Let A be a commutative Banach algebra without
identity_
(i) If T E 6(A), then there exists a unique Te ACA[e])
such that (x) = x A.
e
70
3. The Gel'fand Representation Theory
(ii) If Te E and Te(A) 0, then the equation
T(X) = T (x), x A. defines a complex homomorphism T C
e
Proof. If T E then M = T-l(O) is a maximal regular
ideal in A, and so. by Corollary 1.1.3, there exists a unique
maximal ideal M C A[e] such that M A and M = M n A. Let
e e e
Te E be the unique complex homomorphism of A[e] such that
T-l(O) = M. Since, by Theorem 1.1.5, A and M are maximal
e e e
ideals in A[e] and Me A, it follows that Me A. Thus Te
is not identically zero on A, and so the restriction of T to
e
A, call it TelA' defines a complex homomorphism of A such that
(TeIA)-l(o) = M. Since M uniquely determines a complex homomor-
phism of A, by Theorem 3.1.2(iii), we conclude that TelA = T,
which proves part (i) of the theorem.
On the other hand, suppose Te E
again using the fact that A and M = T-l(O)
e e
in A[e], we easily deduce that T = TelA is
of A, thereby proving part (ii).

and T (A) O. Then,
e
are maximal ideals
a complex homomorphism
o
It should be apparent that this theorem is really just Corollary
1.1.3 translated into the terminology of complex homomorphisms.
If w E is such that weAl = 0, then we see at once
that w(x + ae) = aw(e) = a, x E A and a (C. In particular,
there is evidently precisely one such w E the one defined
by the preceding equation. With this in mind we make the next defi-
nition.
Definition 3.2.2. Let A be a commutative Banach algebra with-
out identity. Then T E will denote the unique complex
GO
homomorphism of A[e] that vanishes identically on A.
These remarks, combined with Theorem 3.2.1, reveal that, if A
is a commutative Banach algebra without identity, then
3.2. The Maximal Ideal Space 71
where the T E
such that T (A) o.
e
tarily.
= U {T ],
CD
have been identified with those T E A(A[e])
e
We shall make use of this observation momen-
Next we wish to define a topology on the maximal ideal space
of a commutative Banach algebra A. Considering as
complex homomorphisms of A we see, from Theorem 3.1.2, that
can be identified with a subset of the closed unit ball of A*, the
space of continuous linear functionals on A. This latter set, by
the Banach-Alaoglu Theorem [L, p. 254], is compact in the weak* topo-
logy on A*. This observation leads naturally to the following
definition:
Definition 3.2.3. Let A be a commutative Banach algebra.
The Gel'fand topology on is defined to be the relative weak*
topology on considered as a subset of A*.
In view of the definition of the weak* topology [L, pp. 240
and 241], we see that a neighborhood base at T E consists of
sets of the form
where e > 0, n is a positive integer, and x
l
,x
2
, . ,x
n
in A
are arbitrary.
The main theorem concerning the Gel'fand topology is the follow-
ing result:
Theorem 3.2.2. Let A be a commutative Banach algebra.
(i) If A has an identity, then with the Gel'fand topo-
logy is a compact Hausdorff topological space.
(ii) If A is without identity, A(A) with the Gel'fand
topology is a locally compact Hausdorff topological space,
72 3. The Gel'fand Representation Theory
A(A[e]) with the Gel'fand topology is the one-point compactifica-
tion of A(A).
Proof. Since the weak* topology is Hausdorff, it is evident
that the Gel'fand topology is always Hausdorff. Suppose that A
has an identity e and, as usual, assume that lieU = 1. From
Theorem 3.1.2(i) we see that UTU = 1, T E A(A). Since the closed
unit ball in A* is compact in the topology, to show that
A(A) is compact in the Gel'fand topology it suffices to prove that
A(A) is weak* closed. So suppose (T 1 C A(A) is a net and x* E A*,
a
IIx*1I < 1, are such that (T) converges to x* in the weak* topo-
a
logy on A*; that is, lim T (x) = x*(x), x E A. We must show that
aa
x* is multiplicative and IIx*1I = 1.
To this end let x,y E A, x 0, y 0, and > O. Since
converges weak* to x*, there exists some a
o
a > a , T belongs to the weak* neighborhood
o a
such that,
U(X*;b;x,y,xy)
{y*ly* EA*, lx*(x)-y*(x) I <6,lx*(y)-y*(y) 1<6, lx*(xy)-y*(xy) 1<6),
where o < () < Thus for
Ix*(xy) - x*(x)x*(y)1 < lx*(xy) - T (xy) I
- a
a > a we have
o
+ IT (X)T (y) - T (x)x*(y)I
ex ex a
+ IT (x)x*(y) - x*(x)x*(y)1
a
< Ix*(xy) - T (xy) I
- ex
+ IIxlllT (y) - x*(y)1
a
+ lIylllT (x) - x*(x)l
a
e e
<-+-+ .....
333
=
3.2. The Maximal Ideal Space 73
as liT II = 1 and
a
is arbitrary, that
ment reveals that
IIx*1I < 1. Consequently we conclude, since ,> 0
x*(xy) = x*(x)x*(y), x,y E A. A similar argu-
x*(e) = 1, whence Ux*U = 1. Thus x* E A(A).
Hence A(A) is closed in the weak* topology, and so A(A) is
compact in the Gel'fand topology.
Now assume that A is without identity. By Theorem 3.2.1 and
the remarks following Definition 3.2.2 we can consider A (A) as a
subset of the set A(A[e]) = A ) U {T], and T E A(A) corresponds
CD
to Te E A(A[e]), as determined in Theorem 3.2.1. Let T E A(A)
and consider the neighborhood U(T;e;x
l
,x
2
, ... ,x
n
) of T in the
Gel'fand topology on ~ A ) . If IT(xk)1 < e, k = 1,2, ... ,n, then
= {w I w E A (A) .1 T (x
k
) - W (x
k
) I < e, k = 1,2, .. n] U (T )
e e e CD
whereas if IT(xk)1 ~ e for some k = 1,2, ,n, then
= (w I W E A (A) , IT (X
k
) - W (X
k
) I < e , k = 1,2, , n)
e e e
In either case it is apparent that
Thus open neighborhoods of points of A (A) in the Gel'fand topology
on A(A) are open sets in the relative topology induced on A(A)
by the Gel'fand topology on A(A[e]). In a similar fashion one can
show that open sets in this relative topology on ~ A ) are actually
open in the Gel'fand topology on A(A).
Thus the Gel'fand topology on ~ A ) and the relative topology
on ~ A ) induced by the Gel'fand topology on A(A[e]) coincide.
74
3. The Gel'fand Representation Theory
Since (T) is a closed subset of A(A[e]) =1 A(A) U (T), it follows
m m
at once that A(A) is an open subset of A(A[e]).
In view of the first portion of the theorem it is now apparent
that A(A) is locally compact in the Gel'fand topology, and A(A[e])
is the one-point compactification of A(A).
o
In the future we shall almost always tacitly assume that the
maximal ideal space A(A) of a commutative Banach algebra A is
endowed with the Gel'fand topology. In particular, then, A(A) is
a locally compact Hausdorff topological space.
3.3. The Gel'fand Representation. Only one step now remains
in our program of homomorphically representing a commutative Banach
algebra A an algebra of continuous functions on the locally
compact Hausdorff topological space A(A): we need to define an
appropriate mapping from A to the continuous functions on A(A).
This is accomplished by the next definition.
Definition 3.3.1. Let A be a commutative Banach algebra. If
..
x E A, then x will denote the complex-valued function defined on
A(A) by = = T(X), T E A(A).
Since the Gel'fand topology is the relative weak* topology on
..
A(A), it is immediately apparent that x is a continuous function
..
on A(A). Moreover, it is easily checked that the mapping x - x,
x A, defines a homomorphism of A onto some algebra of continuous
functions on A(A). Further properties of this mapping are contain-
ed in the next theorem.
Theorem 3.3.1 (Gel' fand ReEresentation Theorem). Let A be a
..
commutative Banach algebra. The mapping
x - x, x E A,
defines a
..
homomorphism of A onto a subalgebra A of C (A(A)). Moreover,
..
0
A
separates the points of A(A), and if A has an identity, then
..
A contains the constant functions.
3.3. The Gel'fand Representation 75
"
Proof. We noted above that the mapping x - x is a homomor-
" "
phism of A onto an algebra A = (x I x E A) of continuous complex-
valued functions on 6(A). Furthermore, the estimates
"
= sup 1-r(x)1
1 A(A)
reveal that each x is a bounded continuous function and that the

mapping x - x is norm decreasing.
If A has an identity, then, by Theorem 3.2.2, 6(A) is com-
"
pact and so A C C
o
(6(A)) = C(A(A)). If A is without identity,
, then A(A[e]) = A(A) U (T) is the one-point compactification of
CD
A(A) . Given x A and e > 0, we see that
U(T ,e,x)
=
(T I T E A(A), IT (x) - T (x)1 < e) U (T }
CD
e e CD CD
= (T T E t.(A),
I T (x) I
< ~ U (T )
CD
"
= (-r T A(A), IX(T)I < e) U (1 )
CD
because 1 (x) =
CD
0, x E A. As usual, T E A(A[e])
e
is the complex
T A(A). Thus, if T E 6(A) does homomorphism corresponding to
not belong to the compact set A(A[e]) - UCT ,e,x) c aCA), then
CD
"
IX(T)I < e.
" "
This, however, says precisely that each x E A belongs
"
Hence in either case, A c C (aCA)).
o
"
To see that A separates the points of 6(A) , consider
T,W E aCA) , T ~ w. Then from the remarks preceding Definition 3.1.1
76
3. The Gel'fand Representation Theory
we see that
-1
T-I(O) "I u,-1(0).
-1
Without loss of generality, suppose
A
x E T (0), but x l w (0). Then, obviously, X(T) = T(X) = 0,
A A
whereas x(w) = w(x) "I o. Thus A separates points.
A
Finally, if A has an identity e, then e(T) = 1, T E
A
from which it follows that A contains the constant functions.
o
Given a commutative Banach algebra A, we shall generally
A
speak of A as the Gel'fand representation of A, and of the map-
A
ping x - x, x E A, as the Gel'fand transformation (occasionally
we shall also refer to the mapping as the Gel'fand representation);
A
x itself will be called the Gel'fand transform of x.
The common notation employed for both the Gel'fand transform
and the Fourier transform is not accidental, as will become apparent
in Section 4.7.
The Gel'fand topology is often defined in a different way from
that of the preceding section. This alternative definition is con-
tained in the next result, whose proof is left to the reader.
Corollary 3.3.1. Let A be a commutative Banach algebra. Then
the Gel'fand topology on is the weakest topology on A(A)
A
such that the Gel'fand transform x is continuous on A (A) for all
x E A.
So far in the development of this chapter there is one obvious
and important question that we have not faced: If A is a commuta-
tive Banach algebra, do there exist any complex homomorphisms of A,
that is, is The answer is not necessarily. That is,
there exist commutative Banach algebras A such that the only homo-
morphism of A into is the zero homomorphism. One rather tri-
vial instance of such a phenomenon is given by the following example:
Let (A,IIU) be a Banach space over \C and define a multipli-
cation in A by xy = 0, x,y E A. It is easily verified that with
this multiplication A becomes a commutative Banach algebra without
3.3. The Gel'fand Representation 77
identity. However, 6(A) = because if 1
on A such that T(xy) = 1(x)r(y), x,y E A,
is any linear functional
2
then T(X) = 1(XX) =
1(0) = 0, x E A, from which it is apparent that T = O.
A less trivial example can be obtained as follows: For each
f E C([O,l]) define
T(f)(t) = f(s) ds
(t ( [0,1]).
Then T (L(C([O,l])). Let A be the norm closure in L(C([O,l]))
of the set of all polynomials in T of the form Ik=lakT
k
, where
n is a positive integer, and a
l
,a
2
, ... ,a
n
in are arbitrary.
Then A is a commutative Banach algebra without identity such that
A(A) To prove the latter assertion it suffices to show that
lim IIT
n
ll
l
/
n
= O. This is the case because, by the Beur1ing-Gel
'
fand
n
Theorem (Corollary 3.4.1), which will be proved in the next section,
if lim IIT
n
\ll/n = 0, then IIrll = 0; that is, 1(T) = 0, r ( A(A).
n CIO
But the elements of A are either polynomials in T without constant
term or limits of such polynomials, whence we deduce that 1(S) = 0,
1 (A(A), for each S (A. Hence the only homomorphism of A into
is the zero homomorphism. It remains then only to show that
lim IIT
n
ll
l
/
n
= o.
n
However, if f E C([O,l]), then
IT
2
(f)(t)1 = feu) du] dsl
If(u)l du] ds
< IIfllCD du) ds
= IIfllCD
More generally, an induction argument reveals that
(t E [0,1]).
(t ( [0,1]; n = 1,2,3, ... ).
78 3. The Gel'fand Representation Theory
Consequently
(f C([O,l]); n = 1,2,3, ... ),
from which we conclude at once that
IIT
n
ll
i/n
< (1.. ) lIn
- n!
Cn = 1,2,3, .. ).
The desired conclusion in now apparent on noting that
lim (lIn!) I/n= 0
n
If A is a commutative Banach algebra with identity, the ano-
maly just discussed cannot occur.
Theorem 3.3.2. Let A be a commutative Banach algebra with
identity. Then
Proof. Since {oj is a proper ideal in A, by Theorem 1.1.3
there exists a maximal ideal in A that contains {oj.
o
The reader should recall that our algebras are always assumed
to contain nonzero elements.
With regard to the Gel'fand representation theory, it is obvious
that a commutative Banach algebra A such that A(A) = is of no
interest. Because of this, and in order to simplify some of the
subsequent development, we shall henceforward only consider commuta-
tive Banach algebras A such that Thus the Gel'fand
representation theory will always have content.
The Gel'fand Representation Theorem provides us with a homomor-
phic mapping of a commutative Banach algebra A onto a subalgebra

A of Co(A(A)). Clearly A is then itself a normed algebra, and

so its uniform closure, call it cl(A) is a commutative Banach alge-
bra with the supremum norm. An question to raise is: What

is the relation of cl(A) to its Gel'fand representation cl(A) ,
that is, what happens if we iterate the Gel'fand transformation.
The answer is that we get essentially nothing new because A(A) is
3.4. The Beurling-Gel'fand Theorem
79

homeomorphic to b(cl(A)) and cl(A) is isometrically isomorphic

to cl(A).

Indeed, define the mapping 6(A) - A(cl(A)) by setting

= w
T
' T E 6(A) , where wT(x) = = T(X), x A. It is
easily verified that the preceding definition actually defines a

complex homomorphism w
T
on cl(A), as A is norm dense in cl(A),
. .. . .
and so w
T
E 6(cl (A)) Further we define a mapping : cl (A) - cl (A)
. - . . . . . .
by setting t[(x) ] = (x) (x) E cICA) . Then some elementary
arguments, whose details we leave to the reader, establish the follow-
ing theorem:
Theorem 3.3.3. Let A be a commutative Banach algebra and

let cICA) denote the supremum norm closure of Ace (6(A)). Then
o

(i)
T A(A) ,
The mapping 6(A) - 6(cl(A)), defined by = w
T
'

is a homeomorphism of 6(A) onto 6(cl(A)).

(ii) The mapping : cl(A) - cl(A), defined by ,[(x) ] =

(x) (x) E cl(A) , is an isometric algebra isomorphism of
.. . .
cl(A) onto cl(A).
Thus we see that nothing is to be gained by repeated applica-
tions of the Gel'fand transformation.
3.4. The Beurling-Gel'fand Theorem. In Section 2.2 we proved
the Spectral Radius Formula (Theorem 2.2.2), which asserts that, if
x is an element of a Banach algebra A, then
limllxnll
i
/
n
= sup = II xll
a

n
The Beurling-Gel'fand Theorem asserts that, if A is a commutative
A
Banach algebra, then IIxlia = lIxllm' x E A. This result will be an
immediate corollary of the next theorem. which connects the range
of the Gel'fand transform of x with the spectrum of x. We denote

the range of x by R(x).
80 3. The Gel'fand Representation Theory
Theorem 3.4.1. Let A be a commutative Banach algebra .

(i) If A has an identity, then a(x) = R(x), x E A

(ii) If A is without identity, then a(x) = R(x) u Col, x E A.
Proof. If A has an identity e and x (A, then x - Ce
is singular for each C E a(x). Thus, by Corollary 1.1.4, each
IC = (Cx - Ce)z I z E A), C E a(x), is a proper ideal in A that
contains x - Ce, and so, by Theorem 1.1.3, for each C E a(x)
there exists a maximal ideal MC E A such that M, IC. If
-1
TC E 6(A) is such that TC (0) = M
C
' then

as e(T
C
) = 1. Thus a(x) c R(x) , x E A.
Conversely, if x E A and C a(x), then x - Ce is regular,
whence, by Corollary 1.1.2, x - Ce belongs to no maximal ideal in

A. Hence T(X - Ce) = X(T) - C 0, T (6(A), which shows that

C R(x).

Therefore a(x) = R(x), x E A, when A has an identity.
Now suppose A is without identity and x E A. If C E a(x)
and C 0, then xl' is quasi-singular in A and so also in
A[e]. Thus, by Proposition 1.1.2, e - x/C is singular in A[e].
Arguing as before, we deduce the existence of some w E 6(A[e])
such that wee - x/C) = 1 - w(x/C) = 0, that is, w(x) = C. Hence,
by Theorem 3.2.l(ii), there exists some T ( 6(A) such that

X(T) = T(X) = w(x) = C. Consequently we see that

a(x) C R(x) U {Ole
The proof of the reverse containment is left to the reader. G
Some useful corollaries of this theorem are given below. The
details are left to the reader.
3.S. Semisimplicity
Corollary 3.4.1 (Beurling-Gel'fand Theorem). Let A be a
commutative Banach algebra. Then
81
n lIn
limllx II = IIxll = IIxli
a CID
(x E A).
n
Corollary 3.4.2. Let A be a commutative Banach algebra and
suppose x E A. Then

(i) x is quasi-regular if and only if 1 t R(x).
(ii) If A has an identity, then x is regular if and only

if 0 ~ R(x).
Corollary 3.4.3. Let A be a commutative Banach algebra with
identity e. If xl,x2, ... x n are in A, then either there exists
some T E. ~ A ) such that Xk(T) = 0, k = 1,2, .... ,n, or there exist
Yl'Y2""'Yn in A such that tk=lXkY
k
= e.
The content of Corollary 3.4.2 should be compared with Theorem
1.6.1, which asserts, among other things, that f E C(X), X being
a compact Hausdorff topological space, is singular if and only if
f(t) = 0 for some t E X.
3.S. Semisimplicity. 1he Gel'fand transformation on a commu-
tative Banach algebra A need not be injective. However, when it
..
is, the Gel'fand representation A of A is clearly an isomorphic
image of A under the Gel'fand transformation. In this case the

study of A via its representation A C C ~ A ) ) is greatly faci-
o
litated. Our concern in this section will be to see when the Gel'fand
transformation is injective. A new and useful notion in this regard
is the concept of the radical of a commutative Banach algebra, which
we shall define. First, however, we wish to give a special name to
those algebras on which the Gel'fand transformation is injective.
Definition 3.S.l. Let A be a commutative Banach algebra. If
the Gel'fand transformation on A is injective, then A is said
to be semisimple.
82
3. The Gel'fand Representation Theory
It is apparent that a commutative Banach algebra A is semi-

simple if and only if x E A and IIxllCD = 0 imply that x = o.
The notion of the radical is defined as follows:
Definition 3.5.2. Let A be a commutative Banach algebra.
The radical of A, denoted by Rad(A) , is defined as the intersec-
tion of all the maximal regular ideals in Ai that is,
Rad(A) = n M.
M(6(A)
Note we are here tacitly assuming, as indicated at the close of
Section 3.3, that 6(A) ~ ~ To be complete we should mention that
in the case 6(A) = ~ one defines Rad(A) = A, and A is then said
to be a radical algebra. Two examples of radical algebras were
given in Section 3.3.
Furthermore, we remark that in the definition of Rad(A) we
are considering 6(A) as the set of maximal regular ideals in A,
and not as the complex homomorphisms of A.
Evidently, Rad(A) is always a closed ideal in A.
Using some of our previous results it is possible to show that
Rad(A) is precisely the set of topologically nilpotent elements in
A, that is, Rad(A) = (x I x A, limnllxnlll/n = oj. Before proving
this result we state the next proposition. The proof, which follows
immediately from Theorem 3.4.1 and its first corollary, is left to
the reader.
Proposition 3.5.1. Let A be a commutative Banach algebra and
let x (A. Then the following are equivalent:
(i) x is topologically nilpotent.

(ii)
IIxll =
o.
CD
(iii)
II
x
ll
o
=
O.
(iv) o(x) = {oj.
3.S. Semisimplicity 83
Theorem 3.S.1. Let A be a commutative Banach algebra and let
x E A. Then the following are equivalent:
(i) x E Rad(A).
(ii) x is topologically nilpotent.
Proof .
-1
If x ~ Rad(A) , then x E T (0), T E ~ A ) ; that is,
= 0, T E ~ A ) . Thus 11;\1 = 0, and so, by Proposition
CD
T(X)

= X(T)
3.S.1, x is topologically nilpotent.
Conversely, if x is topologically nilpotent, then from Pro-

position 3.S.1 we see that IIxli = 0, whence X(T) = T(X) = o.
CD
T E ~ A ) . Therefore x E Rad(A), and the proof is complete.
o
It is now apparent from Proposition 3.S.1 and the comment fol-
lowing Definition 3.S.1 that a commutative Banach algebra A is
semisimple if and only if Rad{A) = (0). Two other necessary and
sufficient conditions for the semisimplicity are given in the next
corollary, whose elementary proof is left to the reader.
Corollary 3.S.1. Let A be a commutative Banach algebra. Then
the following are equivalent:
(i) A is semisimple.
(ii) Rad(A) = (0).
(iii) If x E A is topologically nilpotent, then x = O.
(iv) If x E A is such that a(x) = {oj. then x = o.
A little reflection also reveals that a commutative Banach alge-
bra A is semisimple if and only if the complex homomorphisms of
A separate the points of A.
The notions of semisimplicity and of the radical also have ana-
logs in the context of arbitrary Banach algebras. Since we shall be
84
3. The Gel'fand Representation Theory
concerned primarily with commutati\'e algebras, we have not discussed
these more general concepts. The interested reader is referred to
[N, pp. 162-165; Ri, pp. 55-59]. Similarly, although the Gel'fand
representation theory, as we have developed it, is not valid for
noncommutative Banach algebras, there does exist an analogous theory
for arbitrary Banach algebras based on the study of irreducible re-
presentations. Again we refer the interested reader to [N,Ri].
CHAPTER 4
THE GEL'FAND REPRESENTATION OF SOME SPECIFIC ALGEBRAS
4.0. Introduction. In this chapter we shall examine, in vari-
ous degrees of detail, the Gel'fand representation for some specific
commutative Banach algebras. Besides the value to be gained from
such an investigation with regard to understanding the mechanics of
the Gel'fand representation theory, we shall also apply the material
to prove several interesting theorems. For example, we shall prove
that, if X and Yare compact Ilausdorff topological spaces, then
C(X) and C(Y) are algebraically isomorphic if and only if X and
Yare homeomorphic; we shall obtain an extension of the classical
Riemann-Lebesgue Lemma of Fourier analysis to the context of locally
compact Abelian topological groups; and we shall show that, if f
is a continuous function on r = (z I z ~ , lzl = 1) whose Fourier
series is absolutely convergent, then the reciprocal of f is such
a continuous function provided f never vanishes. The proof of this
last result, generally known as Wiener's Theorem, is one of the
early celebrated accomplishments of the theory of Banach algebras.
4.1. C(X) and ~ X ) . Suppose X is a compact Hausdorff
topological space. Then C(X) with the usual pointwise operations
and the supremum norm is a commutative Banach algebra with identity.
We shall see shortly that A(C(X)) is homeomorphic to X and that
- C(X) is isometrically isomorphic to C(X). Thus, in an obvious
sense, C(X) is its own Gel'fand representation. Similar results
also hold for C (X) when X is a locally compact Hausdorff topo-
o
logical space. We shall utilize this result about the Gel'fand re-
presentation to show that C(X) and C(Y), where X and Yare
compact Hausdorff topological spaces, are algebraically isomorphic
85
86
4. The Gel'fand Representation of Specific Algebras
if and only if X and Yare homeomorphic.
To begin with we note that, if X is a compact Hausdorff topo-
logical space, then a(c(X)) since C(X) is a commutative
Banach algebra with identity. Indeed, it is quite easy to describe
a large collection of complex homomorphisms of C{X): Given t E X,
we define Tt{f) = f(t), f E C{X). It is apparent that T
t
E a(c(x)),
and the maximal ideal in C(X) corresponding to T
t
is obviously
M
t
= = {f I f E C(X), f(t) = oj. Furthermore, we claim that
every maximal ideal in C(X) is of this form; equivalently, every
complex homomorphism of C(X) is T
t
for some t in X.
Theorem 4.1.1. Let X be a compact Hausdorff topological space.
Then the mapping t - Nt = {f I f E C(X), f(t) = 0), t E X, is a
homeomorphism of X onto a(C(X)).
Proof. As already noted, each Nt is a maximal ideal in C(X).
Moreover, since C(X) separates the points of X, it is evident
that the mapping is injective. Now suppose M C C(X) is a maximal
ideal. If M is different from M
t
for each t in X, then,
given t E X, there must exist some f
t
E M such that ft(t) O.
Since f
t
is continuous, there exists some open neighborhood U
t
C X
of t for which ft(s) 0, s E Ute The collection of open sets
{Ut}t E X clearly forms an open covering of the compact space X,
and so there exists a finite subcovering -- that is, there exist
UtI' U
t2
,,U
tn
such that X = Consider the function
f E C(X) defined by
n
f = 1: f
t
'f
t
'
k=l k k
where the bar denotes complex conjugation.
Since M is an ideal, it is apparent that f E M. Moreover,
if sEX, then there exists some j, I j n,
Thus,
such that s E U
t
..
1
4.1. C(X) and C (X)
o
n
f(s) = E f
t
(s)f
t
(s)
k=l k k
n 2
= E 1ft (s)1
k=l k
> o.
87
Consequently f never vanishes on X, and so, by Theorem 1.6.1,
f is regular in C(X). From Corollary 1.1.2 we then deduce that
M cannot be maximal, contrary to assumption.
Hence there exists some t X
the mapping t - M
t
is surjective.
such that M =
Consequently we may identify X and 6(C(X))
Furthermore, we observe that, if f C(X), then
-.
M
t'
that is,
as point sets.
-. -.
f(t) = f(T
t
) = fet),
so that C(X) = C(X). Now by Theorem 3.2.2ei), X = A(C(X)) is a
compact Hausdorff topological space in the Gel'fand topology, and,
by Corollary 3.3.1, the Gel'fand topology is the weakest topology
-.
on A(C(X)) such that all the functions in C(X) = C{X) are continu-
ous. Since the original topology on X was compact and Hausdorff,
we see that the Gel'fand topology is weaker than the original topo-
logy on X, and hence they must coincide [W2' p. 84].
Therefore the mapping t - Mt' t E X,
X onto A(C(X)).
is a homeomorphism from
o
The topological argument that two comparable compact Hausdorff
topologies must coincide will be used again in the sequel.
An immediate corollary is the next result.
Corollary 4.1.1. Let X be a compact Hausdorff topological
space. Then the Gel'fand transformation on C(X) is the identity
mapping of C(X) onto itself.
88 4. The Gel'fand Representation of Specific Algebras
Proof. By Theorem 4.1.1 we may identify with x.
A
On doing this we see at once that, for each f e C(X), f(t) = f(t),
t E X, and so the Gel'fand transformation is the identity mapping.
L
In particular it is apparent that C(X) is isometrically iso-
morphic to its Gel'fand representation and that C(X) is semisimple.
If X is a noncompact locally compact Hausdorff topological
space, then, by applying these results to the Banach algebra of
continuous functions on the one-point compactification of X, one
can readily deduce the following theorem. The details are left to
the reader.
Theorem 4.1.2. Let X be a locally compact Hausdorff topolo-
gical space. Then
(i) The mapping
is a homeomorphism of
t - M
t
= (f I f E Co(X), f(t) = 0), t E X,
X onto (X)).
o
(ii) The Gel'fand transformation on C (X) is the identity
o
mapping of C (X) onto itself.
o
As before, we note that C (X)
o
is semisimple.
In the case that X is compact and Hausdorff we saw trivially

that C(X) is isometrically isomorphic to This result
is actually valid even for C(X), X being a locally compact Haus-
dorff topological space, although in this instance it need not be the
case that 6(C(X)) can be identified with X. Indeed it is obvious
that such an identification can be made only when X is compact.
To be precise we state and prove the following theorem:
Theorem 4.1.3. Let X be a locally compact Hausdorff topologi-
cal space. Then the Gel'fand transformation is an isometric isomor-
phism of C(X) onto
Proof. From the Beurling-Gel'fand Theorem (Corollary 3.4.1) we
4.1. C(X) and C (X)
o
89
see that
(fEe(X,
as IIflU = Uflln. Consequently the Ge1'fand transformation is an
m m
isometric isomorphism of C(X) onto C(X) C In par-
ticular, by the Ge1'fand Representation Theorem (Theorem 3.3.1),

C(X) is a closed suba1gebra of c{a[C(X)]) and contains the
,.
constant functions. To show that C(X) = it suffices,
,.
by the Stone-Weierstrass Theorem [L, p. 327], to prove that C(X)
is closed under complex conjugation.
To see this, suppose first that f E C(X) is real-valued. If
a = c + id E C and d 0, then
If(s) - al
2
= [f(s) _ c]2 + d
2
>0 (s E X).
Thus f - a is bounded away from zero, and so, as is easily seen,
f - a is regular in C(X). Consequently, by Theorem 3.4.1,
. ,.
a a(f) = R(f); that is, f - a never vanishes on Since
a is any complex number with nonzero imaginary part, we conclude
,.
that, if f E C(X) is real-valued, then so is f E C(a[C(X)]).
But if f is any element of C(X), then f = g + ih, where
g,h E C(X) are real-valued, whence
-" . . ,.
(f) = (g - ih) = g ih

= g + ih

= f,

from which it follows at once that C(X) is closed under complex
conjugation.
90 4. The Gel'fand Representation of Specific Algebras

Therefore C(X) =
o
It is evident that each point t in X, X being a locally
compact Hausdorff topological space, defines a complex homomorphism
of C(X) via the formula Tt(f) = f(t), f E C(X). Moreover, it is
not difficult to verify that the mapping t - T
t
, t E X, is a homeo-
morphism from X into It can further be shown that the
image of X in 6(C(X)) under this mapping is dense in 6(C(X)),
and so 6(C(X)), which is a compact Hausdorff space, is a compacti-
fication of X. To be precise it is the Stone-fech compactification
of X. We shall not carry out the details, but instead refer the
reader to [Ri, pp. 123 and 124; WI' pp. 269 and 270].
Finally we turn our attention to proving that C(X) and Cry)
are isomorphic as algebras if and only if X and Yare homeomorphic,
X and Y being compact Hausdorff topological spaces. Without fur-
ther ado we state the indicated theorem.
Theorem 4.1.4. Let X and Y be compact Hausdorff topological
spaces. Then the following are equivalent:
(i) X is homeomorphic to Y.
(ii) There exists an algebra isomorphism of Cry) onto C(X).
Proof. Suppose that X - Y is a homeomorphism of X onto
Y and define T: Cry) - C(X) by
T(f)(t) = f 0
(f E Cry); t ( X).
Then it is easily seen that T is an isometric algebra isomorphism
of Cry) onto C{X). The details are left to the reader.
Conversely, suppose T: C(Y) - C(X) is an algebra isomorphism
of Cry) onto C{X). Given t E X. we define Tt(f) = T(f)(t).
f E C(Y). It is evident that each such T
t
is a complex homomor-
phism of Cry); that is, T
t
E From Theorem 4.1.1 we deduce
4.1. C(X) and Co (X)
91
that there exists a unique E Y such that Tt(f) = f[,(t)],
f E C(Y). In this way we obtain a mapping ,: X - Y. Similarly,
since T-
l
: C(X) - C(Y) exists, we see that for each s E Y the
formula w (f) = T-l(f)(s), f E C(X), defines a complex homomorphism
s
w of C(X), whence, by Theorem 4.1.1, we deduce the existence of a
s
unique t(s) E X such that ws(f) = fetes)], f E C(X). Clearly
: Y - X, and some elementary computations reveal that '0 t(s) = s,
s E Y. Thus ,: X - Y is surjective. Similarly t 0 ,(t) = t,
t E X, from which it follows at once that X - Y is injective.
Hence ,: X - Y is bijective.
To see that X - Y is continuous. let t E X and suppose
o
U c Y is an open neighborhood of Since Y is a compact
Hausdorff topological space, it is normal, and we may appeal to
Urysohn's Lemma [W
2
, p. 55] to deduce the existence of some f E C(Y)
such that = I and = 0, ,(t) l U. But T(f) E C(X),
o
and W = {t I t E X, T(f)(t) = f[,(t)] 01 is an open subset of X
that contains t. Moreover, if t E W, then ,(t) E U, since
o
0, that is, C U. Hence X - Y is continuous.
Therefore X - Y is a homeomorphism, as it is a continuous
bijective mapping of a compact topological space to a Hausdorff
topological space [W
2
' p. 83], and the proof is complete. 0
It should be noted that the equation
liT (f) II = sup ITef) (t) I
CD t EX
= sup I f[,(t)] I
tEX
= IIfll
CD
shows that the isomorphism T is an isometry.
(f E C(Y
The theorem actually remains true even if one requires that T
92
4. The Gel'fand Representation of Specific Algebras
be only a linear isometric isomorphism. This result is usually
called the Banach-Stone Theorem p. 342]. The standard proof
involves the Krein-Mil'man Theorem p. 322] and is more intricate
than the one given here.
An obvious corollary is the next result.
Corollary 4.1.2. Let X and Y be locally compact Hausdorff
topological spaces. Then the following are equivalent:
(i) X is homeomorphic to Y.
Cii) There exists an algebra isomorphism of C (Y)
o
onto Co(X),
4.2.
n
C As indicated in Example 1.2.5, the space of
all n-times continuously differentiable complex-valued functions on
[a,b] is a commutative Banach algebra with identity. Using much
the same arguments as in the preceding section, one can establish
the following theorem. The details are left to the reader.
Theorem 4.2.1. Let E lR, a < b, and let nEil, n > O.
Then
(i) The mapping t M
t
= {f I f E Cn([a,b]), f(t) = 0),
a t < b, is a homeomorphism of [a,b] onto a(Cn([a,b])).
(ii) The Gel'fand transformation on Cn([a,b]) is the identity
mapping of Cn([a,b]) onto itself.
Again it is obvious that Cn([a,b]) is semisimple.
4.3. Let be a positive measure space.
As indicated in Example L is a commutative Banach
aD
algebra with identity. Thus, by the Gel'fand Representation Theorem
(Theorem 3.3.1) and Theorem the Gel'fand representation of
L is some subalgebra of C[aCL where
CD m
a(L = aCL) is the compact maximal ideal space of L
CD CD m
4.3. L (X,S,J,Io)
CD
93

As we shall shortly see, we actually find that L C[6(L )1
CD GO
and that the Gel 'fand transfomation is order preserving; that is.
if f E L (X.S,Il) and f(t) > 0 for ..,-a.lmost all t E X, then
A CD -
f(T) 0, ,. E 4 (L.) Moreover, ACL
CD
) provides us with an example
of a rather complicated maximal ideal space in that it is zero-dimen-
sional. For the sake of completeness we define this term explicitly.
Definition 4.3.1. Let X be a topological space. Then X is
said to be zerodimensional if the topology has a base that consists
of clopen sets -- that is, a base consisting of sets that are both
clos eel and open.
if X is a discrete topological space, it is zero-dimen-
but the converse need not be the case.
The previously indicated results are contained in the next
theorem.
Theorem 4.3.1. Let be a positive measure space. Then
(i) The Gel' fand transformation is an isometric orier preserving
isomorphism of L onto C[6{L
CD
(ii) ACL (X,S,I')
CD
is zero-dimensional.
Proof. By the same argument mutatis mutandis as used in proving
.
Theorem 4.1.3, we see that the Gel'fand transformation is an isometric
isoaorphisll of L (X,S, .. ) onto C[A(L )}; and, in particular, if
CD CD .. ..
f E L (X,S,p) is real valued, then so is f E C[A{t '1 = L
.. J _
But if f E L (X.S,p) is greater than or equal to zero ... -aIWlOst
..
everywhere, then one can evidently write f = g2, where lEt
2
is real valued.. Hence feT) = [geT)] 0, or E ACL.), and 50 the
Gel'and transformation is order preserving.
finally. we shall show that 6(L) is %ero-cliJDensional. To
..
this end let E c: X be measurable and Xs E L.(X,S,I') denote
94
4. The Gel'fand Representation of Specific Algebras
the characteristic function of E. Since = X
E
, it is apparent
A2 A A
that = Xs' whence we deduce that Rex
E
) C (O,l). Let
A A
E = {T I T ( XE(T) = I}. Evidently X
E
= X
E
; that is, the
Gel'fand transform of X
E
is the characteristic function of the

set E C The continuity of X
E
shows at once that E must
be clopen. Conversely, suppose Ec A(L) is a clopen set. Then,
A m
since L = C[A(L )], there exists some f E L
m
A
co 2 CD
such that f = XE' from which it follows that f = f. Thus f
is the characteristic function of some measurable set E C X, and
A
obviously we must have E = E.
Since in the preceding paragraph we allow measurable sets E of
possibly infinite measure, we see that the characteristic functions
of the measurable subsets of X generate a norm-dense subalgebra of
L Consequently, since the Gel'fand transformation is an
CD
isometry, we deduce that the set {X
e
lEe X, E measurable} c C[A(L
m
)]
generates a norm-dense subalgebra of C[ACL )]. It is then easily
CD
seen that the set {E lEe ACL ), E clopen) = {E ,
CD
E c X, E measurable:
forms a base for a zero-dimensional topology T on ACL) that is
CD
weaker than the Gel'fand topology on Since
m
ACL
CD
) is a
compact Hausdorff topological space in the Gel'fand topology, we see,
as in the proof of Theorem that in order to show that T
coincides with the Gel'fand topology it suffices to prove that T is
Hausdorff.
So suppose T
l
,T
2
E A(L
m
), Tl ; T
2
Then there exists some
." A
h E = C[ACL
CD
)] such that h(T
1
); hCT
2
), as C[ACL
CD
)]
separates the points of A(L). Without loss of generality we may
A CD A.
assume that h(T
1
); O. It is then easily verified that f = gg,
.. ".
where g = [h - h(T
2
)]/[hCT
l
) - h(T
2
)], belongs to C[ACL
CD
)] and

fCT
I
) = 1, f(T
2
) = 0, f(T) > 0, T E A(L
m
). Next let k be a finite
linear combination of characteristic functions of clopen subsets of
" .. .
ACL) such that kCT) > 0, T E ACL ), and for which IIf - kll < 1/3.
m - m
This is possible because {XE lEe ACL
CD
) , E clopen) generates a
norm-dense subalgebra of C[ACL )], as seen in the preceding paragraph.
m
4.4. A(D) 9S

(T I T E 6(L ), k(T) > 2/3] is a
m
whereas (T I T E 6(L ), k(T) < 1/3]
m
But then it is apparent that
clopen set containing T
I
,
is a clopen set containing
Thus T is Hausdorff.
T
2
, and these clopen sets are disjoint.
Therefore T coincides with the Gel'fand topology on 6(L)
m
[W
2
, p. 84], and so the Gel'fand topology on 6 (Lm) is zero-dimen-
sional.
o
It is, of course, once again clear that L
m
is semisimple.
describe the Gel'fand topology on 6(L } in somewhat
CD
different terms. Suppose T E 6(L ) and let C denote the maximal
m
connected subset of 6(L ) that contains T. Since the closure of
m
a connected set is connected, we see that C is closed. Suppose
wEe and w T. Since 6(L) is zero-dimensional and Hausdorff,
m
there exists a clopen set Ec 6(L) such that T f, w l E. But
m
then in the relative topology on the connected set C we see that
Ene is open, C - (f n C) is open, and C = (E n C) U [C - (E n C)]
thereby contradicting the connectedness of C. Consequently C = (T].
A topological space with this property, that is, a topological
space such that the maximal connected set C that contains a given
point T of the space is just C = (T), is said to be totally
disconnected. In particular, 6(L) is totally disconnected.
m
The argument above actually shows that, if a topological space
of at least two points is zero-dimensional and To' that is, given
any two distinct points there exists an open neighborhood of one that
does not contain the other, then it must be totally disconnected.
Conversely, it can be shown that a totally disconnected locally
compact Hausdorff topological space is zero-dimensional [HR
1
, p. 12].
4.4. Next we wish to consider the commutative Banach
algebra with identity A(D), that is, the algebra of all continuous
96 4. The Ge1'fand Representation of Specific Algebras
complex-valued functions on D = (z I z Izl < 1) that are
analytic on the interior of o. We shall see that can be
identified with D, which on the surface may not be very surprising.
This will, however, provide us with an example of a commutative
Banach algebra A that contains a subalgebra B whose maximal
ideal space properly contains that of A, that is, such that
and 6(A). Some further observations about
will set the scene for a discussion of finitely generated
algebras in the next section.
then
ACO)
As in the preceding sections, it is apparent that, if z ( 0,
Tz(f) = fez), f A(D), defines a complex homomorphism of
and that M = T-I(O) = (f I f E A(D), fez) = 0).
z z
On the other hand, let
that is, h(z) = z, zED.
Since IiTIi 1, we see that
f E A(D).
h E A(D) denote the identity function,
If T E we set T(h) = ,.
< 1. We claim that Tef) = f(C),
To see this let f (A(O) and for each r, 0 < r < 1, define
fr(z) = f(rz), zED. Then fr is continuous on (z I z Izi < l/r).
In particular, f E ACD) and the power series expansion of f
r k r
about z = 0, say converges uniformly to fr on
(z I z Izi < 1). Thus we see that
T(f )
CD k
=
T[ 1: ak(r)z ]
r
k=O
CD k
= 1: ak(r)T(h)
k=O
CD
=
1: ak(r)c
k
k=O
= f (e) (0 < r < 1).
r
Moreover, an elementary argument involving the uniform continuity of
4.4. A(O) 97
f on 0 reveals that limr_lllf
r
- flieD = O. Thus we have
T(f) = lim T(f )
r-l r
= lim f (C)
r-l r
= f(C),
which is what we set out to prove.
Therefore we see that and 0 can be identified as
point sets, and the usual argument employed twice before concerning
the comparability of compact Hausdorff topologies shows that the
usual topology and the Gel'fand topology on D coincide.
The validity of the next theorem is now apparent.
Theorem 4.4.1. Let D = {z I z E C, Izl 1). Then
<::
(i) The mapping z - M = {f I f E A(O), fez) = 0), z E 0, is
z
a homeomorphism of 0 onto
(ii) The Gel'fand transformation on A(D) is the identity
mapping of A(D) onto itself.
Clearly A(D) is semisimple.
If, as usual, we set r = (z I z 1z1 = 1), then, by means
of the Maximum Modulus Theorem [A, p. 134], it is easily verified
that the mapping A(D) - C(f), defined by = fez),
z E r, f E A(D), is an isometric isomorphism of ACD) into C(I).
Of course, is just the restriction of f to f. In this way
one can consider A(D) as a closed subalgebra B(I) of the commu-
tative Banach algebra C(I). The previous development shows that
A(C(f)) = r D = A(ACD)) = AC8{'l).
.-
Furthermore, we see at once from Theorem 3.4.1 that a(h) = R{h)
= R(h) = D = A(ACD)), where h(z) = z, zED. This observation
98 4. The Gel'fand Representation of Specific Algebras
makes it clear why the term "spectrum of A(D)" is an appropriate
label for the maximal ideal space.
In an obvious sense the function h generates A(D). We shall
see in the next section that an analog of the connection between
a(h) and A(A(D is valid in any commutative Banach algebra with
identity that is finitely generated.
4.5. Finitely Generated Banach Algebras. We begin this section
with two definitions.
Definition 4.5.1. Let A be a Banach algebra with identity e.
A subset E C A is said to generate A if, whenever B C A is a
closed subalgebra such that e E Band E C B, then B = A. The
algebra A is said to be finitely generated if there exists a
finite subset E C A that generates A.
In particular, a commutative Banach algebra A with identity
e is generated by E C A if every element of A is the norm limit
of a sequence of polynomials in e and the elements of E. Thus,
for example, A(D) is generated by E = (h), where h(z) = z, zED.
Definition 4.5.2. Let A be a commutative Banach algebra with
identity and let x
l
,x
2
, ,x
n
be in A. Then the joint spectrum
of x
l
,x
2
, .. ,x
n
is the subset of en defined by
T E A(A)]

= (ex
I
(T),x
2
(T), . ,x
n
(T T E a(A).
If n = 1, then it is evident from Theorem 3.4.1 that the
joint spectrum of a single element reduces to the notion of spectrum
previously introduced.
The following proposition about the joint spectrum is easily
established. The details are left to the reader.
4.5. Finitely Generated Algebras
99
Proposition 4.5.1. Let A be a commutative Banach algebra
with identity and let x
l
,x
2
, ... ,x
n
be in A. Then O(x
l
,x
2
, ... ,x
n
)
is a nonempty compact subset of C
n
that is contained in the
polydisk
Our main result is the next theorem.
Theorem 4.5.1. Let A be a commutative Banach algebra with
identity e that is generated by x
l
,x
2
, .. ,x
n
Then
(i) The mapping T - (T(x
1
),T(x
2
), ... ,T(x
n
)), T E is a
homeomorphism of onto o(x
l
,x
2
, ,x
n
)
(ii)
If
denotes the kth coordinate function in that
A
is, .. ,zn) = zk' k = 1,2, .. ,n,
a(x
1
,x
2
, . ,x
n
), k = 1,2, ,n.
then x
k
= on
A
(iii) If x A, then x is the uniform limit on a(x
l
,x
2
, . ,x
n
)
of polynomials in the n variables zl,z2, ,zn.
A
(iv) If x E A, then x is continuous on a(x
l
,x
2
, .. ,x
n
)
and analytic on the interior of a(x
l
,x
2
, ,x
n
).
Proof. Since the Gel'fand topology on A(A) is the relative
weak* topology, it is evident that the mapping
is continuous from A(A) onto a(x
l
,x
2
, ,x
n
). Moreover, the map-
ping is injective because, if T,W E and
then, since T(e) = wee) = 1 and x
l
,x
2
, .. ,x
n
deduce that T(X) = x E A, whence T = w.
mapping T - (T(X
1
),T(X
2
), ,T(x
n
)) from
generate A, we
Consequently the
to O(x
1
,x
2
, . ,x
n
)
100 4. The Gel'fand Representation of Specific Algebras
is a homeomorphism, as 6(A) is compact and the topology on
a(x
l
,x
2
, . ,x
n
) is Hausdorff.
Using this identification of A(A) with O(x
l
,x
2
, ... ,x
n
), we
see at once that, for each k = 1,2, ... ,n,
It ..
T(X
t
) = Xt(T) = xk[(T(xI),T(x21, . ,T(xn]
(T 6(A,
..
whence x
k
= k = 1,2, . ,n, on a(x
l
,x
2
, .. ,x
n
).
From this result it is apparent that, if p is a polynomial in
n variables, then P(x
l
,x
2
, ... ,x
n
) A and
It
P(x
l
,x
2
,,x
n
) = P(zl,z2,,zn)
Thus, since x
l
,x
2
, .. ,x
n
generate A, given x E A, we see that
for each s > 0 there exists some polynomial Ps of n variables
such that IIx - PS(xl,x2, .. ,xnlll < s. Consequently, since the
Gel'fand transformation is norm decreasing, we see that
It ..
IIx - PS(zl,z2, . ,zn1IlCD < s, and so x is the uniform limit on
a(x
l
,x
2
, .. ,x
n
) of polynomials in zl,z2, .. ,zn.
The last assertion of the theorem is now evident.
[J
Since the function h E A(D) generates A(D), we see that the
first portion of Theorem 4.4.1 is a special case of Theorem 4.5.1.
Loosely speaking, Theorem 4.5.1 says that finitely generated
commutative Banach algebras with identity are like algebras of con-
tinuous functions on compact subsets K of ~ for some n, such
that the functions in the algebra are analytic on the interior of K.
(K is, of course, the joint spectrum of the generators of the Banach
algebra.) The question of whether the Gel'fand representation of such
a Banach algebra consists of all the continuous functions on the
joint spectrum that are analytic on the interior of the joint spec-
trum and the determination of which of these functions belong to the
Gel'fand representation of A are quite delicate and intricate matters.
4.6. AC(I1
101
As an example of one such theorem we state. without proof, the
following result. The notation is as set forth in Example 1.2.4.
Theorem 4.5.2 Theorem). Let K C C be compact.
If C - K is connected. then P(K) = A(K).
A discussion of such material is beyond the intent of this volume.
The interested reader is referred to [B,Ga,S,Wm
1
,l"m
2
] for treatments
of this and related topics.
In view of the previous remarks it seems apprepriate to describe
the maximal ideal spaces of P(K). R(K), and A(K) when K is a
compact subset of C. It can be shown that A(A(K)) = A(R(K)) = K,
whereas A(P(K)) is equal to the complement in C of the unbounded
component of - K (see, for example, [Ga, pp. 27 and 28; S, pp. 273
and 274]).
4.6. AC(I). We recall that AC(f) is the commutative Banach
algebra with identity consisting of all the f E C(I1 with absolutely
convergent Fourier series, that is, such that the Fourier transform

f belongs to Ll(Z) , where the algebra operations are the usual
..
pointwise ones and IIfllAc = IIflll, f E AC (f).
it t
Obviously, if e E f and TtCf) = feel ). f E AC(I1. then
T
t
defines a complex homomorphism of AC(f). Now suppose T E A(AC(f))
. it it
and let C = T(h), where h E AC(I1 lS such that hee ) = e ,
e
it
E f. Evidently h is regular in ACCf) and h-
1
= h. Thus
C 0 and T(h) = I'C. But then we see that
and
whence we conclude that 1'1 = Ii that is, ,E f. The fact that
IIhlll = 1 and = I follows at once from the classical identity
102
4. The Gel'fand Representation of Specific Algebras
2
1
n rt1_n e
ikt
dt = 1 f k 0 J _ or =,
1 In ikt
2n -n e dt = 0 for k E Z, k -j 0
and the definition of the Fourier coefficients given in Example 1.2.9.
However, it is apparent that, for any f E AC(f) , the Fourier
series f(k)e
ikt
converges uniformly to f on r and also
converges to f in AC(r). Hence
CD
f(k)C
k
f(C) = 1:

CD
f(k)T(h)k
= 1:
k=-CD
ca
= T [ t f(k)e
ikt
]
k=-m
= T(f) (f E AC(f).
Consequently, since AC(f) separates the points of r, we see
that the mapping e
it
- M
t
= = {f I f E AC(r) , f(e
it
) = 0)
is bijective from r to A(AC(r. By what should now be a familiar
argument, we further deduce that the mapping is a homeomorphism.
Moreover, identifying r with A(AC(f) as above, we see that, if
A it it it
f E AC(f) , then rCe ) = Tt(f) = fee ), e E f, so that the
Gel'fand transformation is just the identity mapping. Note carefully
A
in the last formula that f denotes the Gel'fand transform, and not
the Fourier transform of f. We summarize the foregoing in the next
theorem.
Theorem 4.6. 1.
t
Let r = {e
1
I -Ti < t < n 1. Then
(i) The mapping e
it
- Nt = (f I f E AC(f) , f(e
it
)
is a homeomorphism of r onto A(AC(f).
it
= 0), e E r,
(ii) The Gel'fand transformation is the identity mapping of
AC (n into C (f)
4.6. AC(f) 103
Again AC(f) is semisimple, but this time, in contrast to the
situation for C(X), X being a compact Hausdorff topological space,
and A(D), the Gel'fand transformation is not an isometry.
One can also use Theorem 4.5.1 to obtain a description of
Indeed, it is easily verified that AC(f) is generated by the func-
tions hand h introduced above, and hence is homeo-
morphic to
a(h,h) = {(T(h),T(h)) I T E = ((T(h),T(h)-l) I T E
However, by the same arguments as before, we see that IT(h)1 =
IT(h)-ll = 1, T E A(AC(f)), and so we deduce that
a(h,h) = {(eit,e-
it
) I e
it
E f).
Thus, using Theorem 4.5.1, we have obtained a description of
A(AC(f)) as a subset of whereas previously we identi-
fied A(AC(f)) with a compact subset of C, namely f. Hence
different approaches to the problem of describing the maximal ideal
space of a commutative Banach algebra may lead to superficially rather
different results.
As the final subject of this section we wish to discuss Wiener's
Theorem. In proving his celebrated general Tauberian Theorem
[Wr
l
, pp. 72-103] Wiener was led to ask: Under what conditions on a
function f E C(f) whose Fourier series is absolutely convergent
is it the case that l/f E C(f) and has a Fourier series that is
absolutely convergent? In terms of the Banach algebra AC(f) this
question clearly reduces to determining when an element in AC(f)
is regular. Since AC(f) C C(f), it is evident from Theorem 1.6.1
that f E AC(f) will be regular in C(f) if and only if f(e
it
) 0,
e
it
E f, but it is not at all obvious in this case that l/f E AC(f).
Our knowledge of however, allows us to show quite simply
that this is indeed the case.
104 4. The Gel'fand Representation of Specific Algebras
Theorem 4.6.2 (Wiener's Theorem). If f E AC(f) is such that
f(e
it
) ~ 0, e
it
e" then -l/f E AC(r).
Proof. From Theorem 4.6.1 we know that f can be identified
with ~ A C f ) ) . On making this identification we see that the assump-
tion f(e
it
) ~ 0, e
it
E f, means that f belongs to no maximal
ideal in AC(f). Hence, by Corollary 1.1.2, f is regular; that
is, l/f E AC(f).
o
This is not, of course, Wiener's original proof. That proof,
which is a good deal more complicated, can be found in [Wr
l
, pp. 86-91].
4.7. L1(G). In this section we shall begin to investigate the
commutative Banach algebra LI(G), where G is a locally compact
Abelian topological group. The study of this algebra will occupy
us periodically throughout the succeeding chapters. We recall that
LI(G) consists of equivalence classes of complex-valued Borel
measurable functions on G that are integrable with respect to a
Haar measure A on G and that the multiplication in Ll(G) is
given by convolution. Before we describe the maximal ideal space
and the Gel'fand representation of LI(G), we wish to develop a
number of other results about Ll(G) that will be useful in this
and subsequent sections.
Definition 4.7.1. Let G be a locally compact Abelian topolo-
gical group and let I < P ~ m. For each s E G the mapping
T L (G) - L (G) is defined by T (f)(t) = f(t - s), f ( L (G)
s P P s P
and t E G, where the formula is understood to apply for almost
all t in G.
For obvious reasons the mappings T, s E G, are called
s
translation operators. Some elementary properties of these mappings
are collected in the next proposition.
Proposition 4.7.1. Let G be a locally compact Abelian topo-
logical group.
(i) If 1 ~ P ~ CD,
f E L (G), s E G.
then T E L(L (G)) and liT (f)1I = IIfll ,
s p s p p
P
(ii) If 1 ~ < CD, then for each f E L (G) the mapping from
p
G to L (G), defined by s ~ T (f), s E G, is uniformly continuous.
p s
Proof. The proof of part (i) is trivial, and that of part (ii)
follows easily after noting that Cc(G) is norm dense in Lp(G),
1 < p < m. The details are left to the reader.
o
If G is a locally compact Abelian topological group and
f E L (G), gEL (G), where 1 < p < CD, IIp + l/q = 1, then we
p q - -
define f * g for s E G by
f * g(s) = IG f(s - t)g(t) dArt),
that is, by the usual convolution formula. It is then easily seen
that f * g is a bounded function on G. Indeed, for almost all
s E G we see, by applying HHlder's inequality and Proposition 4.7.1,
that
I f * g (s) I = I J G T s (1) (t) g (t ) dA (t) I
< liT (1) II ))g))
- s p q
= 11111pllgllq
= II fllpllglIq,
where 1(u) = f(-u).
Somewhat more can be said, as is seen in the next proposition.
106 4. The Gel'fand Representation of Specific Algebras
Proposition 4.7.2. Let G be a locally compact Abelian topo-
logical group.
(i) If 1 < P < m and IIp + l/q = 1, then f * g E C(G),
f E Lp(G) and g E Lq(G).
(ii)
f E L CG)
p
If 1 < P < m and IIp + l/q = 1, then f * g E CoCG) ,
and g E Lq (G).
Proof. To prove part (i) we note that, given
gEL (G), for each s,t E G we have
q
f E L (G)
P
and
If * g(s) - f * g(t}1 = IIG TS(l) (u)g(u) dA(u} - IG Tt(l) (u}g{u) dA{u)1
:5: UTs(l) - Tt(1)lIpllgllq.
The desired conclusion is now immediate from Proposition 4.7.1(ii)
since we may assume,without loss of generality, that I <p < m.
Again the proof of part (ii) follows easily on recalling the
density of C (G) in L (G), I < p < m. 0
C P
It should be observed that the proof of Proposition 4.7.2(i)
actually shows that f * g, f E L (G) and gEL (G), is a bounded
p q
uniformly continuous function on G. We shall make no use of this
latter fact.
Our first result concerning the Banach algebra LICG} is a
necessary and sufficient condition to ensure that LICG) has an
identity. A locally compact Abelian topological group is said to
be discrete if its topology is the discrete topology. In this case
it is evident that the counting measure A on G (that is, the
measure A on G such that, if E C G, then ACE) is the number of
points in E if E is finite and A(E) is plus infinity if E is
infinite) is a translation-invariant regular Borel measure on G,
and hence A is a Haar measure. Following the usual convention,
we shall always assume that our Haar measure on a discrete locally
107
compact Abelian topological group G is the counting measure. In
particular this means that A((SJ) = 1, s E G.
Theorem 4.7.1. Let G be a locally compact Abelian topological
group. Then the following are equivalent:
Ci) LI(G) is a commutative Banach algebra with identity.
(ii) G is discrete.
Proof. Suppose G is discrete and define e (Ll(G) by
e(t) = 1, for t = 0, and e(t) = 0, for t ~ 0, t E G, where we
have used 0 to denote the additive identity in both C and G.
An easy computation then reveals that e * f = f * e = f, f E L1(G),
whence e is an identity for LI(G). Thus part (ii) of the theorem
implies part (i).
Conversely, suppose Ll(G)
we claim that there exists some
has an identity e.
p > 0 such that
inf{A(O) I 0 c G, 0 ~ ~ , 0 open) > p.
In this case
Indeed, suppose this were not the case and let 0 < e < 1. Since
e ~ Ll(G), it is apparent that the formula
~ E ) = IE le(t)1 dA(t),
where E is a Borel subset of G, defines a regular Borel measure
on G that is absolutely continuous with respect to A. Thus there
exists some 6 > 0 such that, whenever E eGis a Borel set for
which A(E) < 0, then ~ E ) < e.
open neighborhood U of 0 in G
In particular there exists an
such that
Su le(t)1 dA(t) < e.
This latter assertion is valid as we are assuming that there exist
open subsets of G with arbitrarily small Haar measure and because
Haar measure is translation invariant.
108
4. The Gel'fand Representation of Specific Algebras
However, from the continuity of the group operations in G
there exists an open neighborhood W of 0 in G such that W = -w
and W + W C U. Denoting the characteristic function of W by Xw,
we see that
Xw(s) = * e(s)
= IG Xw(s - t)e(t) dAft)
= fw+s Xw(s - t)e(t) dA(t)
Consequently, if sEW, then
1 = Xw(s) Iw+s - t)e(t) dAft)
< Iw+w le(t)1 dAft)
le(t)1 dAft)
< e,
(s E G).
contrary to the choice of e. Hence there exists some p > 0 such
that
inf{A(O) I 0 c G, 0 0 open) > p.
But now suppose that 0 eGis open and that 0 has compact
closure. The regularity of Haar measure entails that ACO) If
o contains n distinct points, then clearly there exist pairwise
disjoint open neighborhoods of these n points that are contained
in O. Call them 01,02, ... ,On0 Then
n
ACO) > A( U Ok)
k=l
n
= t A(Ok) > np.
k=l
Since A(O) < m, it follows at once that must be finite. Hence
every open subset of G with compact closure is finite. In particular,
every compact neighborhood of a point in G is finite, whence, since
the topology of G is Hausdorff, we conclude that single points are
open sets.
109
Therefore G
part (ii).
is discrete, and part (i) of the theorem implies
o
On the other hand, Ll(G) always contains what is as an
approximate identity, whether G is discrete or not. We give the
definition of an approximate identity in the context of an arbitrary
commutative Banach algebra.
Definition 4.7.2. Let A be a commutative Banach algebra. A
net (u
a
) C A is said to be an approximate identity if
(i)
(ii)
sup lIu II < go.
a a
lim lIu x - xII = 0
a a
(x E A).
Obviously, if A has an identity e, then (u) is an approxi-
a
mate identity on setting u = e for all a. The converse need not
a
be valid, as seen from the next theorem and Theorem 4.7.1.
Before we can prove the existence of an approximate identity
in Ll (G) we need one preliminary result.
Proposition 4.7.3.
logical group and let
Let G be a locally compact Abelian topo-
A be a Haar measure on G. If U is a
nonempty open subset of G, then A(U) > o.
I
Proof. Since A is translation invariant, we need consider
only nonempty open subsets that contain the identity of G. Suppose
U eGis such a set and A(U) = O. Then A(t + U) = 0, t G.
If KeG is compact, then (t + U l t E K) forms an open
covering of K, and so there exist t
l
,t
2
, ... ,t in K such that
n n
K C vk=l(t
k
+ U). Consequently A(K) < + U) = O.
Since
K is arbitrary, it follows at once from the regularity of A
A = 0, which is a contradiction.
Therefore A(U) > o.
that
o
110 4. The Gel'fand Representation of Specific Algebras
Theorem 4.7.2. Let G be a locally compact Abelian topological
group. Then LI(G) contains an approximate identity.
Proof. Let (u 1 denote the family of open neighborhoods of
a
o in G that have compact closure and are such that U = -u .
a a
Such neighborhoods exist because G is locally compact. Then define
u
a
=
function of U .
a
where as usual Xu denotes the characteristic
In view of Proposftion 4.7.3, we see that each u
ex
is well defined, and, as is easily verified, each u defines a
a
nonnegative element of Ll (G) n LCD(G) such that lIuaU
I
= 1. Moreover,
(u) becomes a net provided we set a > e if and only if U C U
c

a a
10 see that (ual is an approximate identity in Ll(G) we need
only show that limJlu
a
* f - fill = 0 for each fELl (G) Let
f E L
1
(G). Then we see, on applying Fubini's Theorem [Ry, pp. 269
and 270], that
lIu
a
* f - fill = IGIIG f(t - s)ua(s) dA,(s) - f(t)1 dA,(t)
= IGIIG [f(t - s) - f(t)]ua(s) dA,(s)1 dA,(t)
'f(t - s) - f(t)luex(s) dA,(s)] dArt)
= IG[I
G
11s(f)(t) - f(t)l dArt)] ua(s) dA,(s)
= IG 111s(f) - fIl1ua(s) dA,(s).
However, by Proposition 4.7.1(ii), 111s(f) - fill is a uniformly
continuous function on G, whence, given e > 0, there exists some
U
ao
such that, if s E Ucro' then 111s(f) - fli
l
< e. Consequently,
if a > a, then
o
lI
u
a
* f - fill Iu 111 s (f) - fill ua(s) dA (5)
a
< sfu ua(s) dA(s)
a
= elluall
l
III
= c.
Therefore {u
a
) C Ll(G) is an approximate identity. 0
A slight modification of the preceding argument using the fact
that C (G) is norm dense in L (G), I < p < -, reveals that
c p-
lim Ilu * f - fll = 0, f E L (G), where (u 1 is defined as in the
d a p p a
proof of Theorem 4.7.2. The details are left to the reader. It
should be noted that L (G), I < p < -, is a Banach algebra under
p
convolution only if G is compact. This is a nontrivial result
[HR
2
, pp. 469-472]. In any case, however, it is easily seen that
sup lIu II = -, 1 < P < -, when G is not discrete. This follows
a a p
at once on noting that, if G is not discrete, then lim A(U ) = 0
11 a a'
and lIu II = A(U 1- q where IIp + l/q = 1.
a p a'
Let us now turn our attention to describing the maximal ideal
space of L
1
(G). If T E 6{L
I
(G, then we know that T E LI(G)*,
UTII 1, and T(f * g) = T(f)T(g), f,g E L1(G). Since Ll (G)* may
be identified with Lm(G) [HR
I
, p. 148; L, pp. 59 and 60], this implies
the existence of a nonzero element y E L (G) such that
-
T(f) = IG f(t)y(-t) dAft) (f E Ll (G
Moreover, applying Fubini's Theorem [Ry, pp. 269 and 270], we see,
on the one hand, that
T(f * g) = IG f * g(t)y(-t) dAft)
= IG[jG f(t - s)g(s) dA(s)] y(-t) dA{t)
= IG g(s)[jG f(t)y(-t - s) dA(t)] dA(s),
while, on the other hand, we have
T{f * g) = T(f)T(g)
= IG f(t)y(-t) dAft) IG g(s)y(-s) dA(s)
112 4. The Gel'fand Representation of Specific Algebras
= IG g(s) [y{-s)i
G
f(t)y(-t) dArt)] dA(s) (f, g E Ll (G))
From Proposition 4.7.2(i) we know that
IG f{t)y(-t - s) dArt)
(s E G)
defines an element of C(G), whence from the weak* denseness of
Ll(G) in [L, p. 245] we conclude that, for each f E Ll(G),
IG f(t)y(-t - s) dArt) = y(-S)!G f(t)y(-t) dArt)
for almost all s E G. In particular, if fo E LI(G) is such that
T(f ) F 0, we see that
o
J
G
fo(t)y(-t - s) dArt)
y{ -s) = ---------
,fG f 0 (t)y (-t) dA (t)
=
T[Ts{fo)]
T{fo)
for almost all s ( G. Since the expression on the right-hand side
of the preceding equations is evidently a continuous function of s,
we may assume without loss of generality that the element y E L (G)

corresponding to T is itself a continuous function, namely
y(-s) = T[Ts{fo)]/T(f
o
)' s E G.
But then, repeating the previous argument, we deduce from the
identity
IG f{t)y(-t - s) dArt) = IG f{t)y(-s)y(-t) dArt)
that y(-t - s) = y(-t)y(-s), t,s E G. Consequently see that,
for each T E there exists some y E C(G) such that
yet + 5) = y(t)y(s), t,s E G, and
T(f) = IG f(t)y(-t) dArt)
(f L
1
(G)).
Moreover, it is easily verified that y is unique.
113
Conversely, straightforward computations reveal that, if y E Cee)
is such that yet + s) = y(t)y(s), t,s E G, then the formula
T(f) = IG f{t)y(-t) dA(t)
defines an element T E 6(L
l
(G)) provided y is not identically
zero.
Before we summarize this discussion in the next theorem, we
wish to look a little further at such functions y. So suppose
y E C(G) is such that yet + s) = y(t)y(s), t,s E G. If Y is not
identically zero, then yeO) = yeO + 0) = y(O)y(O) shows that
yeO) = 1, and then 1 = yeO) = yet - t) = y{t)y(-t), t E G, reveals
that yet) ~ 0 and y(-t) = l/y(t), t E G. Furthermore, we claim
that ly{t)I = 1, t E G, on the following grounds: if there existed
some t such that ly(t)1 > I, then ly(nt)I = ly(t)I
n
, n = 1,2,3, . ,
contradicts the boundedness of y. Hence ly{t)1 < 1, t E G. Com-
bining this with the fact that y{-t) = l/y(t), t E G, we have
ly(t)1 = 1, t ~ G.
Thus we see that, if y E C(G) is not identically zero and
yet + s) = y{t)y(s), t,s E G, then
(i) yeO) = 1.
(ii) Iy(t) I = 1, t E G.
(iii) y(-t) = l/y(t) = yet), t E G.
That is, each such y is a continuous homomorphism of e into
r = {, I , E ~ "I = 1). Because these homomorphisms of G play
a central role in the study of Ll(e), as they can be identified
with 6(L
l
(G)), we wish to give them a special name.
Definition 4.7.3.
gical group. Then by
morphisms of G into
Let G be a locally compact Abelian topolo-

G we denote the set of all continuous homo-

r = {, , , E ~ I" = 1). If y E G, we say
that y is a continuous character of G.
114
4. The Gel'fand Representation of Specific Algebras
For notational reasons we wish to make a special convention

regarding the elements y E G; namely, if y E G, we shall usually
write ~ t ) = (t,y), t E G. Note that with this notation, y(-t) =
t ~ y ) , t E G. The utility of this notation will become apparent in
the sequel.
Recalling the Gel'fand Representation Theorem (Theorem 3.3.1),
we can summarize our development to this point in the following
theorem:
Theorem 4.7.3. Let G be a locally compact Abelian topological
group.
(i) The formula
T(f) = IG f(t) (t,y) d ~ t )

determines a bijective correspondence between A{LI(G)) and G.

(ii) In the Gel'fand topology on G, G is a locally compact

Hausdorff topological space. If G is discrete, then G is a
compact Hausdorff topological space.
(iii) If f E LI(G), the Gel'fand transform of f is given by

fey) = IG f(t) (t,y) d ~ t ) (y E G).
(iv) The Gel'fand transformation is a norm-decreasing homomor-

phism of LICG) onto a subalgebra LI(G) of C (G) that separates
o.
the points of G. If G is discrete, then LI(G) contains the
constant functions.
As an immediate corollary we have the next result.
Corollary 4.7.1 (Riemann-Lebesgue Lemma). Let G be a locally
compact Abelian topological group. If f (LICG), then

limy_ fey) = 0; that is, given ,> 0, there exists some compact
. . - .
set KeG such that If(y)1 < e, y E G - K.
lIS
Because of the identifications indicated in Theorem 4.7.3 we
"
shall generally speak of G as the maximal ideal space of L
1
(G)
"
and of the Gel'fand transform of f E LlCG) as a function on G.
Note also that the norm-decreasing nature of the Gel'fand trans-
"
formation means explicitly that IIfIlCD:5 IIflll, f E Ll(G).
For the classical groups, for example, r and the Gel'fand
transform on LICG) is precisely the familiar Fourier transform.
" "
We shall see this momentarily when we describe r and For this
reason we shall use the terms "Gel'fand transform" and "Fourier
transform" interchangeably in discussing Ll!G). Also, recalling
Example 1.2.9, it is natural to write LICG) = FLl(G).
Before we proceed any further with the general investigation
"" "
of LICG), let us describe in detail Rand r. Suppose y E
Since (O,y) = 1, there exists some 6 > 0 such that
(s,y) ds O.
Consequently we deduce from
(t + s,y) ds = (s,y) ds (t E IR)
that
(t + s,y) ds
Ct,y) = 6
J
o
(s,y) ds
J
t+6
t (s,y) ds
=
J
6
o Cs,y) ds
(t E lR)
Thus we see at once that y is differentiable on and so
y' (t)
= lim (t + h,y) - (t,y)
h-O h
116 4. The Gel'fand Representation of Specific Algebras
= (t,V) lim (h,V) (O,y)
h-O
= V' CO)Vet)
(t (lR).
Solving this elementary differential equation and using the facts
that vCO) = I and IVCt) I = I, t (!h, we see that there exists some
E lR such that y (t) = (t, V) = t E iRe Conversely, it is ob-
vious that for each E 1f\. the function belongs to R, so
that = {e
igo
I E Since the correspondence - e
ig
., g E
"
is bijective, we can also identify lit with the point set and
we shall often do this. It is now evident that the Gel'fand trans-
form of f E is indeed the classical Fourier transform .
..
The description of f is obtained most easily by utilizing the
preceding results. To do this we note that r is isomorphic to
iV2nZ, where 2n = (2nk I k Z). Then it is easy to check that
.. "
V (f if and only if V and V is periodic with period 2n.
Thus we see that r = (e
ik
. IkE if). As in the case of lri, we
.. . ike
generally identify f with L by means of the mapplng k - e ,
k E oZ.
" ..
To round out this discussion we shall describe L. If V E Z,
then there is some ,E f such that (1,V) = ,. Clearly then we must
have (k,V) = ,k, k E L. Conversely, given ,E r, the mapping
k - ,k, k E Z, is obviously an element of Z. Thus we see that
"
1 = (,Co) I , E fl.
As usual, we generally identify with f.
Allowing for a certain amount of imprecision, we observe that
..".. ..,," "" ..
(JR.) = l[( = lR, (f) = It. = f, and = f = lL.. These observations
are indicative of a general phenomenon for locally compact Abelian
.. "
topological groups G, which, loosely speaking, says that (G) = G.
We shall return to this important result, known as the Pontryagin
Duality in Section 10.5, where we shall discuss it in detail.
It is obvious that for the algebras previously discussed in
117
this chapter the Gel'fand transformation is injective, that is, the
algebras are semisimple. The algebra L (G) is also semisimple,
1
but the proof of this fact is not entirely trivial.
Theorem 4.7.4. Let G be a locally compact Abelian topological
group. Then LI(G) is semisimple.
Proof. In view of Theorem 3.5.1 and Corollary 3.5.1, it suffices
to show that, if f E L
1
(G) is such that = 0, then
f = 0, where by fn we mean the n-fold convolution of f with
itself. We claim first that we may assume, without loss of generality,
that f E LI(G) n Indeed, suppose we know that f E LI(G) n LmCG)
and = 0 imply f = O. Then suppose g E LI (G) is such
that lim IIgnlll1/n = O. Let {u 1 C LI (G) n L CG) be an approximate
n a m
identity; for example, we may choose {u) as in the proof of
a
Theorem 4.7.2. From Theorem 3.5.1 we see that, for each a,
limnliCua * = 0, as Rad[L
I
(G)] is an ideal. However, by the
assumption just made and the fact that lIu * gil < lIu II IIgll1 for
a 0)- aO)
each 0, we deduce that u * g = 0 for each o. But
a
lim Uu * g - gill = 0, whence g = O. Thus to prpve the theorem we
o a 1
need only show that f E LICG) n Lm(G) and n = 0 imply
f = o.
To prove this we note first that, since f E LICG) n Lm(G), we
have f E L
2
CG). Furthermore, setting f*(t) = f(-t), we see that
f* E Ll(G) n Lm(G) whenever f E Ll(G) n Lm(G). From Example 1.2.7
and Proposition 4.7.2 we see that, if f E LICG) n Lm(G) , then
f * f* E LI(G) n LmCG), and f * f* * g E C(G) for each g E LI (G).
Consequently it is evident that F(g) = f * f* * g(O), g E LI(G),
defines a linear functional F on Ll(G), and
Next we define, : LI(G) x LI(G) - by ,(g,h) = F(g * h*),
g,h LI(G). We claim that , is a nonnegative symmetric bilinear
form on Ll(G).
118 4. The Gel'fand Representation of Specific Algebras
Indeed, if g E LI(G) , then
,(g,g) ~ F(g * g*)
= f * f* * g * g*(O)
= (f * g) * Cf* * g*)(O)
= (f * g) * (f * g)*{O)
= Ie f * g(-t)f * g(-t) dAft)
= IG If * g(_t)1
2
dA(t)
> o.
Thus V is nonnegative.
If g,h E tICG), then
yCg,h) = F(g * h*)
Hence is symmetric.
= f f* * g * h*(O)
= (f * g) * (f * h)*(O)
= iG f * g(-t)f * h(-t) d A ~ )
= IG f * g(-t)f * h(-t) dAft)
= .(h,g).
The bilinearity of is easily established, and we conclude
that is a nonnegative symmetric bilinear form on LICG). Conse-
quently the Cauchy-Schwarz Inequality is valid for , [L, p. 373];
that is,
IF(g * h*)1
2
~ F{g * g*)F(h h*)
119
Now let [ual c L
1
(G) n L.(G) be the approximate identity con-
structed in Theorem 4.7.2. Then u*; U aDd
a a
F(u * u*) = f * f* * u * u*(O)
a a' a a
< IIf * ffrU.Uu
a
*
< Hf * f*ii.
for each at as Uualll = 1. Combining this observation with the
of F and the Cauchy-Schwarz Inequality, we deduce that
Setting K = (llf f*li.)1/2 and applying this inequality to
PCg if g*) 0, we see that
IF(g)1 < K[F(g * g*)1/2]
K
1
+
l
/
2
[F(g * g* * g * g*)1
/
4
J
Continuing in this fashion, we find that
IF(g)l Sxl+l/2+".+1/2n[F([g * g*J
2n
)1/2n+l]
I /
n n / n+l
:s Kl+1 2+ ... +1 2 [lif * f*llCD
Il
(g * g*)2 Ill) 1 2
for each g E L
1
(G) and n = 0,1,2, . Thus we have
n 1/2
n
,Feg)I < 1C
3
[limll{g * g*)2 III ]1/2 (g E L1 eG).
n
In particular, if g = f * f*, then g * g* E Rad[L
1
(G)J, as
the radical is an ideal, and we conclude that
o = Fef f*)
= f * f* * f * f*(O)
120 4. The GeI'fand Representation of Specific Algebras
However, since f E L
1
(G) n Le(G) , we see from Proposition 4.7.2(i)
that f * f* E C(G), and so the preceding identity implies that
f * f*(t) = 0, t E G. Hence
o = f * f*(O) = IG If(t)12 dA(t) ,
from which we deduce that fCt) = 0 for almost all t E G; that is,
f = o.
Therefore LICG) is semisimple.
Some rather easy corollaries of this theorem are worth mention-
ing explicitly. The first is just a rephrasing of the theorem.
Corollary 4.7.2. Let
gical group. If f E LI(G)
G be a locally compact Abelian topolo-
.. ..
and fCY) = 0, y E G, then f = o.
Corollary 4.7.3. Let G be a locally compact Abelian topolo-
..
gical group. Then G separates the points of G .
..
Proof. Suppose G does not separate the points of G; that
is, suppose there exist t,s (G, t s, such that (t,y) = (s,y),
.. ..
y E G. Clearly then (t - s,Y) = 1, y E G.
But then for any f E LI(G) we would have
..
[Tt_s(f)] (y) = fG Tt_s(f)(u) (u,y) dA(u)
= IG f[u - (t - s)](u,y) dA(u)
= IG f(u)(u + t - s,y) dA(u)
= (t s,y) J
G
feu) (u,Y) dA(u)
.. ..
= f (y) (y E G).
Since LI(G)
fELl CG),
A
121
is semisimple, this would imply that Tt_s(f) = f,
which is obviously false if t s.
Thus G separates the points of G.
A
o
The set of continuous characters G can also be given a group
A
structure under which it is an Abelian group. Namely, if y,w E G,
one defines (t,y + w) = (t,y)(t,w) and (t,-y) = (-t,y), t ( G.
A
It is easily seen that with these operations G does indeed become
an Abelian group. The details are left to the reader. The identity
A
element of the group 0 is the continuous character Yo such that
y (t) = 1, t E G. We shall always use y to denote this element
o A 0
of G.
A
Considered as the maximal ideal space of LI(G), G is endowed
with a topology -- namely, the Gel'fand topology -- and a natural
A
question to ask is whether the Gel'fand topology on G and the group
A A
structure of G combine to make G a topological group. That this
is indeed the case will be seen in the corollary to Theorem 4.7.5.
A
First, however, we wish to use the group structure of G to derive
two further corollaries of Theorem 4.7.4.
Corollary 4.7.4. Let G be a compact Abelian topological group
and suppose Haar measure A on G is normalized so that A(G) = 1.
A
Then G is a complete orthonormal set in the Hilbert space L
2
CG).
A A
Proof. Clearly G C L
2
(G) , as G C C(G) and G is compact.
A
If y = Yo E a, then
IG (t,y) dAft) = ACG) = 1,
A
whereas, if yEO and y Yo' then there exists some s E G such
that (s,y) 1. Hence, since A is translation invariant, we see
that
IG (t,y) dAft) = IG (t + s,y) dAft)
= (S,y)J
G
(t,y) dA(t),
122 4. The Gel'fand Representation of Specific Algebras
which can clearly be valid only if

Consequently, if y,w E G, we see that
IG (t,y)(t,w) dA(t) = IG (t,y - w) dA(t)
= I for y = W
= 0 for y w

Thus G is an orthonormal set in L
2
(G) .

To show that a is complete it suffices to prove that, if
f E L
2
(G) and
fG f(t) (t,y) dA(t) = 0
"
(y E G),
then f = 0 [L, p. 398]. But the left-hand side of this equation

is just fey) since L
2
(G) C LI(G), as G is compact. The desired
conclusion now follows immediately from the semisimplicity of LI(G).O
In general, when G is a compact Abelian topological group,
we shall always assume that Haar measure A is normalized so that
A(G) = 1.
Cerellary 4.7.5. Let G be a locally compact Abelian topolo-

gical group. If G is discrete, then G is compact, and if G is
"
compact, then G is discrete.
Proof. If G is discrete, then, by Theorem 4.1.1, has

an identity, whence a is compact by Theorem 3.2.2. On the other

is compact, then G C LI(G).
"
hand, if G Hence, by Corollary 4.1.4,
we deduce that, for each w E G,
"
w(y) = IG (t,w)(t,y) dA(t) = I for y = w
"
= 0 for y w
(y E ct).
123

Thus the characteristic function of each singleton set (w), w E G,

is a continuous finction on G, and so (w) is an open set for

each w E G.

Therefore G is discrete.
The converse of the results of Corollary 4.1.5 is also valid,
and we shall come back to this point in Sections 1.3 and 10.5.
o
Now let us return to the question raised before Corollary 4.1.4:

Is the algebraic group G a topological group in the Gel'fand topo-
logy? Since the Geltfand topology is Hausdorff, we need only show
that the group operations are continuous in the Gel'fand topology in
order to obtain an affirmative answer. To accomplish this we shall
first describe an alternative description of the Gel'fand topology

on G which is of considerable interest in itself. The fact that

G with the Gel'fand topology is a locally compact Abelian topologi-
cal group will be an easy consequence of this description.
Theorem 4.7.5. Let G be a locally compact Abelian topological
group. For each a > 0 and each compact set KeG define

U(K,a) = (y lyE G, I(t,y) - 11 < a, t E K).
Then the family of sets (U(K,e)), where e > 0 is arbitrary and
KeG is an arbitrary compact set, forms a neighborhood base at

y E G for the Gel'fand topology on G.
o
Proof. It suffices to show that every such set U(K,a) is

open in the Gel'fand topology on G and that every open neighbor-
hood of y in the Gel'fand topology contains some such set.
o

We first observe that the function from G X G - f, defined by

(t,y) - (t,y) = yet), t E G, yEO, is continuous. Here (t,y)

denotes a point in the product space G X G, and G X G is given

the product topology. Indeed, suppose (t) C G and ~ ) C G
~ a
are nets that converge to t (G and y E OJ respectively. If
124 4. The Gel'fand Representation of Specific Algebras
A A A
f (Ll(G) is such that fCy) ~ 0, then, since f ( C (0), there
A 0
exists some et
o
such that a > et
o
computations reveal that for each
implies f y ~ ~ O. Elementary
s E G we have
A A
(s,y)f(w) = [T (f)] (w)
-s
A
(w E G),
whence, recalling the development preceding Theorem 4.7.3, we con-
clude that
A
(s,y)
=
[T -s (f)] (y)
(s ( G)
A
fey)
and
A
(s'YJ
[T -s (f)] - Y ~
(s E G; a > a ).
=

f(YJ
0
Clearly then to show that lim (t ,y ) = (t,y)
a a.a
we need only prove

that lima[T_ta(f)] (Ya) = [T_t(f)] (y) since limaf(Ya) = fey).
But
A
+ I [T_t(f)] (Ya) - [T_t(f)] (y)1
< IIT_
t
(f) - T_
t
(f)1I
1
a
from which we at once deduce the desired conclusion via Proposition

4.7.I(ii). Thus the mapping (t,y) ~ (t,y) from G X G to r is
continuous.
Now suppose s > 0 and KeG is compact, and consider U(K,s).
If w E U(K,s), then from the continuity of the function (t,y) ~ (t,y)
A
t E G,y E G, and the fact that I(t,w) - 11 < s, t E K, we deduce
that, for each t E K, there exists an open neighborhood V
t
of t
125
,.
in G and an open neighborhood W of w in G such that, if
u" t
(s,Y) V
t
x Ww,t' then I(s,y) - 11 < i. The sets {V
t
I t E K)
clearly form an open covering of the compact set K, and so there
exist t
l
,t
2
, .. ,t
n
in K for which K C Let W = nk=IWw,tk.
Evidently W is an open neighborhood of w in the Gel'fand topology,
and if t K, then - 11 < i, y (W; that is, We U(K,i).
Since E U(K,s) is arbitrary, we conclude that U(K,e) is open
in the Gel'fand topology.
On the other hand, suppose that W is an open neighborhood of
Yo in the Gel'fand topology. since the GeI'fand topology is
the relative weak* topology on A(L
1
(G)), we may assume without
loss of generality that there exist f
l
,f
2
, ,f
n
in LI(G) and
> 0 such that
,.,. ,.
{y I y.E. G, Ifk(y) - fk(yo)1 < e, k = 1,2" . ,n).
Let gk ec(G) be chosen so that llfk - gklll < e/3, k = l,2" ... n,
and consider
If yEW. then
o
whence W c w.
o
,. ,.
+ Igk(y
o
) - fk(y
o
) I
,. ,.
II
f
k
- gkll l + Igk(Y) - gk(Y
o
) I
+ lig
k
- fklll
e e e
<-+-+-
333
= (k = 1,2, ... ,n),
126
4. The Gel'fand Representation of Specific Algebras
Now let KeG be any compact set such that gk(t) = 0, t t K,
k = 1,2, ... ,n. For example, K can be taken to be the union of the
"
supports of gk' k = 1,2, .. ,n. Then for each y G we have
19k(y) - gk(YO) 1 = lJ
G
gk(t) (t,y) dAft) - SG gk(t) (t,yo) dA(t)l
~ I K 19k(t)ll(t,y) - 11 dA(t)
< sup l(t,y) - llllgklli (k = 1,2,. .. ,n).
tEK
Consequently, if e' = e/(6max k=I,2, ... ,n!lgklll) and y E U(K,e'),
then
and so U (K, e') ewe '\'.
o
Therefore the sets U(K,e), > 0 and KeG compact, form
a neighborhood base at Yo for the Gel'fand topology. 0
"
A moment's reflection reveals that the Gel'fand topology on G
is that topology in which convergence of a net of continuous charac-
ters of G is equivalent to the convergence of the net of continuous
characters, qua functions, on compact subsets of G. It is often
"
useful to think of the Gel'fand topology on G in this way.
"
The promised result about G can now be proved.
Corollary 4.7.6. Let G be a locally compact Abelian topolo-
"
gical group. Then G with the Gel'fand topology and the group
operations
(t,y + ~ = (t,y)(t,w),
"
Ct,-y) = (-t,y)
(t E G; y,w G),
is a locally compact Abelian topological group.
4.8. M(G)
127

Proof. Since G is a locally compact Hausdorff topological
space in the Gel'fand topology, we need only show that the group
operations are continuous with respect to this topology. This is
easily verified by using the description of the Gel'fand topology
provided by Theorem 4.7.5 and observing that
I(t,y + w) - 11 < I(t,y + w) - (t,y)1 + I(t,y) - 11
= I (t ,y) (t,w) (t,y)1 + I(t,y) - 11

= ICt,w) - 11 + I(t,y) - 11
(t E G; y,w E G)
and that

I(t,-y) - 11 = I(t,y) 11
= I(t,y) - 11
(t E G; yEO).
The details are left to the reader.
[j

When G is considered as a locally compact Abelian topological
group, we shall call it the dual group of G.
.. ..
Following Theorem 4.7.3 we observed that ~ = ~ r = Z, and
- .
L = f. The reader should verify that the Gel'fand topology on ~

f, and L coincides with the usual topology on F ~ Z, and r,
respectively. Thus as dual groups of ~ r, and Z we see that
... .
lR = IF\, r = Z, and Z = r.
4.8. M(G). As indicated in Example 1.2.8, the space M{G) of
bounded regular complex-valued Borel measures on a locally compact
Abelian topological group G is a commutative Banach algebra with
identity. The algebra operation is, of course, convolution of
measures. Consequently, by Theorem 3.2.2, ~ M G ) ) is a compact

Hausdorff topological space. If y E G, then it is easily verified
that
128 4. The Gel'fand Representation of Specific Algebras
E M(G))
"
defines an element T E aCMCG)), and so G can be identified with
a subset of a (M (G) )

In the case that G is discrete it is evident that C = a(M(G)),
"
as M(G) = LI(G).
subset of a(M(G))
But if G is not discrete, then G is a proper
"
and, moreover, is not dense in a(M(G)). We
shall not prove these results, but cite them only to illustrate
that the structure of the maximal ideal space can be more complicated
than the discussion of the preceding sections might indicate. For
some further discussions of these results and the maximal ideal space
of M(G) the reader is referred to [DkRa, pp. 1-28; pp. 176-192;
H, p. 143; HR
l
, pp. 368 and 369; Ri, pp. 328-330].
Nevertheless, M(G) is still a semisimple algebra, and indeed
it can be shown, as we shall do in Section 10.5, that, if E M{G)
and
"
= IG (t,y) = 0
"
(y E G),
" "
then = O. The function on G defined here is casily seen
"
to be a bounded uniformly continuous function on 0 and it is gener-
ally called the Fourier-Stieltjes transform of
CHAPTER 5
SEMISIMPLE COMMUTATIVE BANACH ALGEBRAS
5.0. Introduction. In this chapter we shall concentrate our
attention mainly on various semisimple commutative Banach algebras.
First we shall obtain some results concerning the case in which the
Gel'fand transformation of a semisimple commutative Banach algebra
is either a topological isomorphism or an The chief tool
in obtaining these results will be a theorem asserting that the
..
spectral radius IIxll = IIxll defines an equivalent norm on a semi-
a CD
simple commutative Banach algebra A if and only if IIxll
2
< Kllx
2
11,
x E A, for some constant K > O.
In Section 5.2 we shall see how to make the Gel'fand representa-
..
tion A of a semisimple commutative Banach algebra A into a Banach
algebra that is isometrically isomorphic to A and has the same
maximal ideal space as A.
Following this we shall discuss several results about homomor-
phisms and isomorphisms between commutative Banach algebras. The
most interesting of these results asserts that an algebra homomor-
phism of a commutative Banach algebra into a semisimple commutative
Banach algebra is automatically An easy corollary of
this theorem will show that the norm in a semisimple commutative
Banach algebra is essentially unique.
Finally, we shall employ some of these results, together with
our knowledge of topological zero divisors garnered in Section 1.6,
to show that in a self-adjoint semisimple commutative Banach algebra
with identity an element is singular if and only if it is a topolo-
gical zero divisor.
129
130
s. Semisimple Commutative Banach Algebras
5.1. The Gel'fand Representation of Semisimple Commutative
Banach Algebras. From the definition of semisimplicity we know that
the Gel'fand transformation of a semisimple commutative Banach alge-
bra A is injective, and so the Gel'fand representation is a sub-
algebra of C [aCA)] that is isomorphic to A. Furthermore, by the
o
Beurling-Gel'fand Theorem (Corollary 3.4.1), we see that, if A is
a semisimple commutative Banach algebra, then the spectral radius
..
IIxllo = IIxll of x in A defines a norm 1111 on A, called the
a
spectral radius norm. It is evident that A is a normed algebra
with the spectral radius norm and that IIxll < IIxli, x E A. However,
0-
(A,IIlt
o
) need not be a complete normed linear space; that is,
(A,IIlI
a
) need not be a Banach algebra. Obviously, if 11.11
0
and
1111 were equivalent norms on A (that is, if there existed some
K > 0 such that IIxll Kllxllo' x E A), then {A, II "
0
) would also
be a Banach algebra, and our first theorem provides a necessary and
sufficient condition for this. We shall then use this result to
derive conditions under which the Gel'fand transformation is a topo-
logical isomorphism. Similar arguments will give us conditions
under which the Gel'fand transformation is an isometry.
Theorem 5.1.1. Let A be a semisimple commutative Banach
algebra. Then the following are equivalent:
(i) 1111 and 11.11
0
are equivalent norms on A.
(ii) There exists some constant K > 0 such that IIxll
2
< Kllx
2
11,
x A.
Proof. If 1111 and 11.11
0
are equivalent norms, then there
exists some Kl > 0 such that IIxli < Klllxllo' x E A. Consequently
On setting
part (ii).
(x A).
we see that part (i) of the theorem implies
Conversely, suppose lIxU
2
Kllx
2
U, x A, for some K > o.
5.1. The Ge1'fand Representation 131
Then applying this estimate repeatedly, we see that, for each x E A,
IIxll KI/2I1x
2
111/2
< Kl/2[Kl/2UX4111/2] 1/2
222
= Kl/2+l/2 IIx
2
111/2
K
l / 2+ 1/22 + ... + 1/
2n
ll 2nlll/2n
< .. < x
- -
Cn = 1,2,3, ).
Thus, by the Beurling-Gel'fand Theorem (Corollary 3.4.1) or the
Spectral Radius Formula (Theorem 2.2.2), we deduce that IIxll < KUxll ,
- a
x E A. This estimate, combined with the fact that IIxU
a
< IIxll,
x E A, shows that 1111 and 1I'lI
a
are equivalent norms. Hence
part (ii) implies part (i).
o
The first application of this theorem is the next corollary.
Recall that a mapping is said to be a topological isomorphism if it
is both an isomorphism and a homeomorphism.
Corollary 5.1.1. Let A be a semisimple commutative Banach
algebra. Then the following are equivalent:

(i) The Gel'fand representation A of A is a closed subal-
gebra of (Co[fl(A)],UU
m
).
(ii) There exists a constant K > 0 such that IIxU
2
Kllx
2
U.
x E A.
(iii) The Gel'fand transformation is a topological isomorphism

of A onto A.
Proof. If part (i) holds, then the Gel'fand transformation is
evidently a continuous bijective linear mapping of A onto the

Banach space A C C [fl(A)]. Thus an application of the Open Mapping
o
Theorem [L, p. 187] reveals that the inverse of the Gel'fand trans-

formation is a continuous linear mapping from A to A, and so

there exists some Kl > 0 for which IIxll K111xllm = K1llxlla. x E A.
132
S. Semisimple Commutative Banach Algebras
Consequently 1111 and 11.11
0
are equivalent norms on A, whence,
by Theorem 5.1.1, we conclude that there exists some K > 0 such
that IIxll
2
< Kllx
2
l1, x 'A. Thus part (i) implies part (ii).
If part (ii) holds, then Theorem 5.1.1 shows that 1111 and
11.11
0
are equivalent norms, from which it follows immediately that
the Gel'fand transformation is a topological isomorphism, as

IIxlia = lIxllm, x E A. Hence part (ii) of the corollary implies part
(iii).
To show that part (iii) implies part (i) is even easier and is
left to the reader.
o
Before we can state the next corollary we need a definition.
Definition 5.1.1. Let A be a commutative Banach algebra.

A is said to be self-adjoint if whenever x E A then x E A,
where the bar denotes complex conjugation.
Thus A is self-adjoint if and only if the Gel'fand representa-

tion A is closed under complex conjugation. It is left as an exer-
cise for the reader to verify that the Banach algebras C (X), X
o
being a locally compact Hausdorff topological space, and Ll(G),
G being a locally compact Abelian topological group, are self-adjoint,
whereas A(D) is not.
It should be observed if A is a self-aajoint commutative
Banach algebra, then an appeal to the Stone-'Veierstrass Theorem

[L, p. 332] shows that the Gel'fand representation A is dense in

(Co , I; ltD). In particular, Ll (G) is dense in Co (G) , where
G is a locally compact Abelian topological group.
The proof of the next corollary utilizes the general fact about
self-adjoint algebras just mentioned. The details are left to the
reader.
..
S.2. A as a Banach Algebra
133
Corollary S.1.2. Let A be a self-adjoint semisimple commuta-
tive Banach algebra. Then the following are equivalent:
(i) There exists a constant K > 0 such that Ilxll
2
< Kllx
2
11,
x A.
..
Cii) A = C [deAl].
o
The final result of this section gives a necessary and suffi-
cient condition for the Gel'fand transformation to be an isometry.
Theorem S.1.2. Let A be a commutative Banach algebra. Then
the following are equivalent:
- (ii) The Gel'fand transformation is an isometry of A onto A.
Proof. If IIxU
2
= IIx
2
U, x E A, then clearly for each x E A
I
2n I 2
n
we would have Ix I = IIxll , n = 2 ~ . 3 .... Hence, by the 8eurling-
Gel'fand Theorem (Corollary 3.4.1), we deduce that
2
n
1/2
n
..
IIxll = limUx II = IIxll
G)
(x E A),
n
from which it is apparent that the Gel'fand transformation is an
isometry.
The converse is immediate on noting that
2 -2 .. 2 2
IIx II = IIx II = IIxll = IIxll
CD CD
(x E A)
whenever the Gel'fand transformation is an isometry.
c
It is worth remarking that, if the Gel'fand transformation is
- an isometry, then A must be semisimple and A must be a closed
subalgebra of C [A(A)).
o
..
5.2. A ~ ~ Banach Algebra. If A is a commutative Banach
algebra, then the Gel'fand Representation Theorem (Theorem 3.3.1)
134
s. Semisimple Commutative Banach Algebras

shows that A is a normed subalgebra of C [A(A)]. It need not,
o
however. be a Banach algebra with ~ h supremum norm. Nevertheless,

if A is semisimple, A can be normed in such a way that it does
become a Banach algebra. This technique is often useful in studying
various problems.
Theorem 5.2.1. Let A be a semisimple commutative Banach alge-

bra and define IIxli. = IIxll, x E A. Then
(iJ
1111.

is a norm on A.

(ii) (A,IIII.) is a semisimple commutative Banach algebra .

(iii) A (A) is homeomorphic to A(A).
(iv) The Gel'fand transformation is an isometric isomorphism of

A onto (A,IIU.).

Proof. The verification that
nlI.
is a norm on A and that
(A,IIII.) is a commutative Banach algebra is routine and is left to
the reader. The semisimplicity of A is required to show that IIU.
is well defined.

If T E A(A), define w
T
on A by w (x)
T
= T(X), x E A .

Clearly w
T
is a complex homomorphism of A; that is,

wTEA(A).
Moreover, the mapping from A(A) to A (A) so defined is bijective.
Indeed, if w
T
= W ,
1 T2
which implies that TI
(Theorem 3.3.1). Thus

then X(T
I
) = TI(X) = T
2
(X) = X(T
2
), x E A,

= T
2
, as A separates the points of A (A)
the mapping is injective.

On the other hand, if w E A(A) and one defines T(X) = w(x),

x E A, then obviously T E A(A) and wT(x) = T(X) = w(x), x E A.
Hence the mapping is surjective.
Furthermore. if (T J C A(A) is
a
then, since A C C [A(A)], we see at once that
.0. .
a net that converges to T E A(A),

{w (x)] converges
Ta
to wT(x) for each x E A. from which

of the Gel'fand topology on A (A) that (w
T
] converges to w
T

a
it f o l l o ~ s by the definition
5.3. Homomorphisms and Isomorphisms
135
..
Consequently the mapping T w
T
of 6(A) onto 6(A) is continuous.
A similar argument proves the continuity of the inverse mapping .
..
Therefore 6(A) and 6(A) are homeomorphic
..
The semisimplicity of (A,IIII .. ) and part Civ) of the theorem
are now apparent.
o
One of the advantages of this construction is that it enables
us to replace a semisimple commutative Banach algebra A by another
such algebra B, where the algebra B is an algebra of continuous
functions under pointwise operations that is isometrically isomor-
phic to A and has the same maximal ideal space as A. The algebra
..
B, as we have just seen, is B = A with the norm 1111.. Thus, for
example, if A = Ll(G), G being a locally compact Abelian topologi-
cal group, then LI(G) is isometrically isomorphic to the semisimple
.. ..
commutative Banach algebra L1CG) of functions in C CG) with the
.. 0
norm IIfll .. = IIf1l1, f (; LI(G). A special case of this was mentioned
.. ..
in Example 1.2.9 when G = f and G = Z. There we denoted LICf)
..
by FLICf). Similarly in this case we may write LICG) = FL1CG)
when G is an arbitrary locally compact Abelian topological group.
5.3. Homomorphisms and Isomorphisms of Commutative Banach
Algebras. In this section we wish to prove several results concern-
ing homomorphisms and isomorphisms between commutative Banach algebras.
A corollary of the first of these results, which asserts that an
algebra homomorphism of any commutative Banach algebra into a semi-
simple commutative Banach algebra is continuous, will show that the
norm on a semisimple commutative Banach algebra is essentially unique.
This assertion will be made precise in Corollary 5.3.1.
First, however, we prove the indicated theorem.
Theorem 5.3.1. Let A and B be commutative Banach algebras
and suppose A is semisimple. If T : B A is an algebra homo-
morphism, then T is continuous.
136 S. Semisimple Commutative Banach Algebras
Proof. Since T is a linear mapping and A and Bare
Banach spaces, we may appeal to the Closed Graph Theorem [L, p. 189]
to deduce the continuity of T. To this end, suppose {x
k
) is a
sequence in B and x E Band yEA are such that
and
lim/lT(X
k
) - yllA = 0,
k
where II'II
B
and U'II
A
denote the norms in B and A, respectively.
We must show that T(x) = y.
Let T E ~ A ) and set w
T
= ToT. Since T is a homomorphism,
it is easily seen that either w
T
= 0 or w
T
E ~ B ) . In either
case we have limkw,,(x
k
) = wT(x). lienee we see, on the one hand,
that for each T E A(A)
"
= T(x) (T)
and, on the other hand,
" "
lim T(x
k
) (T) = yeT),
k
" "
as
Thus T(x) = y, whence, from the semi-
simplicity of A, we conclude that T(x) = y.
Therefore, T is a closed linear mapping and hence, by the
Closed Graph Theorem [L, p. 189], a continuous mapping.
c
Corollary 5.3.1. Let A be a commutative Banach algebra under
the norms 11'11
1
and 11'11
2
. If (A,IIII
I
) is semisimple, then 11'11
1
and 1i'11
2
are equivalent norms on A.
Proof. Consider the mapping T : (A,II'1I
2
) - (A,li'H
I
), defined
by T(x] = x, x E A. F.vidently T is an algebra homomorphism and,
by Theorem 5.3.1, is continuous. Thus there exists some K > 0 such
5.3. Homomorphisms and Isomorphisms 137
that IIxlll < Kllxll2, x A.
are both Banach spaces, this
[L, p. 190], that 11.11
1
and
However, since (A,IIlI
l
) and (A, 11.11
2
)
implies, via the Two Norm Theorem
11.11
2
are equivalent. 0
Thus the meaning of our previous remark concerning the essential
uniqueness of the norm in a semisimple commutative Banach algebra
should now be clear. Corollary 5.3.1 is actually valid even if the
assumption of commutativity is dropped, but the proof requires more
of the general theory of Banach algebras than we have at our dis-
posal. A proof of this result can be found in [Jo].
In Theorem 4.1.4 we proved that two compact Hausderff topologi-
cal spaces X and Yare homeomorphic if and only if C(X) and
C(Y) are algebraically isomorphic. Our next theorem shows that
of this equivalence is valid in general for commutative
Banach algebras.
Theorem 5.3.2. Let A and B be eommutative Banach algebras.
If there exists an algebra isomorphism of B onto A, then 6(A)
is homeomorphic to 6(B).
Proof. Suppose T: B A is a surjective algebra isomorphism.
As before, for each T E 6(A) we define w
T
E 6(B) by
WT(X) = (T 0 T)(x) = T[T(x)] (x E B).
Note that w
T
0, because if wT(x) = T(T(x)] = 0, x E B, then
T = 0, as T is surjective, contradicting the fact that T E 6(A).
Evidently the mapping = w
T
' T E 6(A), maps 6 (A) into 6(B).
r.toreover, is bijective.
Indeed, if = then
whence we conclude that Tl = T
2
, as
w E 6(B), then, as above, we see that
Tl[T(x)] = T
2
[T(x)], x E B,
T is surjective; and if
T (y) = W[T-l(y)], yEA,
W
defines an element T E A(A). Clearly = w, and so is
UI W
bijective.
138 5. Semisimple Commutative Banach Algebras
-1
Furthermore, and are continuous. For instance, suppose
(T ) C A(A) is a net that converges to T E A(A). Then, since
a
A C Co[A(A)], we see that for each x E A

lim = lim T [T(x)] = lim T(x) (T ) = T(x) o(T) =
a a a a
whence we conclude that is continuous. Similarly is con-
tinuous.
Therefore A(A) A(B) is a homeomorphism.
A few remarks about this theorem are in order. First, if A
is semisimple, then, by Theorem 5.3.1, T is continuous, and T

o
is completely determined by the equation T(K) (T) = where
T E A(A) and x E B, and is the homeomorphism defined in the
proof of the theorem. Thus, if Tl and T2 were surjective alge-
bra isomorphisms from B to A which defined the same homeomorphism
then Tl = T
2
" This follows at once from the semisimplicity of

A and the equations Tl(x) (TJ = = T
2
(x) (T), T E A(A) and
x E B.
Contrary to the situation for C(X) discussed earlier, the
converse of Theorem 5.3.2 need not be valid. In particular, if
: A(A) ACB) is a homeomorphism between A(A) and A(B), then

T : B A, defined by T(x) (T) = T E A(A) and x E B,
need not define a surjective algebra isomorphism. The crux of the
difficulty is that, given a homeomorphism A(A) - A(B), it is

not at all evident that x 0 (A when x E B. The question of
precisely which homeoMorphisms between A(A) and ACB) induce
isomorphisms between B and A is generally a rather intricate
one, even in the case that A = B. For some specific algebras the
answers are known, for instance, when A = C(X) and B = Cry), X
and Y being compact Hausdorff topological spaces. We shall not
discuss the problem any further, but content ourselves describ-
ing the situation for LIOR) " Recall that A(LIOR =
5.4. Characterization of Singular Elements
Theorem 5.3.3. If T is a mapping from LIOR) to itseff,
then the following are equivalent:
139
(iJ T Ll (lH) - Ll (lR) is a surjective algebra isomorphism.
(ii) ~ R-R is a homeomorphism of R onto itself of the
form C9(t) = at + b, t E IR, for some a,b E IR, a ~ 0, and
.. ..
T(f) (t) = f[C9(t)], t E Rand fELl (lR).
A proof of this result can be found in [Ka, pp. 217-219]. Some
further discussion of such isomorphism problems is available in
[Ru
l
, pp. 77-96].
5.4. Characterization of Singular Elements in Self-adjoint
Semisimple Commutative Banach Algebras. In our earlier discussion
of topological zero divisors we proved (Proposition 1.6.2) that, if
x is a topological zero divisor in a commutative Banach algebra
A with identity, then x is singular. For the special algebra
C (X), X being a compact Hausdorff topological space, we sa\.; in
Theorem 1.6.1 that f E C(X) is a topological zero divisor if and
only if f is singular. For an arbitrary commutative Banach algebra
A with identity it need not be the case that a singular element is
a topological zero divisor, but if A is self-adjoint and semi-
simple, then it is true. We shall prove this in the theorem of
this sectioR.
Before we can do this, however, we need some preliminary re-
sults about self-adjoint semisimple commutative Banach algebras.
Suppose A is such an algebra. Given x A, we know that there
.. ..
exists some yEA such that y = x. Since A is semisimple, this
element y is clearly unique. ~ h s observations lead to the follow-
ing definition:
Definition 5.4.1. Let A be a self-adjoint semisimple commu-
tative Banach algebra. If x ~ A, then we denote by x* the uni-
.. ...
que element of A such that x* = x.
140
5. Semisimp1e Commutative Banach Algebras
We shall need the next proposition.
Proposition 5.4.1. Let A be a self-adjoint semisimple commu-
tative Banach algebra. Then
(i) (x + y)* = x* + y*, (xy)* = x*y*, and (ax)* = ax*,
x,y (A and a ~
(ii) The mapping * : A - A, defined by * : x - x*, x E A,
is an antilinear topological isomorphism of A onto itself.
Proof. Part (i) and the fact that * is a surjective anti-
linear isomorphism are easily verified. We shall prove explicitly
only that * is a topological isomorphism. To this end we intro-
duce a new norm in A by defining IIxlll = IIx*lI, x (A. It is a
routine exercise to check that 11'11
1
is indeed a norm. _For example,

if IIxlll = 0, then IIx*1I = 0, and so x* = O. Hence x = 0, which

implies that X = 0, and so x = 0, as A is semisimple. More-
over, 11'11
1
is a complete norm on A.
Indeed, suppose {xnJ is a Cauchy sequence in (A, UI;l)' Then
evidently {x*) is a Cauchy sequence in (A,II'II), and so there
n
exists some yEA such that lim IIx* - yll = O. But then, if
n n
x = y*, we see that
limllx
n
- xIII =
limllx*
x"l1
n
n n
=
limllx* - yll
n
n
= o.
Thus (A,II'II
I
)
is a Banach space.
Consequently we see that (A, Ii '11
1
)
commutative Banach algebras and (A, Ii 'Ii)
conclude, by Corollary 5.3.1, that 11'11
1
and
(A,II'II)
are both
norms. Thus, in particular, there
that K2lixli < IIx* II ~ Klllxll, x E A,
isomorphism.
exist
and so
is semisimple, whence we
and 1111 are equivalent
Kl > 0 and K2 > 0 such
* is a topological
c
5.4. Characterization of Singular Elements 141
One further observation is in order before we can state and
prove the indicated theorem. If A is a commutative Banach algebra,
then x A is a topological zero divisor if and only if there
exists a sequence {y} C A such that inf lIy II > 0 and
n n n
lim lixy II = O. The proof of this assertion is left to the reader.
n n
Theorem 5.4.1. Let A be a self-adjoint semisimple commuta-
tive Banach algebra with identity e. If x E A, then the follow-
ing are equivalent:
(i) x is a topological zero divisor.
(ii) x is singular.
Proof. The implication from part (i) to part (ii) is contained
in Proposition 1.6.2. Conversely, suppose x E A is singular.
A
Then _0 ( o(x) = R(x), by Theorem 3.4.1. Set y = xx*. Clearly
1 = = and so 0 C R(y) and R(y) C Thus we see
that -lIn oCy), n = 1,2,3, ... , so that y + (l/n)e is regular,
n = 1,2,3,.... Since lim liy + (l/n)e - yli = 0, we conclude that
n
y (bdy(A-
1
), and so, by Corollary 1.6.1, y is a topological
zero divisor. Hence there exists some sequence (y) C A such that
n
inf Ily II > 0 and 1 im lIyy II = 1 im Iixx*y II = O.
n n n n n n
If inf IIx*y II > 0, then clearly x is a topological zero
n n
divisor. On the other hand, suppose inf lix*y II = O. Then there
n n
exists a subsequence (y ) of (y) such that limkllx*y II = O.
nk n nk
However, in view of Proposition 5.4.1, there exist Kl > 0 and
K2 > 0 such that K2lizll IIz*1I < KlllzlI, z E A. Consequently, we
see that lim.!Ixy* II = O. Furthermore,
k. nk
Thus, once again, we see that x is a topological zero divisor.
U
CHAPTER 6
ANALYTIC FUNCTIONS AND BANACH ALGEBRAS
6.0. Introduction. Let A be a commutative Banach algebra
with identity and suppose p is a polynomial. Then it is obvious
that p(x) makes sense for any x E A and belongs to A. Moreover,
the Polynomial Spectral Mapping Theorem (Theorem 2.2.1) entails

that p(x) (T) = p[X(T)], T E A ) . Our main concern in this chap-
ter will be to show that such a phenomenon is valid for a much lar-
ger class of functions than polynomials. Thus, we shall see that,
if x E A and f is analytic in some open set containing a(x),

then there exists some y = f(x) E A such that f(x) (T) = f[x(T)],
T e 6(A). The discussion and proof of this result will take up
Section 6.1.
The topics in the succeeding sections are essentially applica-
tions of the results of Section 6.1. Besides a generalization of
the Polynomial Spectral Mapping Theorem, we shall establish a suffi-
cient condition for a commutative Banach algebra with compact maximal
ideal space to have an identity; prove a generalization of Wiener's
Theorem (Theorem 4.6.2); show that, if f is a nonconstant entire
function and x E A is such that f(x) = 0, then there exists a
nonconstant polynomial p such that p(x) = 0; and give a descrip-
tion of the connected component of the identity in the set of regular
elements in A.
6.1. Analytic Functions of Banach Algebra Elements. Let A
be a commutative Banach algebra with identity, let xe A, and
suppose f is a complex-valued function defined and analytic on
some open set 0 ~ a(x). Our main goal in this section will be to
142
6.1. Analytic Functions 143
show how to define f(x) as an element of A such that
A A
f(x) (T) = f[x(T)], T E 6(A). We shall accomplish this through a
series of lemmas culminating in the desired result. Afterward we
shall briefly discuss, without proof, some generalizations of the
development.
To begin we first need a number of elementary definitions.
Definition 6.1.1. A omplex-valued function Crt) = +
a t b, is said to be a smooth arc if
(i) dC/dt exists and is continuous on [a,b].
(ii) and dll/dt have no common zero on [a,b].
Furthermore, C is said to be a regular curve if there exists
a = to < tl < ... < t
n
_
l
< tn = b such that C restricted to
[tk,t
k
+
l
] is a smooth arc, k = O,I,2, . ,n - 1.
In the definition of a smooth arc we mean, of course, that the
appropriate one-sided derivatives exist at t = a and t = b.
Definition 6.1.2. Let C{t), a < t b, be a regular curve.
Then C is said to be simple if there exist no points a tl < t
z
< b
such that ,(tl) = '(t
2
), and , is said to be a simple closed regular
curve if C(a) = C(b) and there exist no points a < tl < t2 < b
such that ,(tl) = C(t
2
). The disjoint union of a finite number of
simple closed regular curves is said to be a regular contour.
Thus a simple regular curve , does not cross itself, and a
simple closed regular curve meets itself only at its end points.
Now suppose A is a commutative Banach algebra with identity
and let x (A. Then o(x) is a nonempty compact subset of C,
which may, however, consist of several disjoint pieces and whose
complement need not be connected. Nevertheless, if 0 C C is an
open set such that 0 a(x), then it is not difficult to verify
144 6. Analytic Functions and Banach Algebras
that there exists a regular contour y that satisfies the following
properties, where n(Y,a) denotes the winding number of the regular
contour y with respect to the complex number a [A, pp. 114-117]:
(1) yeO - o(x).
(2) If a E o(x), then n(Y,a) > 0; that is, o(x) lies
inside the regular contour y.
(3) If a E o(x), then n(Y,a) = 1; that is, each point
a E o(X) lies inside precisely one of the simple closed regular
curves that make up y.
(4) If a 0, then n(y,a) = 0; that is, no point
a E - 0 lies inside of the regular contour y.
Definition 6.1.3. Let A be a commutative Banach algebra
with identity, x E A, and suppose 0 o(x) is an open set. If
(4) ,
O.
Y is a regular contour that satisfies properties (1) through
then Y is said to be a spectral contour for a(x) lying in
An application of the Cauchy Integral Formula [A, p. 119] imme-
diately establishes the first lemma.
Lemma 6. 1. 1.
identity and let
o a(x) and y
then
Let A be a commutative Banach algebra with
x C A. If f is analytic in some open set
is any spectral contour for a(x) lying in 0,
f(a) - _1_ '" f(,) d'"
- 2ni 'y , - a
(a E a(x)).
We wish Rext to extend this result so that we can replace a
in the above equation by x. To do this we clearly need to define
the integral
where e denotes the identity of A, and show that it belongs to A.
6.1.Analytic Functions
145
With this in mind suppose A is a commutative Banach algebra
with identity e, x E A, ~ a(x) is open, and f is analytic on
0, and let y be a spectral contour for a(x) lying in O. To
make the argument a bit simpler we shall assume for the moment that
y is a simple closed regular curve C = C(t), a <t < b. If C E y,
then C ~ a(x), and so (Ce - x)-l exists, and, moreover, the
-1
mapping C - (Ce - x) is continuous from y to A, since inver-
sion in A is a continuous operation (Corollary 1.4.3). Since y
-1
is compact, we then deduce that the mapping C - (Ce - x) , C E y,
is uniformly continuous.
Let a = to < tl < ... < t
n
_
1
< tn = b be any partition of
[a,b] and set 'k = 'Ct
k
), k = O,1,2, ... ,n. Then considering the
finite sums
we deduce, by essentially the same argument as for scalar functions
[Go, pp. 52-54], that
exists in A. The details are left to the reader. The limit so
obtained is denoted by
1 J -1
f(x) = 2ni y f(C)(Ce - x) dC.
Note that the argument works equally well for a spectral contour y
on repeating the preceding development for each of the finitely
many simple closed regular curves making up y.
These observations prove the first portion of the next lemma.
Lemma 6.1.2. Let A be a commutative Banach algebra with
identity
o a(x)
e and let x (A. If f is analytic in some open set
and y is a spectral contour for a(x) lying in 0, then
146 6. Analytic Functions and Banach Algebras
1 J -1
f(x) = 2ni y f(C)(Ce - x) dC
A A
exists and is an element of A. Moreover, f(x) (T) = f[x(T)],
T E 6(A) .
Proof. We need only prove the final assertion of the lemma.
Let T E 6(A). Then we see that, for C a(x) ,
-1
1 = T[(Ce - x)(Ce - x) ]
-1
= T(Ce - x)T[CCe - x) ]
= [C - T(X)]T[(Ce - xl-I],
-1 -1
whence we deduce that T[CCe - x) ] = [C - T(X)] . Consequently,
for any of the approximating sums to f(x), say
we have
1 n-I -1
(2ni) E f(Ck)CCke - x) (C
k
+
I
- C
k
),
k=O
1 n-l -1
T[G
2ni
) E f(C
k
) (Cke - x) (C
k
+
1
- C
k
)] =
k=O
1 n-l -1
(2ni) 1: f(C
k
) [C
k
- T(X)] eC
k
+
I
- C
k
),
k=O
whence we conclude from the definition of f(x) and the continuity
of T that
A
f(x) (T) = T[f(x)]
A
= 21. S f(C)[C - T(X)]-1 dC
n1 y
= f[TeX)]
A
= f[X(T)],
since X(T) E R(x) = a(x), by Theorem 3.4.1.
o
6.1. Analytic Functions
147
Our definition of f(x) seems to suffer from one disadvantage:
the value f(x) appears to depend on the choice of the spectral
contour y. That this disadvantage is illusory follows from the
next lemma.
Lemma 6.1.3. Let A be a commutative Banach algebra with
identity e and let x A. If f is analytic on some open set
o a(x) and Y
I
and Y2 are spectral contours for a(x) lying
in 0, then
1 J -1 1 r -1
2ni Y f(C)(Ce - x) dC = 2ni Jy f(C)(Ce ~ x) dC
I 2
Proof. Suppose x* is any continuous linear functional on A.
Then, as seen previously (for example, in the proofs of Theorems
1.5.1 and 2.1.2), we see that the mapping C - x*[(Ce - x)-l] is
analytic on C - a(x). Thus it follows at once from the same sort
of limiting argument as in Lemma 6.1.2 and the classical Cauchy
Integral Theorem [A, p. 145] that
1 J -1 1 J -1
x*[2ni Y
I
f(C)(,e - x) dC] = 2ni Y
l
f(C)x*[(Ce - x) ] dC
that
= ~ i Iy f(CJx*[(Ce - xl-I] dC
2
1 r -1
= x*[2ni &y f(C)(Ce - x) dC].
2
A consequence of the Hahn-Banach Theorem [L, p. 90] then implies
Thus we see that the definition of f(x) does not depend on
which spectral contour is utilized.
Our next lemma describes what f(x) is for some particular
choices of f.
148 6. Analytic Functions and Banach Algebras
Lemma 6.1.4. Let A be a commutative Banach algebra with
identity e and let x C A. Then
(i) If f(C) = 1, , . ~ , then f(x) = e.
Cii) If fCC) = C, C ~ , then f(x) = x.
(iii) If f(C) = E:=OanC
n
, , ~ , is an entire function, then
CD n
f(x) = Ln=Oanx, where the series in A converges absolutely in
norm.
Proof. Clearly we need only prove part (iii), and we may take
o = Ii.: => o(x). In view of Theorem 3.4.1 we see that, if r > !lxII,
then the circle y = (C I Ici = r) is a spectral contour for o(x).
Moreover, if C (V, then, by Corollary 1.4.3 and the comments
following it or by direct computation, we see that
CD
(Ce - x)-l = ,-lee _ X)-l = ,-1 1:: (x)n
C n=O ,
and that the convergence is absolute and uniform in norm on V.
Hence we deduce that
f(x)
1 . -1
= 2ni .I
y
f(C)(Ce - x) dC
CD
= 1: [ ~ J fCC) dC]x
n
n=O 2n1 V C
n
+
l
CD f(n) (0) n
= 1: I X
O
n.
n=
CD
\ n
= 1: a x ,
n=O n
by an application of the Cauchy Integral Formula [A, p. 120].
c
It is apparent from the definition of f(x) that the mapping
from f - f(x) is linear. However, more can be said, as seen in
the next lemma.
6.1. Analytic Functions 149
Lenuna 6. 1. 5 . Let A be a conunutative Banach algebra with
identity e and let x E A. If f and g are analytic on an
open set O:Ja(x), then
(i) (f + g) (x) = f(x) + g (x).
(ii) (af) (x) = af(x), a E; \l;.
(iii) (fg) (x) = f (x) g (x) .
Proof. We need only prove part (iii). Let VI and Y
2
be
distinct spectral contours for a(x) lying in 0 such that Y
I
lies inside V
2
, that is, such that n(v
2
,C) > 0, , E VI' We
note that, if z (Y
2
and C E Y
l
, then
(Ce - x)-l(ze - x)-l(z - C) = (Ce - x)-l(ze - x)-l[(ze - x) - (Ce - x)]
-1 -1
= (Ce - x) - (ze -x) .
Using this observation we see that
I J' -1 1 r -1
f(x)g(x) = [-2 . f(C) (Ce - x) dC] [2n' & Y g(z) (ze - x) dz]
TTl Y
I
1 2
= (2;i)2 SV
l
J
V
2
f(C)g(z)CCe - x)-lCze - x)-1 dz dC
-1 -1
= 1 J' J f(C) ( l[CCe - xl - (ze - x) ] dz dC
(2ni)2 Y
l
Y2 g z z - ,
- 1 S f -1[ 1 ~ g(z) d ]
- 2TTi Y
I
(C)(Ce - x) 2ni Y
2
z _ C z dC
1 J -1 1 J f(C)
- 2ni Y
2
g(z)(ze - x) [2ni Y
l
z _ CdC] dz.
Now g is analytic on and inside Y
2
, so for each C E Y
1
we have,
by the Cauchy Integral Formula [A, p. 119],
gCC) = J g(z) dz
2TTI Y
2
z - C
as Y
l
lies inside Y2' On the other hand, f(C)/(z - ') is analy-
tic on and inside VI as a function of C for each z E Y2' whence
150 6. Analytic Functions and Banach Algebras
we conclude, from the Cauchy Integral Theorem [A, p. 145], that
for each z E Y2
1 j. f(C) dC = 0
2ni Y 1 z - C
Consequently we have
I S -1
f(x)g(x) = 2ni Y f(C)g(C)(Ce - x) dC = (fg)(x),
1
and the lemma is proved. 0
The reader should note the crucial, if tacit, role plated by
Lemma 6.1.3 in the proof of the preceding result.
We can summarize the preceding development in Theorem 6.1.1.
The proof of part (iii) is left to the reader.
Definition 6.1.4. Let A be a commutative Banach algebra
with identity and let x A. The collection of all complex-valued
functions f that are defined and analytic on some open set
o a(x) will be denoted by A(x).
The open set 0 in the preceding definition may, of course,
depend on f. Evidently A(x) is an algebra under pointwise oper-
ations of addition, scalar multiplication, and multiplication.
Theorem 6.1.1. Let A be a commutative Banach algebra with
identity and let x E A. Then the mapping f - f(x), f E A(x), is
an algebra homomorphism from A(x) into A such that

(i) If f E A(x) , then f(x) (T) = f[X(T)], T E
(ii) If f(C) = C is an entire function, then
f(x) = and the convergence is absolute in norm.
(iii) Suppose (f) c A(x) and f E A(x) are analytic on an open
n
set 0 a(x). If the sequence (f) converges uniformly to f on
n
each compact subset of 0, then lim IIf (x) - f(x)1I = o.
n n
6.2. Consequences of the Preceding Section 151
Portions of this theorem can be extended to the case where a(x)
is replaced by the joint spectrum O(x
l
.x
2
x
n
) of finitely many
elements of A. We shall not give a proof of this theorem since it
involves the techniques of the theory of functions of several complex
variables, and a complete discussion would necessitate a rather
lengthy digression. We shall, however, state the indicated result.
Proofs are available. for example, in [HOI' pp. 101-114; Hr, pp. 68-70;
Ri, pp. 156-162; S. pp. 60-72].
Theorem 6.1.2 Let A be a commutative
Banach algebra with identity and let x
l
,x
2
' .. x
n
be in A. If
f is a function of n complex variables that is defined and analy-
tic on some open set 0 a(x
l
'x
2
' .. x). then there exists some
n.
yEA such that yeT) = f[x
1
(T),X
2
(T) . X
n
(T)], T E A(A).
Theorem 6.1.1 also gives a partial answer to a more general
question. If A is a commutative Banach algebra and f is defined

on some open set 0 then we say that f operates in A if

fox E A whenever x E A and R(x) c O. Theorem 6.1.1 then asserts

that functions that are analytic on open sets in operate in A
for any commutative Banach algebra A with identity. On the other
hand, if A = C(X), X being a compact Hausdorff topological space,
then it is easily verified that any continuous function defined on

an open subset of operates in A.
The question of determining precisely which functions operate

in A for a particular algebra A is in general a difficult one,
and we shall not pursue it here. The interested reader is referred
to [Ka, pp. 235-249; Ri. p. 167; Ru
l
, pp. 137-155].
6.2. Consequences of the Preceding Section. In this sec-
tion we wish to establish a number of results that are simple conse-
quences of the development in the preceding section. In particular
we shall obtain generalizations of the Polynomial Spectral Mapping
Theorem (Theorem 2.2.1) and Wiener's Theorem (Theorem 4.6.2).
152 6. Analytic Functions and Banach Algebras
Theorem 6.2.1 (Spectral Mapping Theorem). Let A be a commu-
tative Banach algebra, let x E A, and suppose f is a complex-
valued function that is defined and analytic on some open set
a(x). Then
(i) If A has an identity, then f[a(x)] = a(f(x.
(ii) If A is without identity and f(O) = 0, then there
exists some f(x) E A such that f[a(x)] = a(f(x.
Proof. Part (i) follows immediately from Theorem 6.l.l(i) on

recalling that a(x) = R(x) and a(f(x = R[f(x)] A has
an identity.
To establish part (ii) we need first to prove the existence of
a suitable element f(x) E A. The development of the preceding sec-
tion is not directly applicable, as A is without identity. However,
by Theorem 2.1.1, we know that a(x) = aA(x) = aA[e] (x), so we can
apply Theorem 6.1.1 to the algebra A[e] to deduce the existence

of an f(x) E A[e] such that f(x) (T) = f[X(T)], T E
However, if T E denotes,as usual, the unique complex
CD
homomorphism on A[e] that vanishes identically on A, we see that

f(x) (T
CD
) = f[X(T
CD
)] = f(O) = 0 by the assumptions on f. Hence
f(x) E A.

The definition of f(x) also reveals that f(x) (T) = f[x(T)],
T E which, combined with the facts that f(O) = 0,

a(x) = R(x) U {oj, and a(f(x = R[f(x) ] U {oj, implies that
f[a(x)] = a(f(x.
o
A similar result is the following corollary:
Corollary 6.2.1. Let
suppose is compact.
and f is a complex-valued

on some open set 0 R(x),

yeT) = f[X(T)], T E A(A).
A be a commutative Banach algebra and

If x E A is such that X(T) 0, T E
function that is defined and analytic
then there exists some yEA such that
6.2. Consequences of the Preceding Section 153
Proof. If A has an identity, then is automatically
compact, and the corollary follows at once from either Theorem 6.l.l(i)
,.
or 6.2.1. So assume that A is without identity. Since x is
,.
continuous, we see that R(x) is a compact subset of that does
not contain zero. Thus there evidently exist open sets 01 and 02
in q; such that
,.
(a) R(x) C 01 C 0.
(b) 0 02.
(c) 01 n 02 =
Having chosen such sets, define
g(C) = f(C) for C 01'
g(C) = 0 for C E 02.
Clearly g is analytic on the open set 01 U 02 a(x), and
g(O) = O.
Consequently, by Theorem 6.2.l(ii), there exists some y = g(x)
,. ,.
in A such that yeT) = g[X(T)), T E from which it follows
,.,. ,.
at once that yeT) = f[X(T)), T E as R(x) cOl 0
As we know from Theorem 3.2.2, if A is a commutative Banach
algebra with identity, then is compact. It is natural to
ask or not the converse assertion is valid. Some support
for such a conjecture is given by the next result.
CorGllary 6.2.2. Let A be a semisimple commutative Banach
algebra. If is compact and there exists some x E A such
,.
that X(T) 0, T E A(A), then A has an identity.
,.
Proof. Since R(x) is a compact set in that does not con-
"
tain zero, there exists an open set R(x) such that 0 o.
Then f(,) = I'" , E 0, is analytic on 0, and so, by Corollary
154 6. Analytic Functions and Banach Algebras

6.2.1, there exists some yEA such that yeT) = f[X(T)] = l/X{T).
T ~ A(A). It is then immediate from the semisimplicity of A that
xy is an identity for A.
o
Actually. one can prove that, if A is a semisimple commutative
Banach algebra and A(A) is compact, then A has an identity. The
proof. however, requires the use of the theory of functions of sever-
al complex variables and will not be discussed here. The interested
reader is referred to [Ri, pp. 167-169; 5, pp. 75-77]. An appropri-
ate analog of the ilov-Arens-Calderon Theorem (Theorem 6.1.2) for
algebras without identity and the ilov Idempotent Theorem [Ri, p. 168;
5, p. 73] are the main tools needed to establish the indicated re-
sult. The semisimplicity of A is necessary for the validity of
the result.
Another easy application of Theorem 6.1.1 yields the following
generalization of Wiener's Theorem (Theorem 4.6.2):
"t
Theorem 6.2.2 (Wiener-Levy Theorem). Let f = (e
l
I -n < t < n).
If h E AC(f) and f is a complex-valued function defined and
analytic on some open set 0 ~ R(h), then there exists some g AC(f)
such that g = f 0 h.
Proof. The result follows immediately from Theorem 6.I.l(i)
on recalling that the Gel'fand transformation on AC(f) is the
identity mapping of AC(f) into C{f) (Theorem 4.6.1).
Wiener's Theorem is, of course, the special case of this result
it
where h(e ) ~ 0, -n ~ t < n. and f(C) = lIe. , E ~ .
6.3. Zeros of Entire Functions. As another application of the
results of Section 6.1 we shall show that, if x is an element of
a commutative Banach algebra A with identity and f is a noncon-
stant entire function such that f{x) = 0, then there exist some
nonnegative integer nand b
k
~ k = 0,l,2, ... ,n such that
6.4. Connected Component of the Identity 155
tk=Obkx
k
= 0, where not all of the b
k
are zero.
Theorem 6.3.1. Let A be a commutative Banach algebra with
identity and let x (A. If there exists a nonconstant entire func-
tion f such that f(x) = 0, then there exists a nonconstant poly-
nomial p such that p(x) = o.
Proof. From Theorem 6.1.1 we know that f(x) E A for any

entire function f and that f(x) (T) = f[x(T)], T E Hence,

since f(x) = 0 by assumption, we see that f[x{T)] = 0, T E A(A);
that is, f vanishes identically on a(x). Since a(x) is compact
and the zeros of a nonconstant entire function are isolated
[A, p. 127], we conclude that a(x) is finite, say
Let m
k
denote the multiplicity of the zero of f at
k = 1,2, . ,r, and define
r
p(C) = n (C - a )mk.
k=l k
Then, as is easily verified, the function g defined by
gCC) = fCC)/pCC), C (C, is an entire function that never vanishes

on o(x); that is, g[X(T)] 0, T 6(A).
However, by Theorem 6.1.1 and Corollary 3.4.2(ii), we see that
g(x) (A is regular and f{x) = p{x)g{x) = O. Therefore p(x) = 0.
0
6.4. The Connected Component of the Identity in A-I. If A
is a commutative Banach algebra with identity e, then we have seen
-1
previously that A , the set of regular elements in A, is an
open subgroup of A. We now wish to show, at least in the case that
A is semisimple, that the connected component of e in A-I is
precisely the set of those elements of A of the form
x E A. We recall that the connected component of point in a topolo-
gical space is the largest connected set that contains the given
point.
156 6. Analytic Functions and Banach Algebras
In what follows, exp will denote the entire function
m n
exp(C) = ~ n = O In!, C (Cj that is, exp is the complex exponential
function. We need two lemmas before the indicated theorem. The
proof of the first lemma is left to the reader.
Lemma 6.4.1. Let A be a semisimple commutative Banach algebra
with identity. If x,y A, then
(i) exp (x) A.
(1
1 ) ( ) - w n
l
I
exp x - Ln=OX n ..
(iii) exp(x + y) = exp(x)exp(y).
Lemma 6.4.2. Let A be a semisimple commutative Banach algebra
with identity e. If x E A and lie - xII < 1, then there exists
some y (A such that exp(y) = x.
Proof. Since lie - xII < 1, we see from Theorem 1.4.1 that

x is regular, whence, by Corollary 3.4.2{ii), R(x) is a compact
set that does not contain zero. Let 0 be an open subset of C

such that 0 ~ R(x) and 0 ~ O. Then fCC) = 10gC, C 0, is
analytic on 0, and so, by Theorem 6.1.1, y = f(x) E A and
.. .
yeT) = log X(T), T (6(A). But exp{y) E A and

exp{y) (T) = exp[Y(T)] = exp[log X(T)] = X(T) (T E 6(A,
from which we conclude that exp(y) = x, as A is semisimple.
o
We can now state and prove the desired theorem.
Theorem 6.4.1. Let A be a semisimple commutative Banach
algebra with identity e and let exp(A) = (exp(x) I x Al. Then
-1
exp(A) is the connected component of e in A .
Proof. From Lemma 6.4.1 we see that exp(x)exp(-x) = exp{O) = e,
~ d ~ -1 E h"
x ~ A, an so e ~ exp(A) cA. Moreover, if x A, t en 1t lS
easily verified that wet) = exp{tx), 0 < t < 1, defines a continuous
6.4. Connected Component of the Identity 157
curve in exp(A) from e to exp(x), from which it follows at
once that exp(A) is connected. Thus exp(A) is contained in the
-1
connected component of e in A . To show that exp(A) is this
connected component it suffices to prove that exp(A) is both open
and closed in A-I.
To this end suppose y = exp(x) and let z (A be such that
IIY - zll < lIlly-III. Then
lie - y-l
z
l1 = lIy-l
y
_ y-l
z
l1
~ lIy-
1
11l1y - zll
< 1,
and so, by Lemma 6.4.2, there exists some w E A such that
-I
exp(w) = y z. Consequently
z = y[exp(w)] = exp(x)exp(w) = exp(x + w),
whence z ~ exp(A), and exp(A) is open.
On the other hand, suppose y E A-I and y E cl[exp(A)]. Then
there exists some z = exp(x) such that lIy - zll < lIlly-III. Arguing
as before, we conclude that y-I Z (exp(A), and so y-l E exp(A).
Thus y E exp(A) and exp(A) is closed.
Therefore exp(A) is the connected component of e in
-1
A 0
Actually the theorem and the lemmas are valid in any commutative
Banach l g e b ~ with identity. We have proved the results only for
semisimple algebras because the arguments in Lemmas 6.4.1(iii) and
6.4.2 are less technical in this case. The lemmas can be established
in the general case via a power series argument. See, for example,
[8, pp. 49 and 50].
We shall not pursue the discussion of A-I and exp(A) any
further, but content ourselves with mentioning one of the most
158 6. Analytic Functions and Banach Algebras
important theorems on the subject. Note that exp(A) is evidently
a subgroup of A-I.
Theorem 6.4.2 (Arens-Royden). Let A be a commutative Banach
algebra with identity. Then the quotient group A-1/exp(A) is
isomorphic to HI[a(A),Z], the first ~ e h cohomology group of a(A)
with integer coefficients.
-1
Using this result one can prove, for instance, that A is
either connected or has an infinite number of connected components.
Discussions of the Arens-Royden Theorem and some of its consequences
can be found in [Ga, pp. 88-91; 5, pp. 98-104; Wm
2
, pp. 88-96].
More generally, other treatments of the role of analytic func-
tions in the study of Banach algebras are available in [B, Ga, Ho
l
,
Hr, Lb, Ri, 5, wm
l
, Wm
2
].
CHAPTER 7
REGULAR COMMUTATIVE BANACH ALGEBRAS
7.0. Introduction. If X is a compact Hausdorff topological
space, then it is well known that X is a completely regular topo-
logical space and a normal tOP91ogicai space. The first assertion
is equivalent to saying that for each closed set E C X and each
point t E X, t t E, there exists some f E C(X) such that
< f(s) < 1, s ( X, fet) = 1, and fes) = 0, sEE. The second
assertion, combined with Urysohn's Lemma, reveals that, if El C X
and E2 C X are disjoint closed sets, then there exists some
f E C(X) such that 0 ~ f(s) ~ 1, s ( X, f(s) = 1, s (E
l
, and
f(s) = 0, s E E
2
. Recalling that the Gel'fand representation of

the commutative Banach algebra A = C(X) is just A = C(X), we see

that these observations say precisely that A has the indicated
properties. In particular, if A = C(X), X being a compact Haus-

dorff topological space, then A separates disjoint closed subsets
of A(A). Our goal in this chapter is to determine sufficient condi-
tions on an arbitrary commutative Banach algebra A such that the

elements of A possess separation properties analogous to those
described here.
In order to accomplish this we shall introduce a new topology
on ~ A ) , the hull-kernel topology, and show that, when this topo-
logy coincides with the Gel'fand topology on A(A) , then, whenever
E C A(A) is closed and T E A(A), T ~ E, there exists some x E A

such that X(T) = 1 and xew) = 0, wEE. Furthermore, in this
case, we shall also see that, if K C A(A) is compact, E C A(A)
is closed, and K n E = ~ , then there exists some x E A such that

X(T) = 1, T E K, and X(T) = 0, T ~ E.
159
160 7. Regular Commutative Banach Algebras
A commutative Banach algebra A that satisfies either of the
foregoing separation properties will be termed regular or normal,
respectively. The hull-kernel topology and regular commutative
Banach algebras play a central role in the study of the ideal struc-
ture of commutative Banach algebras, as we will see in the next
chapter.
7.1. The Hull-Kernel Topology and Regular Commutative Banach
Algebras. The main concerns of this section will be to define and
investigate the elementary properties of the hull-kernel topology
on A (A) and to determine when this topology coincides with the
Gel'fand topology. We begin with a number of preliminary definitions.
Definition 7.1.1. Let A be a commutative Banach algebra. If
E C A(A) , then the kernel of E, denoted by keEl, is defined as
k(l:) = n M =
MEE
if E ~ ~ , while k ~ ) = A. If I C A is an ideal, then the hull
of I, denoted by h(I), is defined as
h(l) = [M I M A(A), M IJ
= (T I T E A(A), T-I(O) ~ IJ.
For each x E A, the ~ set of x, denoted by Z(x), is defined
as
Z(x) = {T
..
T E. A(A), X(T) = 0)
= {M I M E A(A), x EM),
whereas if I C A is an ideal, then the ~ set of I, denoted
by Z(I), is defined as
Z(I) = fl Z(x) = {T
xEI
..
T (A(A), X(T) = 0, x ( I)
= {M M E a(A), I eM).
7.1. Hull-Kernel Topology 161

It is evident from the definitions and the continuity of x
that Z(x) and Z(I) are closed subsets of and Z(I) = h(I).
Moreover, we have the following proposition:
Proposition 7.1.1. Let A be a commutative Banach algebra.
(i) If I C A is an ideal, then h(l) = h[cl{I)].
(ii) heAl = and h(O) =
(iv) If E C then
..
keEl = (x I x A, X(T) = 0, TEE) = {x I x A, Z(x) El.
(v) If E C then keEl is a closed ideal in A.
(vi) = Rad(A).
(vii) If I C A is an ideal, then cl(I) C k[h{I).
(viii) If I C A is an ideal, then h{I) = h(k[h(I)]); that is,
h = hkh.
(ix) If E C then keEl = k(h[k(E)]); that is, k = khk.
(x) If E C then h[k(E)] E.
(xi) If El C E2 C and El C E
2
, then k(E
l
)::> k (E
2
)
and h[k(E
l
)] C h[k(E
2
)].
(xii) If El C and E2 c then h[k(E
l
) n k(E
2
)] =
h [k (E
l
U E
2
)].
Proof. All of the proofs are rather elementary. We shall give
only one proof and leave the rest to the reader.
Suppose
h (k [h (I) ] )
I C is an ideal. We shall show that h(l) =
If M E h(k[h(I)), then M k[h(I)] I by part (vii),
162 7. Regular Commutative Banach Algebras
whence M E h(I). Conversely, suppose T h(k[h(I)) = Z(k[h(I)).
"
Then there exists some x E k[h(I)] such that X(T) 0, while
A
from part (iv) we see that x(w) = 0 W E hel). Hence
T h(I)"
and so h(l) = h(k[h(I)]), which proves part (viii).
o
As the reader should guess from the preceding proof, in discus-
sing hulls and kernels it is often of considerable advantage to keep
in mind the descriptions of these objects both in terms of maximal
regular ideals and in terms of complex homomorphisms.
The comment preceding Proposition 7.1.1 shows that h(l) is a
closed subset of 6(A) for each ideal leA. The idea behind the
hull-kernel topology is to use such closed sets, that is, hulls of
ideals, as the closed sets in a topology. With this in mind we make
the following definition:
Definition 7.1.2. Let A be a commutative Banach algebra. If
E c 6(A) , then the hull-kernel closure of E, denoted by E, is
defined to be E = h[k(E)].
The reader should recall that the closure of E in the Gel'fand
topology is denoted, as usual, by cl(E).
In order to show that the hull-kernel closures of sets in 6(A)
are actually the closed sets for some topology on 6(A) we must
show that the correspondence E - E, E C 6(A) , is a closure operation
[DS
l
, pp. 10 and 11]; that is, we must show that the operation of
forming the hull-kernel closure satisfies the following conditions:
(a) 6(A) = 6(A).
(b) E C E.
(c)
E = E.
Cd) (E
l
U E
2
)- = El U E
2

7.1. Hull-Kernel Topology
163
Evidently conditions (a) and (b) hold by Proposition 7.1.I(vi)
and (x), respectively, and
E = h[k(h[k(E)])] = h[k(E)] = E,
by Proposition 7.1.1(ix).
So suppose EI C A(A) and E2 C A(A). From Proposition 7.1.I(xi)
we see that
(k = 1,2),
whence El U E2 C (E
I
U E
2
)-. Conversely, suppose M belongs to
~ A ) - (E
I
U E
2
) and let u be an identity modulo M. Clearly
k(E
k
) ~ M, k = 1,2, and u ~ M, 9y Proposition l.l.l(i). Thus,
since M is the null space of a nonzero continuous linear functional
and hence is of codimension one [L, p. 68], we see that there exist
v,w ~ M, x e keel)' and y e k(E
2
) such that u = x + v = y + w.
2 2
However, u - xy = u - (u - v,(u - w) = (v + w)u - VW, which
2
shows that u - xy M, as M is an ideal, while xy E keEl) nk(E
2
).
Now, if M E (E
I
U E
2
) = h[k(E
l
U E
2
)], then, from Proposi-
tion 7.1.1(xii) we see that M h[k(E
1
) n k(E
2
)]. Thus, in this
2 2
case, we conclude that xy M, whence u E M, as u - xy M.
However, since u is an identity modulo M, we have u
2
- u M,
and so u E M, which is a contradiction. Therefore M (E
I
U E
2
)
and (E
I
U E
2
)- = El U E
2
.
Thus we see that the correspondence E - E, E C ~ A ) , is a
closure operation and so it can be used to define a topology on ~ A ) .
Definition 7.1.3. Let
Then the topology on ~ A )
E - E = h[k(E)], E C ~ A ) ,
A be a commutative Banach algebra.
determined by the closure operation
is called the hull-kernel topology.
The relation of the hull-kernel topology to the Gel'fand topo-
109)' on ~ A ) is contained in the next theorem. We recall that a
topology is said to be Tl if every singleton set is closed.
164
7. Regular Commutative Banach Algebras
Theorem 7.1.1. Let A be a commutative Banach algebra. Then
the hull-kernel topology on A(A) is a Tl topology which is weaker
than the Gel'fand topology. Moreover, if A has an identity, then
A(A) in the hull-kernel topology is compact.
Proof. Since h[k([M))] = (M), M (A(A), and h[k(E)], E C A(A),
is always closed in the Gel'fand topology, we see at once that the
hull-kernel topology is Tl and weaker than the Gel'fand topology.
A bit more work is required to show that 6(A) is compact in the
hull-kernel topology when A has an identity.
If A has an identity e, then in order to show that deAl
is compact in the hull-kernel topology it suffices to prove that,
if (E) C A(A) is a family of hull-kernel closed sets with the
a
finite intersection property, then n E Evidently, if (E)
aa a
is such a family of sets, then E # and k(E) is a proper
a a
closed ideal in A. Let I denote the closed ideal in A generated
by U k(E); that is, I is the norm closure of the linear subspace
a a
in A generated by Uak(Ea'. It is easily seen that I is a closed
ideal, and, moreover, I is proper. Indeed, if I = A, then there
would exist a
l
,a
2
, . ,a and x
k
(k(E ), k = 1,2, ... ,n, such
n
that lie -1:k=lxkl: < 1. Hence, by Theorem 1.4.1, is regular.
Let y be the inverse of Ik=IX
k
. Clearly e = Ik=IXkY and
xky E k(E
Ok
), k = 1,2, ... ,n. Thus we see at once that the norm
closed ideal J generated by is all of A, as e J.
We claim that If this were not so, there would
exist some M E A(A) such that M k(E ), k = 1,2, ... ,n, and

hence M J = A. This, contradicts the properness of M.
Thus = D, whence = D since EG\ = EOk
= h[k(E )], k = 1,2, ... ,n. The latter assertion is contrary to the

assumption that (E)
a
has the finite intersection property, and so
we conclude that I is proper.
Hence, by Theorem 1.1.3, there exists some M E A(A) such that
M I k(Ea' for each a. Consequently from Proposition 7.1.1(iii)
7.1. Hull-Kernel Topology 165
we see that M = heM) c n h[k(E )] = n E, whence n E ~ ~ .
a a aa aa
Therefore a (A) is compact in the hull-kernel topology.
o
The next most obvious question to raise is: ~ ~ e n do the hull-
kernel and Gel'fand topologies on ~ C A coincide? From Theorem
7.1.1 and an argument used several times previously we see that, if
A has an identity, then the topologies will coincide precisely
when the hull-kernel topology is Hausdorff. This observation is
actually true in general, as is shown by the next result.
Theorem 7.1.2. Let A be a commutative Banach algebra. Then
the following are equivalent:
(i) The hull-kernel topology and the Gel'fand topology on a(A)
coincide.
(ii) The hull-kernel topology on a(A) is Hausdorff.
(iii) If E c a(A) is closed in the Gel'faBd topology and
A
T a(A), T ~ E, then there exists some x E A such that X(T) = I
A
and x(w) = 0, wEE.
Proof. Obviously part (i) implies part (ii), and from the
remark preceding the theorem we see that part (ii) implies part (i)
if A has an identity. If A is without identity, then consider
the algebra A[e]. By the comment following Definition 3.2.2 and
Theorem 3.2.2(ii) we have a(A[e]) = a(A) U (T), where T E a(A[e])
co CD
is the unique complex homomorphism of A[e] such that T (x) = 0,
co
X (A. Denoting hulls and kernels computed with respect to aCA[e])
by hand k, respectively, it is easily verified that
e e
h [k (E)] = h(k[E n 6(A)]) U (T )
e e co
(E c: a (A [ e ]) ) .
From this it is apparent that the hull-kernel topology on a(A)
coincides with the relative hull-kernel topology inherited from
A(A[e]), and the same is true of the Gel'fand topology by Theorem
166
7. Regular Commutative Banach Algebras
3.2.2(ii). Since the hull-kernel topology on A(A) is assumed to
be Hausdorff, it follows that the hull-kernel topology on A(A[e])
is Hausdorff, whence, as before, the hull-kernel and Gel'fand topo-
logies on A(A[e)) coincide. Hence these topologies on A(A) also
coincide, and part (ii) implies part (i) in the case that A is
without identity.
Now suppose the hull-kernel and Gel'fand topologies on A(A)
coincide and let E CA(A) be closed and T E A(A) - E. Since
E = E = h[k(E)], we see that T l h[k(E)], and so there exists

some x E keEl such that X(T) O. However, since x keEl,

x(w) = 0, wEE, and part (i) implies part (iii).
Conversely, suppose part (iii) holds. Thus, if E C A(A) is
closed in the Gel'fand topology and T E A(A) - E, then there exists

some x E A such that X(T) = 1 and x(w) = 0, wEE. Hence
x E keEl and T h[k(E)] = E, from which we deduce that, if T E,
then T E; that is, if E is closed in the Gel'fand topology,
then E is closed in the hull-kernel topology. Consequently the
Gel'fand topology is weaker than the hull-kernel topology, which,
combined with Theorem 7.1.1, shows that the two topologies coincide.
Therefore part (iii) implies part (i), and the proof is complete.
O
Recalling the discussion in the introduction to this chapter

we see that the Gel'fand representation A of a commutative Banach
algebra A separates points and closed sets precisely when the
hull-kernel topology is Hausdorff. We wish to single out such alge-
bras with a special name.
Definition 7.1.4. Let A be a commutative Banach algebra.
Then A is said to be regular if the hull-kernel topology on A(A)
is Hausdorff.
In view of the topological remarks in the introduction the
reader may wonder why we do not call such Banach algebras completely
7.2. Some Examples 167
regular instead of regular. Actually this is done by some authors
(see, for instance, [Ri, pp. 83-96]), but the term "regular" seems
to be more widely used, and for this reason we prefer it.
Before we show that the Gel'fand representation of a regular
commutative Banach algebra also separates compact sets from closed
sets, we shall consider the question of regularity for several
specific algebras.
7.2. Some Examples. As mentioned in the introduction, every
compact Hausdorff topological space X is normal, whence we see
that the commutative Banach algebra C(X) is regular as the Gel'fand
transformation on C(X) is just the identity mapping (Theorem 4.1.1
and Corollary 4.1.1). Similarly, if X a locally compact Haus-
dorff topological space, then C (X) is a regular algebra. To see
o
this, in view of Theorem 7.1.2, we need only show that, if E C X
= is closed, then it is closed in the hull-kernel topology;
that is, E = h[k(E]). Denoting the compactification of
X by X+, we know that C (X)[e] = C(X+), and so, by the regularity
+ 0
of C(X), we deduce that
E = cl(E) = cl (E) n X
e
= he[ke(E)] n X
= (h[k(E)] U {T )) n X
m
= h [k (E)] ,
where cle(E) denotes the closure of E in X+. Thus Co (X) is
regular. Using Theorem 7.1.2(iii), it is also easily seen that,
if a,b a < b, then Cn([a,b]), n = 1,2,3, . , is regular.
On the other hand, the commutative Banach algebra A(D) is
not regular. Indeed, consider the set E = {lIn I n = 2,3, .. ).
Evidently E C D and cl(E) = E U (oj. But recalling that the
Gel'fand transformation on A(D) is again the identity mapping,
168
7. Regular Commutative Banach Algebras
we see that
keEl = (f I f E A(D), f(l/n) = 0, n = 2,3, . }
= (0),
since the zeros in an open set of a nonconstant analytic function
are isolated [A, p. 127]. Hence E = h[k(E)] = herO)) = A(A(D)) =
D cl(E). Thus the Gel'fand and hull-kernel topologies do not coin-
cide for A(D), and so A(D) is not regular.
As our final example we wish to show that LI(G), G being a
locally compact Abelian topological group, is regular. However,
this result is not as simply proved as the previous assertions. In
order to do it we shall have to appeal to another fundamental theorem
of harmonic analysis, Plancherel's Theorem, whose proof will not be
given until Section 10.4.
Before stating this result we wish to remind the reader of some
of the discussion in Section 4.7. We saw there that the maximal
..
ideal space A(L
I
(G)) is identifiable with the dual group G of G,
that is, with the group of continuous homomorphisms of G into
r = {, I , (Q;, 1,1 = I). Moreover, with the Gel'fand topology,
..
G is even a locally compact Abelian topological group, so, in parti-
..
cular, we can speak of Haar measure on G. Furthermore, the
Gel'fand transform, which in this instance we refer to as the Fourier
transform, is defined by
..
fey) = SG f(t) (t,y) dArt)
..
(y E G; f E L
1
(G)).
Plancherel's TheoTem assures us of the existence of a dense linear
.. ..
subspace Vo of L
2
(G) such that Vo C LI(G) n L
2
(G) , Vo C L
2
(G),
and on which the Fourier transform is an L2 isometry. More pre-
cisely we have the following theorem:
7.2. Some Examples 169
Theorem 7.2.1 (Plancherel's Theorem). Let G be a locally
compact Abelian topological group and let
on G. Then there exists a Haar measure
A be a given Haar measure
..
~ on G and a linear
subspace V of
0
L
2
(G) such that
(i) Vo C LI(G) n L
2
(G).
(ii) V is norm dense in
L2 (G) .
0
,.

(iii) V is norm dense in L
2
(G).
0
,.
(iv)
IIfll2 = II
f
1i
2
,
f E V .
0
.. ..
(v) The mapping
f - f, f E Yo'
from V to V can be
0 0
..
uniquely extended to a linear isometry of L
2
(G) onto
L2 (G) .
The extensiop of the Fourier transformation on V to all of
o
L
2
(G) will be called the Plancherel transformation, and we shall
,.
denote the Plancherel transform of f once again by f. Due to
the way the Plancherel transform is defined, it shares a number of
the formal properties enjoyed by the Fourier transform.
lar, it is not difficult to show that, if f E L
2
(G),
&,. ,.,. ,.
In particu-
then
fey) = f(-y) and [(.,w)f] (y) = fey - w), y,w E G, where, of
course, the identities are interpreted to hold only almost every-
where with respect to 11. The details are left to the reader.
As we shall subsequently see, the proof of Plancherel's Theorem
is a nontrivial matter. However, in the case of compact Abelian
topological groups the result is an easy consequence of our discus-
sion in Section 4.7 and standard results of Hilbert space theory.
Indeed, if G is compact Abelian and Haar measure A on G is
normalized so that A(G) = 1, then from Corollary 4.7.4 we see that
,.
G is a complete orthonormal set in the Hilbert space L
2
(G). By
,.
Corollary 4.7.5 the dual group G is discrete, and we may assume
..
that Haar measure ~ on G is normalized so that ~ y } ) = 1,
..
Y (G. Then, since L
2
(G) C LICG), we see that the Fourier trans-
formation is defined on all of L
2
(G), and, by a standard theorem
170 7. Regular Commutative Banach Algebras
of Hilbert space theory [L, p. 405], we have
IIfll2 = [ t.lf(Y)12]1/2
yEG
= [Sa If(y)1
2
dTJ(y)]1/2

= IIfll2
Moreover, the Riesz-Fischer Theorem [L, p. 404] shows that the map-

ping f - f, f E L
2
(G), is surjective, and so Plancherel's Theorem
is proved in the case of compact Abelian topological groups. Here,
of course, Vo = L
2
(G).
OUr concern now, however, is not to prove Plancherel's Theorem,
but to use it. The first consequence of the theorem we wish to men-
tion is the next corollary.
Corollary 7.2.1. (Parseval's Formula). Let G be a locally
compact Abelian topological group and let A be a given Haar measure

on G. If Haar measure ~ on G is so chosen that Plancherel's
Theorem is valid, then

SG f(t)g(t) dA(t) = Sa f(y)g(y) dTj(y)
Proof. Apply Plancherel's Theorem to the identity
o
The next consequence we need -- one that is of considerable
interest in its own right -- says that the Gel'fand representation
_. ..
of Ll(G) is precisely L
2
(G) * L ~ C G ; that is,. f E Ll(G) if
and only if there exist g,h E L
2
(G) such that f = g * h. Note

that the equality here holds pointwise since both f and g * h
-
belong to Co(G).
Theorem 7.2.2. Let G be a locally compact Abelian topological
group and let A be a given Haar measure on G. If Haar measure ~

on G is so chosen that Plancherel's Theorem is valid, then

LICG) = L
2
(6) * L
2
(6).
7.2. Some Examples
171
Proof. If f E Ll(G), then we can write f = glh
l
, where
gl,h
l
E L
2
(G). For instance, define hI and gl almost everywhere
as hl(t) = If(t)1
1
/
2
, gl(t) = e-iargf(t)lf(t)ll/2. Let g and h
A
in L
2
(G) denote the Plancherel transforms of gl and hI' respec-
tively. Then, by Parseval's Formula and the observations following
A
Plancherel's Theorem, we have, for each w E G,
A
few) = IG f(t) (t,w) 4A(t)
= IG gl{t)h1(t) (t,w) dArt)
= Ia g(y)h(w - y) d ~ y )
= g * hew).
A
Conversely, suppose g,h L
2
(G)
such that the Plancherel transform of
is h. Evidently f = g
l
h
1
E Ll(G),
and let gl,h
1
(L
2
(G) be
gl is g and that of hI
and a repetition of the pre-
A
ceding argument reveals that f = g * h.
.. A A
Therefore LI(G) = L
2
(G) * L
2
(G).
The regularity of L
1
(G) will be a corollary of the next
theorem.
o
Theorem 7.2.3. Let G be a locally compact Abelian topologi-
cal group, let A be a given Haar measure on G, and suppose Haar
A
measure ~ on G is so chosen that Plancherel's Theorem is valid.
A A
If KeG is compact and U eGis measurable and such that
o < ~ U ) < GO, then there exists some f L
1
(G) such that
A
(i) fey)
=
1, Y E K.
A
(iiJ fey) = 0, ylK+U - u.
A A
(iii) o fey) ~ 1, Y E G.
172 7. Regular Commutative Banach Algebras
..
Proof. Clearly Xu/ij(U) and X
K
_
U
belong to L
2
(G) , where, as
usual, X
E
denotes the characteristic function of E. Hence, by
Theorem 7.2.2, there exists some f (LI(G) such that
..
f = [Xu/Tj(U)] * X
K
-
U
. Thus we see that
1
fey) = U ) faxK-u(y - w)Xu(w) d ~ w )
= ~ t U ) J
u
xK-U(y - w) dl1(w)
..
(y ~ G).
Hence, if y E K, then y - w ~ K - U, w E U, and so
.. I J"
fey) = Tl{U) U dTj(w) = I,
whereas, if Y ~ K + U - U, then y - w ~ K - U, w E U, whence
.. .. ..
fey) = O. Finally, it is obvious that 0 ~ fey) ~ 1, y E G.
o
Corollary 7.2.2. Let G be a locally compact Abelian topolo-
gical group.
that W:J K,
..
.. .
If KeG is compact and W eGis open and such
then there exists some f E LlCG) such that
(i) fey) = 1, Y E K
..
(ii) fey) = 0, y ~ w.
.. ..
(iii) 0 ~ f (y) 5 1, Y G.
Proof. In view of Theorem 7.2.3 and the fact that the Haar
measure of a nonempty open set is positive (Proposition 4.7.3), it
..
suffices to find an open neighborhood U of y in G such that
o ..
K + U - U C W. As usual, y denotes the identity of G. To do
o
this we obviously may assume, without loss of generality, that Yo
is in K. Moreover, we claim that it suffices to show only that
..
there exists some open neighborhood V of y in G for which
o
K + V C W.
Indeed, suppose such a neighborhood V exists. Then, by the
continuity of the group operations, for each y E V there exists

an open neighborhood U of y in G such that U C V and
y 0 y
7.2. Some Examples 173
Y + U c V. Set U = U (VU. Clearly U is an open neighborhood
Y. Y Y
of y in G and U C V. Furthermore,
o
U C V = U (y + U ) c V + U U = V + U,
yE.V y y(V y
whence we conclude that U - U C V. Thus K + U - U C K + V C W,
as we wished to show.
Hence it remains only to prove the existence of an appropriate
neighborhood V. Again from the continuity of the group operations
one deduces for each y E. K the existence of an open neighborhood

V of y in G such that y + V + V C W. Evidently the family
y 0 y y
{y + Vy lyE K} forms an open covering of K, and so, since K
is compact, there exist v
1
.v
2
... yn in K such that
K c ur
k
l(Y
k
+ V ). Let V = nP IVy' Then V is an open neighbor-
= Yk. K= k
hood of Yo in G, and if y E K, there exists some k, 1 ~ k < n,
such that Y (Y
k
+ V
Yk
' Thus
Y + V C Yk + V + V C Yk + V + V C W,
Yk Yk Y
k
from which we conclude that K + V C w.
o
Corollary 7.2.3. Let G be a locally compact Abelian topolo-
gical group. Then Ll(G) is a regular commutative Banach algebra.

Proof. Suppose E eGis closed and w E. G - E. Then K = {w}
is compact, and there exists an open neighborhood W of w such
that W n E = " as E is closed. By Corollary 7.2.2 there exists

some f E LI(G) such that few) = 1 and fey) = 0, Y E E, whence,
by Theorem 7.1.2, we conclude that L1(G) is regular. 0
For particular groups G it is possible to prove the regularity
of LI(G) by direct arguments, thereby avoiding the use of Plancher-
el's Theorem. For example, suppose G = If E ~ is closed
and t l E. then let 6 > 0 be such that (t - 6,t + 6) n E = ,.
000
Define h to be the continuous tent function
174 7. Regular Commutative Banach Algebras
1
h{s) = 2n (o +
s) for -6 < s < 0,
1
s) for o < s < 6, h{s) = - (o -
2n
h{s) = 0 for
lsi
> 0,
and set
it u ius
feu) = e 0 h{s)e ds
(u E lR).
..
Then f E LlOR), and direct computation reveals that f{t) = h{t - to)'
A
t E IR. Hence f{t) = 6 ; 0 and f{t) = 0, tEE, which shows
o
that Ll (IR) is regular.
7.3. Normal Commutative Banach Algebras. In this section we
shall show that the Gel'fand representation of a regular commutative
Banach algebra A separates not only points and closed sets in
but also compact sets from closed sets. Before we can prove this
result, however, we need to establish some facts about the Gel'fand
representation theory of ideals and quotient algebras. Once again
these results are of considerable interest by themselves.
Theorem 7.3.1. Let A be a commutative Banach algebra and
let I C A be a closed ideal. Then
(i) a(l) is homeomorphic to a(A) - h{I).
(ii) If x E I
o
and x denotes the Gel'fand transform of x
o
as an element of the commutative Banach algebra I, then x is the
0
restriction of x to 6(A) - h(I); that is, x = xI
6
{A)-h{I).
o
(iii) If.1 denotes the Gel'fand representation of I, then
o
I =
(iv) 6(A/I) is homeomorphic to h(I).
(v) If
o
(x + I) denotes the Gel'fand transform of x + I in
o -
A/I, then (x + I) is the restriction of x to hel); that is,
o
ex + I) = xih(I)'
7.3. Normal Algebras 175
o
(vi) If (A/I) denotes the Gel'fand representation of A/I,
then (A/I)
o

= Alh(I)'
(vii) A/I is semisimple if and only if I = k[h(I)].
Proof. Suppose T E Then it is evident that T restrict-
ed to I is either a complex homomorphism on I or is identically

zero. But T(X) = 0, x E I, if and only if X(T) = 0, x E I, that
is, T E h(I). Thus every T E A(A) - h(l) defines a complex homo-
morphism on I via restriction to I. On the other hand, suppose
T (A(I). Then there exists y E I such that T(Y) = 1. De-
fine w: A - C by w(x) = T(xy), x E A. Clearly w is linear,
and if x,z E A, then
w(xz) = T(XZY) = T(XZY)T(Y) = T(XYZY) = T(xy)TCZY) = w(x)w(z),
whence w E Obviously w(x) = T(X), x E I, and w h(I).
Moreover, it can be shown that w is unique: Suppose w' E A(A) is
such that w'(x) = T(X), x E I. Then for any x E A we have
w'(X) = W'(X)T(y) = w'(x)w'(y) = w'(xy) = T(xy) = w(x),
and so w' = w. Thus we see that the restriction of T E A(A) - h(I)
to I defines a bijective mapping between A(I) and A(A) - h(I).
Furthermore, from the definition of this mapping it is apparent that
o 0
x = xIA(A) -h(I)' x E I, where x denotes the Gel'fand transform
of x E I. Finally, it is now clear that the weakest topology on
o
ACI) that makes all of the functions x continuous is just the
relative Gel'fand topology on A(A) - h(I), from which we conclude
that A(I) and A(A) - h(l) are homeomorphic. This proves parts
(i) through (iii).
Next let A - A/I denote the canonical homomorphism; that
is, = x + I, x E A. If T E then it is evident that
W = T 0 is a complex homomorphism on A, where 0 denotes the
usual composition of mappings. Moreover, if x E I, then
176 7. Regular Commutative Banach Algebras
w(x) = T 0 q>(X) = T(X + I) = T(I) = 0,
whence we see that w E h(I). Conversely, if w E h(I), then we
define T on A/I by setting T(X + I) = w(X) , x E A. This is
well defined since, if x,y E A and x - y E I, then
T[(X + I) - (y + I)] = T(X - Y + I) = w(x - y) = 0,
as w E h(I). Thus T E 6(A/I), and w = T Q q>. In this way we
obtain a bijective mapping from 6(A/I) to h(I).
o
Furthermore, it is easily verified that, if (x + I) denotes
o ..
the Gel'fand transform of x + I E A/I, then (x + I) = xlh(I)
and that 6(A/I) and h(I) are homeomorphic. This proves parts
(iv) through (vi).
o
Finally, A/I is semisimple if and only if (x + I) (T) = 0,
Q
T (6(A/I), implies x E I. But it is readily seen that (x + I) =
if and only if x E k[h(I)]. Since k[h(I)] ~ I, by Proposition
7.1.1(vii), we conclude that A/I
I=k[h(I)].
is semisimple if and only if
o
Before we establish the theorem indicated at the beginning of
this section we shall make a definition and prove some preliminary
results.
Definition 7.3.1. Let A be a commutative Banach algebra.
We say that A is normal if, whenever K C 6(A) is compact,
E C 6(A) is closed and K n E = ~ there exists some x E A such
that
..
(i) X(T) = 1, T E K .

(ii) X(T) = 0, TEE.
Corollary 7.2.2 says precisely that I.I(G), G being a locally
compact Abelian topological group, is normal. We shall see that the
7.3. Normal Algebras
177
same is true of every regular commutative Banach algebra.
Lemma 7.3.1. Let A be a regular commutative Banach algebra.
If T E ~ A ) , then there exist some x E A and some open set

U C b(A) containing T such that x(w) = 1, w E U.

Proof. Clearly there exists some yEA such that yeT) ~ o.
Let U be an open neighborhood of T with compact closure such that

yew) ~ 0, w cl(U). Such a neighborhood exists, as ~ A ) is locally
compact. Since A is regular, we have cl(U) = h(k[cl(U))). Set
I = k[cl(U)). Then I is a closed ideal, and from Theorem 7.3.1
o
we see that A(A/I) = h(l) = cl(U) and (y + I) = yICl(U). In
particular, (y + I) never vanishes on A(A/I), and ~ A / I ) is
compact. Moreover, A/I is semisimple, as k[h(I}) = k[cl(U)) = I.
Consequently we may apply Corollary 6.2.2 to deduce that A/I has
an identity.
o
Thus !here exists some x E A such that (x + I) = xlh(I) = I,
that is, x(w) = 1, w E cl(U), which completes the proof.
o
Corollary 7.3.1. Let A be a regular commutative Banach alge-
bra. If K C ~ A ) is compact, then there exists some x E A such

that X(T) = 1, T E K.
Proof. From Lemma 7.3.1 we see that for each T E K there
exist some x
T
E A and some open neighborhood U
T
of T such that

xT(w) = I, w E UTe Obviously the family {U
T
I T E K) is an open
covering of K, and so there exist T
I
,T
2
, .. ,T
n
such that
A straightforward calculation then reveals that
n
K C Uk=IU
Tk
.

x = x 0 x 0 0 x
Tl T2 Tn
in A is such that X(T) = 1, T E K,
where x 0 y = x + y - xy.
o
It is perhaps worthwhile pointing out that this corollary, com-
bined with Corollary 6.2.2, gives us another sufficient condition
for a Banach algebra to have an identity.
178 7. Regular Commutative Banach Algebras
...-
Corollary 7.3.2. Let A be a semisimple regular commutative
Banach algebra. If A (A) is then A has an identity.
This corollary, combined with Theorem 4.7.1, yields a partial
converse to Corollary 4.7.5.
Corollary 7.3.3. Let G be a locally compact Abelian topolo-
"
gical group. If G is compact, then G is discrete.
Proof. Since LI(G) is a semisimple regular commutative
"
Banach algebra and A(Ll(G)) = G, we conclude from Corollary 7.3.2
that LI(G) has an identity. But, by Theorem 4.7.1, this occurs
if and only if G is discrete.
o
The normality of regular commutative Banach algebras is an imme-
diate consequence of the next theorem.
Theorem 7.3.2. Let A be a regular commutative Banach algebra.
If K C A(A) is compact, E C A(A) is closed, K n E and I
is any closed ideal in A for which h(l) = E, then there exists
some x E I such that
"
(i) X(T) = 1, T E K

(ii) X(T) = 0, TEE.
Proof. Let I be a closed ideal such that E = h(I); for
example, I could be keEl. By Theorem 7.3.1, A(I) is homeomor-
phic to A(A) - h(l) = A(A) - E. Since this latter set is open in
A(A), it follows at once from the regularity of A that the hull-
kernel topology on A(I) is Hausdorff, and so, by Theorem 7.1.2, 1
is a regular commutative Banach algebra.
Clearly K C A(A) - E = A(I) is compact, and therefore from
Corollary 7.3.1 we deduce the existence of some x I whose Gel'fand
o
transform x is identically one on K.
But, by Theorem 7.3.1,
o "
x = xlA(A) -h(l)' from which we see that
"
X(T) = 1, T K. However,
7.3. Normal Algebras 179

since x E I, we have X(T) = 0, T E h(l) = E, and the proof is
complete.
o
Corollary 7.3.4. If A is a regular commutative Banach algebra,
then A is normal.
In the case that 6 (A) is compact it is immediate that the
-
Gel'fand representation A of a regular commutative Banach algebra
A even separates disjoint closed sets. This observation should
help explain the origin of the term "normal" as applied to Banach
algebras.
CHAPTER 8
IDEAL THEORY
8.0. Introduction. One of the more interesting portions of the
theory of commutative Banach algebras is the study of the structure
of closed ideals. As we shall see, any attempt to completely describe
the closed ideals in an arbitrary commutative Banach algebra is hope-
less, although the task may be relatively easy for certain specific
algebras. Nevertheless a number of questions concerning closed ideals
can be investigated fruitfully.
The first such question we shall take up is to determine suffi-
cient conditions under which a proper closed ideal I in a commuta-
tive Banach algebra A is contained in some maximal regular ideal.
That is, when is it the case that h(l); for each proper closed
ideal I in A? Of course, as we saw previously, if A has an
identity, then this is certainly the case, and similarly every proper
regular closed ideal is always contained in a maximal regular ideal.
However, in algebras without identity it is not entirely clear when
the desired phenomenon occurs. We shall prove in the first section
that, if A is a semisimple regular commutative Banach algebra such
that the elements of A whose Gel'fand transforms have compact sup-
port are norm dense in A, then h(l) for each proper closed
ideal I in A. We shall see that one important example of such a
Banach algebra is LI(G), where G is a locally compact Abelian
topological group.
This result has a number of interesting consequences. For exam-
ple, we shall show that the linear subspace of LI(G) spanned by
the translates of some f E LI(G) is dense in LI(G) if and only if
180
8.1. Tauberian Commutative Banach Algebras 181
A
f never vanishes. Moreover, the indicated result will be used to
establish some Tauberian theorems. We shall prove two such theorems,
one due to Wiener and one to Pitt.
The second question about ideals that we shall consider at some
length is the determination of conditions under which the hull of a
proper closed ideal I uniquely determines I. Since we know that
in a regular commutative Banach algebra h(I) is closed and
h(k[h(I)]) = h(l), we see that I and k[h(I)] are two closed
ideals with the same hull, and we shall see that the question of when
h(l) determines I is equivalent to determining when I = k[h(I)].
In other words, the problem will be to determine when a proper closed
ideal I is the intersection of the maximal regular ideals contain-
ing I. We shall refer to this as the problem of spectral synthesis.
Although it is, in general, a very difficult problem, some general
results are available. For instance we shall see that, if A is a
semisimple regular commutative Banach algebra that satisfies certain
additional requirements known as Ditkin's condition, then I = k[h(I)]
for each proper closed ideal I in A such that the topological
boundary of h(l) contains no nonempty perfect set. The proof of
this result, which appears in Section 8.5, is rather intricate and
involves a number of results of independent interest.
Although the problem of spectral synthesis is, in general, diffi-
cult, its solution for certain algebras is quite elementary. We
shall, for instance, solve the problem completely for C(X), X being
a compact Hausdorff topological space, and for Ll(G), G being a
compact Abelian topological group. The problem for Ll(G) when G
is noncompact is, however, exceedingly complex.
In the final section of this chapter we shall briefly mention
some other questions and results concerning closed ideals.
8.1. Tauberian Commutative Banach Algebras. We have already
observed that, if A is a commutative Banach algebra, then every
proper regular ideal in A is contained is some maximal regular
182 8. Ideal Theory
ideal. However, it need not generally be the case, when A is with-
out identity, that every proper closed ideal is contained in a r-axi-
mal regular ideal. The main result of this section will be to show
that this does occur whenever A is a semisimple regular commutative

Banach algebra such that {x I x E A, x E C (A(A))) is norm dense
c
in A. In particular, we shall see that, if G is a locally compact
Abelian topological group, then every proper closed ideal in Ll(G)
is contained in some maximal regular ideal. Some further applica-
tions of these results will be discussed in the next section.
To begin we make some definitions and some elementary observa-
tions.
Definition 8.1.1. Let A be a commutative Banach algebra and
suppose E C A(A) is closed. Then I (E) will denote the set of
0
all x E A for which x vanishes identically on some open set
Ox C A(A) such that Ox E; and Jo(E) will denote the set of

x E IoCE) such that x Cc(A(A)).
Given a closed set
that Jo(E)
cl[Io(E)).
and I (E)
o
Moreover, it
E C A(A), it is evident that J (E) C I (E),
o 0
are ideals in A, and that cl[Jo(E)) =
follows from the definition of k[h(I))
that, if I is any closed ideal such that h(l) = E, then
k[h(I)) Cl[Io(E)). If A is semisimple and regular, then cl[Io(E)]
is actually the smallest closed ideal in A whose hull is E. This
is the content of the second part of the next theorem.
Theorem 8.1.1. Let A be a regular commutative Banach algebra
and let E C 6(A) be closed. Then
(i) h(cl[I
o
(E)]) = h(cl[Jo(E)]) = E.
(ii) If A is semisimple, then cl[Io(E)] C I whenever I C A
is a closed ideal such that h(l) = E.
Proof. It is apparent that E C h(cl[Jo(E)]). On the other
hand, if T E, then let U be an open neighborhood of T with
8.1. Tauberian Commutative Banach Algebras
compact closure such that cl(U) n E = ,. Since A is regular,

there exists some x E A such that X(T) 0 and x(w) = 0,
w E - U.

Thus the support of x lies in cl(U)

compact, and x vanishes identically on the open set
o = - cl(U)
x
and so is

which contains E. Hence x E J eE) C cl[1 (E)], and X(T) o.
o 0
Consequently T h(cl[Jo(E)]), and so E = h(cl[Io(E)]).
183
To prove part (ii) of the theorem, suppose I C A is any closed

ideal such that hel) = E. If x E Jo(E), then X E C and
c
there exists an open set 0 E on which x vanishes identically.
x
If K denotes the compact support of x, we see at once that
K n E = ,. Thus, by Theorem 7.3.2, there exists some y E I such

that yeT) = I, T E K, and yeT) = 0, TEE. However, it is easily

verified that X(T) = X(T)y(T). T E whence, since A is semi-
simple, we conclude that x = xy E I.
Therefore Jo(E) C I, and so cl[Io(E)] = cl[Jo(E)] C I, as
I is closed.
o
The result indicated in the introduction is a simple corollary
of this theorem.
Definition 8.1.2. Let A be a commutative Banach algebra.
Then A is said to be Tauberian if (x l x E A, E C is
c
norm dense in A.
Corollary 8.1.1. Let A be a Tauberian semisimple regular
commutative Banach algebra. If I C A is a proper closed ideal,
then there exists some maximal regular ideal MeA such that
M I.
Proof. Suppose I is a closed ideal that is contained in no
maximal regular ideal; that is, E = h(I) = ,. Then clearly

Jo(E) = (x I x E A. x E and so, by Theorem 8.1.1, we have
A = cl[Jo{E)] C I, as A is Tauberian. Hence I is not proper. 0
184 8. Ideal Theory
Since all ideals in an algebra with identity are regular, Corol-
lary 8.1.1 has nontrivial content only in the case of algebras without
identity. An important collection of such algebras that are Tauber-
ian are the algebras Ll(G).
Theorem 8.1.2. Let G be a locally compact Abelian topological
group. Then Ll(G) is a Tauberian semisimple regular commutative
Banach algebra.
Proof. We need only prove that Ll(G) is Tauberian, as the
other assertions were established in Theorem 4.7.4 and Corollary
7.2.3. Appealing to Plancherel's Theorem (Theorem 7.2.1), we see

at once that the set of f E L
2
(G) whose Plancherel transform f

is equal almost everywhere to some element of C (G) is a norm-dense
c
linear subspace of L
2
(G). We denote this subspace by More-
c
over, if g,h E L
2
(G), then g * h E C (G), whence we deduce that
c
gh E Ll(G) n L
2
(G) and (gh) = g * h. Note that (gh) here is
actually the Fourier transform of gh. Furthermore, each f ( Ll(G)
can be written as f = f
l
f
2
where fk E L
2
(G) , k = 1,2, from
which it follows easily that (gh I g,h is norm dense in
L1(G), since is norm dense in L
2
(G).
However, (f I f E LI(G), f Cc(G)) (gh I g,h
whence we conclude that LI(G) is Tauberian. 0
Corollary 8.1.2. Let G be a locally compact Abelian topolo-
gical group. If I C Ll(G) is a proper closed ideal, then there
exists some maximal regular ideal Me LI(G) such that M I.
Of course, if G is discrete. then the corollary is trivial,
as Ll(G) has an identity.
Recalling the definition of the translation operators Ts de-
fined in Definition 4.7.1, we can apply Corollary 8.1.2 to obtain
the following interesting theorem:
8.1. Tauberian Commutative Banach Algebras 185
Theorem 8.1.3. Let G be a locally compact Abelian topological
group, let f E LI(G), and suppose I denotes the closed linear
subspace of LI(G) spanned by {Ts(f) I s E G). Then the following
are equivalent:

(ii) fey) 0, E G.
Proof. We claim first that I is a closed ideal in LI(G).
To see this it suffices to show that g * f I for each g L
1
(G)
since g * Ts(f) = Ts(g * f), s E G, and I is invariant under
the translation operators
( L (G) is such that
T , s E G.
s
To this end we note that, if
CD
then
IG dArt) = 0
f * = IG f(s - t),(t) dA(t)
= IG f(s + dArt)
= IG dArt)
= 0
that is, f * = O. Hence for any g LI(G) we have
iG g * dArt) = (g * f) *
= g * (f *
= o.
(h E I),
(s E G),
Thus, since the dual space of LI(G) can be identified with Lm(G)
[OSI' pp. 289 and 290; E
l
, pp. 215-220, 239 and 240], a consequence
of the Hahn-Banach Theorem [L, p. 90] shows that g * f I for
each g E LI(G). Hence I is a closed ideal in L
1
(G).
186 8. Ideal Theory
Now if I Ll(G), then I is a proper closed ideal in Ll(G),
whence by Corollary 8.1.2 there exists some maximal regular ideal
Me L1(G) such that M I. Thus from Theorem 4.7.3 we see that
A
there exists some y E G such that M = {g I g E Ll(G), g(y) = 0).
A
In particular, we would have fey) = 0, contradictiong the assump-

tion that f never vanishes. Thus part Cii) of the theorem implies
part (i).
A

Conversely, suppose fCY) = 0 for some y E G. An elementary
A A
computation reveals that T (f) (w)
s.
= (-s,w)f(w), s E G and w E G,
whence we deduce that T (f) (y) = 0,
A s
that hey) = 0, h E I, and so I is
A
ideal M = {g I g E L1CG), g(y) = oj.
part Ci) implies part (ii).
s E G. It is then obvious
contained in the maximal regular
Hence I is proper, and so
o
The fact established in the preceding proof that the closed
linear subspace I spanned by {Ts(f) I s (G) is a closed ideal
is a special case of a generally valid result in Ll(G): a closed
linear subspace I of LleG) is an ideal if and only if it is
translation invariant; that is, if and only if Ts(g) E I, s E G,
whenever gEl. We state this result as the next theorem leaving
the proof to the reader.
Theorem 8.1.4. Let G be a locally compact Abelian topologi-
cal group and let I C Ll(G). Then the following are equivalent:
(i) I is a closed ideal.
Cii) I is a closed translation-invariant linear subspace.
This is also an appropriate point to mention some approximation
results for LI(G) that are easy corollaries of the fact that Ll(G)
is Tauberian.
Corollary 8.1.3. Let G be a locally compact Abelian topologi-
cal group. If f E L1(G) and & > 0, then there exists some
A A
V E LI(G) such that v E Cc(G) and IIf - f * vIII < &.
8.1. Tauberian Commutative Banach Algebras 187
Proof. From Theorem 4.7.2 we know that LlCG) contains an
approximate identity, and so there exists some u E LICG) such that
IIf - f * ull
l
< 1/2. Then, s!nce LICG) is Tauberian, there exists
some v E LI CG) such that v ( C
C
and lIu - vIII < 1/2l1f Il
1

Note that we may assume that f 0 since the result is trivially
valid in this case. Hence
IIf - f * vIII IIf - f * ull
l
+ IIf * u - f * vIII
IIf - f * ull
l
+ IIflllliu - vIII

<-+-
2 2
= I.
o
Corollary 8.1.4. Let G be a compact Abelian topological
group. If f E LICG) and I > 0, then there exists a finite linear
combination of continuous characters of G, say Ey EG ayC ,y),
a
y
where only finitely many a
y
are nonzero, such that
IIf - f * ayc,y)]U
I
< c.
Proof. 8y Corollary 8.1.3 there exists some v E LlCG) such

that v . C
c
(G) and IIf - f * vIII < . Since G is discrete by
Corollary 4.7.5, the support of v must be finite, say Yl'Y2' 'y
n
It is then evident that v = Ik=lvCYk)X(Yk)' where X(Yk) denotes
the characteristic function of (Yk), whence, from the semisimplicity
of Ll(G) (Theorem 4.7.4) and Corollary 4.7.4, we conclude that

v = 0
A finite linear combination of continuous characters of a locally
compact Abelian topological group G is usually called a trigono-
metric polynomial. The origin of the terminology is clear on consi-
dering the group G = f.
Utilizing these corollaries it is not difficult to prove the
following result, the details being left to the reader:
188 8. Ideal Theory
Corollary 8.1.5. Let G be a locally compact Abelian topolo-
gical group. Then LI(G) contains an approximate identity (u
a
)
such that c C Moreover, if G is compact, then LI(G)
Ot c
contains an approximate identity consisting of trigonometric poly-
nomials.
Theorem 8.1.2 and the idea of the proof of Corollary 8.1.3 also
provide us with the following useful result:
Corollary 8.1.6. Let G be a locally compact Abelian topolo-

gical group. If KeG is compact and e > 0, then there exists
some f E LI(G) such that
(i)

f E C (G).
c
..
(ii) fey) = 1, Y E K

Proof. If G is compact, then from Corollary 7.3.3 we see
that G is discrete, and so Ll(G) has an identity e. Clearly
.. .
= I, Y E G, and lIeU
I
< 1 + e. On the other suppose
G is noncompact. Then there exists an open set W C G such that

W K, and cl(W) is compact. If E = G - W, then E is closed
and K n E =" whence, by Theorem 7.3.2 or Corollary 7.3.4, we

deduce the existence of some h Ll(G) such that hey) = 1, Y E K,
. .. .
and hey) = 0, y E E. In particular, h E Cc(G). Now let 6 >
be such that 6 < min(cllhlll,e).
From the proof of Theorem 4.7.2 we see that there exists some
u C LI(G) such that lIulil = 1 and IIh - h * ull
l
< 6/3, and, since
Ll(G) is Tauberian (Theorem 8.1.2), there exists some g E Ll(G)

such that g E C (G) and !lu - gli
l
< 6/3I1hIl1. Set f = h + g - h * g.
c ..
Evidently f E LI(G), f E Ce(G), and
- . . ..
fey) = hey) + g(y) - h(y)g(y)

= 1 + g(y) - gCy) = 1
(y E K).
8.2. Two Tauberian Theorems
189
Moreover,
< IIglll +
IIh - h
* gill
< IIglll + IIh -
h
* ull
l
+ IIh
* u - h *
< lIull
l
+
6 6
IIhlllliu - gill
311
h
l;1
+ - +
3
6 6 6
< I
+ 311hlll
+-
3
+3
< 1
e e e
+-+-+-
333
= 1 + e.
gill
c
8.2. Two Tauberian Theorems. The theorem of Tauber, from which
the name "Tauberian theorem" originates, is the following: Suppose
(a
k
) is a sequence of complex Rumbers such that the power series
~ O ak,k converges in (, 1 , E Q;, 1'1 < I). If limk_CDka
k
= 0
and
lim Ek=O akr
k
= a,
r-I
O<r<l
then Ek=o a
k
= a; that is, if a
k
= o(l/k) and tk=oa
k
is Abel
summable to a, then Ek=o a
k
converges to a. In 1932 Wiener
put this theorem, as well as all other theorems of a similar nature
known at the time, into a much more general framework involving
limits of convolutions. We shall content ourselves with proving
one of the main results of Wiener and a consequence of this result
due to Pitt. For further discussion of Tauberian theorems the read-
er is referred to [HR
2
, pp. 509-513jPij Wr
l
; Wr
2
]. In particular,
a proof using Theorem 8.2.1 of Littlewood's improvement of Tauber's
theorem, which replaces ~ = o(l/k) by ~ = O(l/k), can be found
in [HR
2
, pp. 512 and 513; Wr
I
, pp. 104-111].
190 8. Ideal Theory
For the sake of completeness we make one definition before giv-
ing Wiener's theorem.
Definition 8.2.1. Let G be a locally compact Abelian topolo-
gical group and suppose g is a complex-valued function defined on
G. Then limt_mg(t) = a if, given ~ > 0, there exists a compact
set KeG such that 19(t) - al < ~ t E G - K. A function g is
said to be slowly oscillating if, given ~ > 0, there exist a com-
pact set KeG and a compact neighborhood U of 0 in G such
that (g(t) - g(s)l < ~ t - s U and t E G - K.
Obviously every uniformly continuous function on G is slowly
oscillating.
Theorem 8.2.1 (Wiener's Tauberian Theorem). Let G be a locally
compact Abelian topological group and let g E Lm(G). Suppose there
exist some h E LI(G) and some a E C such that
,.. ..
(i) hey) ~ 0, y E G.
(ii) lim
t
-tg * h(t) = aJG h(u) dA.(u).
Then limt_cx$ * f(t) = aJG feu) dA,(u) for each f in LI(G).
Proof. We remark first that the limits are meaningful since
g * f, fELl (G), is a bounded continuous function on G by Proposi-
tion 4.7.2. Denote by I the set of all f E Ll(G) such that
lim
t
-tg * f(t) = aJG feu) dl.(u). Clearly I is a closed linear
subspace of LI (G), and I ~ (OJ because h E 1. Moreover, I
is translation invariant since, if f E I, then
lim g * T (f) (t) = lim T (g * f) (t)
t-t S t-m s
= lim g * f (t - s)
t-t
= lim g * f(t)
t-m
8.2. Two Tauberian Theorems 191
= aiG feu) dA.(u)
= aiG Ts(f)(u) dA.(u)
(s E G)
Thus, in particular, I contains the closed linear subspace of

LI(G) spanned by {Ts(h) I s E G). Since hey) ~ 0, y E G, it
follows immediately from Theorem 8.1.3 that I = LI(G). This com-
pletes the proof.
o
Wiener's original result was, of course, only for G = R, and
his proof is quite intricate. It was in proving this theorem that
Wiener needed the result about absolutely convergent Fourier series,
now generally known as Wiener's Theorem (Theorem 4.6.2), which we
have already discussed. Wiener's proof is available in [Wr
l
, pp. 72-
97] (see also [Pi, pp. 43-92]).
We note that Corollaries 8.1.1 and 8.1.2 and Theorem 8.1.3 are
also referred to as Wiener's Tauberian Theorem.
It may not be completely clear why Wiener's Tauberian Theorem
carries such a name, and we shall not pursue the reasons in any de-
tail. The following theorem should, however, provide some support
for the terminology:
Theorem 8.2.2 (Pitt). Let G be a locally compact Abelian
topological group and let g be a bounded, measurable, slowly oscil-
lating function on G. Suppose there exist some hELl (G) and
some a E C such that

(i) hey) ~ 0, y E G.
(ii) lim
t
~ * h(t) = aiG h(u) dA.(u).
Proof. Considering g as an element of L (G),
CII
we see at
once from Wiener's Tauberian Theorem that
192 8. Ideal Theory
lim g f{t) = aiG feu) dA. (u)
t .. CD
Now, given > 0, since g is slowly oscillating, there exist a
compact set KI C G and a compact neighborhood U of 0 G such
that Ig(u) - g(v)1 < e/2, u - v E U and u E G - K
I
. Consequently,
if fo = Xu/A.(U), then fo (L1(G) and
Ig(t) - g * fo (t) I = 19(t) - rlur J
U
get - s) dA(s) I
~ 1 ~ ) Iu 19(t) - get - s) I dA.(s)
e
<-
2
(t E G - K
I
).
But since
lim g * fo (t) = aI
G
fo (u) dA.(u)
t-m
= a,
there exists some compact set K? C G such that
..
Therefore, if K = Kl U K
2
, we see that 1get) - al < i,
t E G - K, which proves the theorem.
o
8.3. The Problem of Spectral Synthesis. Given a commutative
Banach algebra A, it would obviously be of considerable interest
to be able to describe all the closed ideals in A. In general this
is apparently a hopeless task, although we shall see shortly that
for some specific Banach algebras rather simple descriptions of the
closed ideals may be available. Some progress can be made on the
general question by considering the related problem of determining
when a closed subset of the maximal ideal space 6(A) can be the
hull of more that one closed ideal in A. In this and the following
sections we shall concentrate our attention mainly on this question.
8.3. The Problem of Spectral Synthesis
193
First, let us see what relation this sort of problem has to the
description of closed ideals. Let A be a semisimple regular commu-
tative Banach algebra and suppose E C 6{A) is closed. Then from
Theorem 8.1.1 we see that E = h(cl[1 (E)]) = h(cl[J (E)]) and that
o 0
cl[lo(E)]
recall that
is the smallest closed ideal in A whose hull is

I (E) consists of all the x ~ such that x
o
identically on some open set o C A(A)
x
such that 0 ~ E.
x
E. We
vanishes
On the other hand, from the definition of the hull-kernel topo-
logy and Proposition 7.1.1 we see that keEl is a closed ideal in
A such that h[k(E)] = E, and if I C A is a closed ideal for
which h(l) = E, then I C keEl; that is, keEl is the largest
closed ideal whose hull is E.
Thus, if there were precisely one closed ideal I such that
h(l) = E, then we would have to have 1= cl[I (E)] = keEl = k[h(I)].
o
In particular, I would consist of those x E A such that X(T) = 0,
T h(l) = E.
We summarize these observations in the following proposition:
Proposition 8.3.1. Let
Banach algebra. If E C A(A)
equivalent:
(i) cl[Io(E)] = keEl.
A be a semisimple regular commutative
is closed, then the following are
(ii) If I and J are closed ideals in A such that h(l) =
h(J) = E, then I = J.
(iii) If I is a closed ideal in A such that h(l) = E, then
I = k[h(I).
However, it may very well be the case that, given a closed set
E in 6(A), there exist distinct closed ideals I and J such
that h(l) = h(J) = E, equivalently, so that CI[lo(E)] ~ keEl.
We shall mention an example of this in L
I
QR3) following Theorem 8.3.4.
194 8. Ideal Theory
With this in mind we make the following definition:
Definition 8.3.1. Let A be a semisimple regular commutative
Banach algebra. Then a closed set E in ~ A ) is said to be a
set of spectral synthesis if I = keEl is the only closed ideal in
A such that h(l) = E.
Proposition 8.3.1 provides several equivalent formulations of
this definition. In particular we note that, if E is a set of spec-
tral synthesis for A and I is a closed ideal such that h(I) = E,
then I = k[h(I)]; that is, I is the intersection of all the maxi-
mal regular ideals in A containing I.
The problem of determining which closed sets in A(A) are sets
of spectral synthesis is called the problem of spectral synthesis.
The development of Section 8.1 provides us with a necessary
and sufficient condition that the empty set be a set of spectral
synthesis.
Theorem 8.3.1. Let A be a semisimple regular commutative
Banach algebra. Then the following are equivalent:
(i) A is Tauberian.
(ii) ~ is a set of spectral synthesis.
Proof. If
E = ,
then it is immediately apparent that J (E)
0

is precisely the set of all x A such that x C ~ A ) ) .
c
But
A is Tauberian if and only if cl[J (E)] = cl[I (E)] = A, whence
o 0
we conclude, since
keEl = ~ ) = A, that
E =
is a set of
spectral synthesis if and only if A is Tauberian.
o
In general the problem of spectral synthesis is a complicated
one. Before we pursue it further in the abstract we wish to consi-
der several specific examples. First, we shall describe completely
the closed ideals in C(X), X being a compact Hausdorff topological
8.3. The Problem of Spectral Synthesis
space. An immediate corollary of this result will be that every
closed subset of X is a set of spectral synthesis for C(X).
Theorem 8.3.2. Let X be a compact Hausdorff topological
space and suppose I c CeX). Then the following are equivalent:
(i) I is a closed ideal in C(X).
(ii) There exists a closed set E C X such that I = keEl =
{f I f C(X), f(t) = 0, t E EJ.
Proof. Obviously part (ii) of the theorem implies part (i),
195
so suppose that I is a closed ideal in C(X). Ultimately we shall
wish to apply the Stone-Weierstrass Theorem [L, p. 332] to I, and
with this is mind we shall first show that I is closed under com-
plex conjugation.
Indeed, let f I and suppose s > O. Then define gs C(X)
by gs(t) = f(t) [f(t)]2/(S + If(t)1
2
), t E X. Clearly gs E I, as
I is an ideal, and
Hence
clf(t)1
c + If(t)1
2
<./s
-2
The validity of the last estimate is easily verified.
(t E X).
(t E X).
196 8. Ideal Theory
Consequently we see that lig - 11 < .//2, from which it fol-
t: CX)-
lows that f E I, as I is a closed ideal. Thus I is closed
under complex conjugation.
Now let E = h(I). From Theorem 7.3.1 we see that I is a
commutative Banach algebra whose maximal ideal space a(I) is homeo-
morphic to X - h(I) = X - E and that the Gel'fand transform of f,
as an element of the Banach algebra I, is just the Gel'fand trans-
form of f, as an element of C(X), restricted to X - E; that is,
the Geltfand transform of f E I is just the restriction of f to
X-E. Thus, since X - E is open and each f in I vanishes
identically on E = h(I), we see that, if f E I, then f is a
continuous function on X vanishing identically on E. It is appar-
ent that the set of all such functions in C(X) can be identified
with C (X - E). Hence the proof will be complete if we can show
o
that I, considered as a subalgebra of C (X - E), is actually all
o
of Co(X - E).
Evidently I is a closed subalgebra of C (X - E), which is
o
closed under complex conjugation. Moreover, I separates the points
of X - E because, if t,s E X - E, t s, then, since C(X) is
regular, there exists some g E C(X) such that get) = 1, g(s) = o.
Since t E = h(I), there exists some f E I such that f(t) O.
Then fg E I, fg(t) 0, and fg(s) = 0, and so I separates
points. Finally, there exists no t X - E such that f(t) = 0,
f E I, because such a t would have to be in h(l) = E. Thus all
the hypotheses of the Stone-Weierstrass Theorem [L, p. 332] are ful-
filled, and so I = C (X - E) = (f I f E C(X), f(t) = 0, t E EJ.
o
This completes the proof.
Corollary 8.3.1. Let X be a compact Hausdorff topological
space and suppose E C X. Then the following are equivalent:
(i) E is closed.
(ii) E is a set of spectral synthesis for C(X).
c
8.3. The Problem of Spectral Synthesis 197
Moreover, every proper closed ideal I in C(X) is the intersection
of the maximal ideals containing I , that is, I = k[h(I)].
Proof. If E is closed and I is a closed ideal in C(X)
such that h(l) = E, then, by Theorem 8.3.2 and the regularity of
C(X), we see that there exists some closed set F C X such that
I = kef) and E = h(l) = h[k(F)] = F, that is, I = keEl. Thus
E is a set of spectral synthesis for C(X).
o
The analog of Theorem 8.3.2 is also valid for LI(G) provided
G is a compact Abelian topological group. Recall that, since G

is compact, the dual group G is discrete (Corollary 4.7.5), and

hence every subset of G is closed.
Theorem 8.3.3. Let G be a compact Abelian topological group
and suppose Ie LI(G). Then the following are equivalent:
(i) I is a closed ideal in LI(G)

(ii) There exists a set E C G such that I = keEl =

(f I f (LI(G), fey) = 0, y E EJ.
Proof. As before, we need only prove that part (i) implies
part (ii). So let I be a closed ideal and let E = h(l). Then
from Proposition 7.1.I(vii) we see that I C k[h(I)] = keEl.

To establish the reverse containment we note that, if y E G - E,

then (.,y) E I. This follows from the fact that, if y E G - E =

G - h(I), then there exists some gEl such that g(y) = 1. Since
G is compact, we see that (.,y) L
1
(G) and
g *
(,y)(t) = IG (t - s,y)g(s) dA(s)

=
(t,y)g(y)
=
(t,Y) (t E G).
Thus (.,y) E I, as I is an ideal.
198 8. Ideal Theory
Consequently I contains all the trigonometric polynomials of
the form 1:
y
(G _ E a
y
(. ;v) . Hence JI if f ( k [h (1)], then we see
that, for any trigonometric polynomial 1:
y
E a a
y
(. , y),

= 1:.a fey) (t,y)
yEG
Y

~ a"f(y) (t,y)
yEG-E
=
(t ( G),
..
since fey) = 0, y E E = h(I) = h(k [h(l)]). Thus f * [1:
y
E aay (. ,y)]
belongs to I for each f k[h(I)] and each trigonometric poly-
nomial. But then, appealing to either Corollary 8.1.4 or 8.1.5, we
conclude that f E I, as I is closed, that is, k[h(I)] C I.
Therefore I = k[h(I)] = k(E), and part (i) implies part (ii).O
Corollary 8.3.2. Let G be a compact Abelian topological group .

If E C G, then E is a set of spectral synthesis for Ll(G). More-
over, every proper closed ideal I in LI(G) is the intersection
of the maximal regular ideals containing Ii that is, I = k[h(I)].
An examination of the proof of Theorem 8.3.3 reveals that, if
I is a closed ideal in LI(G), G being a compact Abelian topologi-
cal group, then every f E I is the limit in LI(G) of trigono-

metric polynomials of the form EYEG-h(I) ayf(Y)C"Y) ~ d that
all such polynomials belong to I. Thus, since oCf) = Ref) or

oCf) = Ref) u (oj, fELl (G), depending on whether G is finite
or infinite, we see that the elements of I are determined by those

y E G such that fCY) ~ 0; that is, by those y E G such that

fCy) E oCf) and fCy) ~ O. Hence, loosely speaking, we see that
an element fEr can be recovered or synthesized from a knowledge
of the nonzero numbers in oCf). These remarks should supply some
insight into the origin of the term "spectral synthesis".
8.3. The Problem of Spectral Synthesis
199
This basis for the terminology is even more apparent if one
considers the Banach algebra L
2
(G) in place of Ll(G), G being
a compact Abelian topological group, since then one has all the
machinery of Hilbert space theory available. For a related, but
somewhat different, motivation the reader is referred to [Ru
l
, pp.
183-186].
If G is a noncompact locally compact Abelian topological
group, then the structure of the closed ideals of LICG) is vastly
more complicated than in the case of compact groups, and it is no

longer the case that every closed subset of G is a set of spectral
synthesis. Moreover, in this case there exist no known necessary

and sufficient conditions for a closed set E C G to be a set of
spectral synthesis. Since it would entail a considerable digression
to investigate in any detail the problem of spectral synthesis in
LICG), we shall content ourselves here with some observations on
the problem without any proofs. Some additional results will be
proved in Section 8.6.
First we note that the failure of spectral synthesis in LICG),
G being a noncompact locally compact Abelian topological group, is
universal; that is, it occurs for all such groups. This is a cele-
brated result of Malliavin, which we state as the next theorem.
Theorem 8.3.4 CMalliavin). Let G be a noncompact locally
compact Abelian topological group. Then there exists some closed

set E C G that is not a set of spectral synthesis for LICG).
Prior to Malliavin's theorem it was known that certain groups
contained closed sets that were not sets of spectral synthesis. For
example, L. Schwartz showed that E = {Ct
l
,t
2
,t
3
) I t ~ + t ~ + t ~ = 11
is such a set in JR3. The analogous assertion is actually valid
in ~ n ~ 3, but false in Rand Ff. On the other hand, sets
that fail to be sets of spectral synthesis can be fairly complicated .

For instance, if G is noncompact and 0 eGis open and nonempty,
then 0 contains a closed subset E that is not a set of spectral
200 8. Ideal Theory
synthesis and such that E is homeomorphic to the Cantor ternary
set. Nevertheless the Cantor ternary set itself is a set of spec-
tral synthesis in Ll (lR)
Simple operations with sets of spectral synthesis may not pro-
duce sets of spectral synthesis. For example, the intersection of
sets of spectral synthesis may not be such a set. Indeed, the sets
and
E2 = {(t
l
,t
2
,t
3
) I t ~ + t ~ + t; ~ I}
are both sets of spectral synthesis in L l ~ ) but, as we have
noted, E = El n E2 is not. On the other hand, the union of two
disjoint sets of spectral synthesis is again a set of spectral syn-
thesis. It is unknown whether the same is true for nondisjoint sets
of spectral synthesis.
More complete discussions of the problem of spectral synthesis
in Ll(G) can be found, for instance, in [HR
2
, pp. 484-605; Ru
l
,
pp. 157-191].
We now wish to return to the problem in general semisimple
regular commutative Banach algebras.
8.4. Local Membership in Ideals. In the next section we shall
prove Ditkin's Theorem, which provides the most general known suffi-
cient condition for a closed set to be a set of spectral synthesis.
First, however, we need some preliminary results which supply suffi-
cient conditions for an element x in a semisimple regular commuta-
tive Banach algebra A to belong to a given ideal I in A. The
role of these results in the problem of spectral synthesis will be-
come apparent in the next section. We begin with some definitions.
8.4. Local Membership in Ideals
201
Definition 8.4.1. Let A be a commutative Banach algebra and
let I be an ideal in A. An element x (A is said to belong
locally to I at the point T a(A) if there exist an open neigh-

borhood U of T and some y I such that x(w) = yew), w E U.
If a(A) is noncompact, then x is said to belong locally to I
at infinity if there exist a compact set K c a(A) and some y E I

such that x(w) = w ( a(A) - K.
The theorem of this section asserts that, if A is a semisimple
regular commutative Banach algebra and x A belongs locally to
an ideal I at every point in a(A) and at infinity when a (A) is
noncompact, then x belongs to I. Before proving this result we
need a lemma about continuous functions on normal topological spaces.
Lemma 8.4.1. Let X be a normal topological space and suppose
A C C(X) is a subalgebra such that
Ca) A contains the constant functions.
(b) If E
l
,E
2
are disjoint closed subsets of X, then there
exists some f E A such that f(t) = 1, t (E
I
, and f(t) = 0,
t E
2

If E C X is closed and E C where Uk C X is open,
k : 1,2, ,n, then there exist hk A, k : 1,2, . ,n, such that
(ii) = 0, t E X - Uk' k = 1,2, ... ,n.
Proof. The lemma is evidently valid for n = 1 on setting
El = E, E2 = X - U
l
and applying the hypotheses of the lemma. If
n = 2, then E C U
I
U U
2
, and so E n (X - U
2
) C U
l
is closed.
Since X is a normal topological space, there exists an open set
U C X such that E n eX - U
2
) cue cl(U) cUI (see, for example,
[W
2
' p. 49]). Then cl(U) and X - U
1
are disjoint closed subsets
of X, and so there exists some hI A such that hI (t) = I,
202 8. Ideal Theory
t E cl(U), and hl(t) = 0, t E X - U
l
Moreover, it is apparent
that E and X - (U U U
2
) are also disjoint closed subsets of X,
whence there exists some f E A such that f(t) = I, tEE, and
f(t) = 0, t X - (U U U
2
). Set h = f - fh
2 1
Clearly h2 E A,
and if tEE, then
= 1,
as f(t) = 1, t E. Furthermore, hl(t) = 0, t E X - U
1
' by the
choice of hI' whereas if t E X - U
2
' then either t E (X - U
2
) nu
or t E X - (U U U
2
). In the first case t E cl(U), and so
h
2
(t) = f(t) - f(t)hl(t)
= f(t) - f(t) = 0,
as hl(t) = 1, t E cl(U). In the second case h
2
(t) = 0, as
f(t) = 0, t E X - (U U U
2
). Thus hk(t) = 0, t E X - Uk' k = 1,2.
Now suppose the conclusion of the lemma is valid for n - 1,
where n > 3, and suppose E C Then F = E n (X - Un) is
closed and F C Again appealing to the normality of X,
we deduce the existence of an open set U such that
n-I
Feu C cl(U) C U Uk"
k=l
Then, by the induction hypothesis, there exist gk E A, k = 1,2, ... ,n-I,

such that = 1, t E cl(U), and gk(t) = 0, t ( X - Uk'
k = 1,2, . ,n - 1. Moreover, E C U U U, and so, applying the
n
lemma with two open sets, we see that there exist f and h in
n
A such that f(t) + h (t) = I, t (E, and f(t) = 0, t E X - U,
n
and hn(t) = 0, t ( X - Un. Hence on setting hk = fg
k
, k = 1,2, .. ,n-l
it is easily verified that tk=lhk(t) = 1, tEE, and = 0,
t E X - Uk' k = 1,2, .. ,n.
This completes the proof of the lemma.
o
8.4. Local Membership in Ideals
203
The reader should observe that the lemma is actually valid with-
out any assumptions of continuity on the functions in A. However,
our applications of the lemma will only be in the case where A
consists of continuous functions.
Theorem 8.4.1. Let A be a semisimple regular commutative
Banach algebra and suppose I C A is an ideal. If x E A belongs
locally to I at every point of A (A) and at infinity when A(A)
is noncompact, then x E I.
Proof. We claim that, without loss of generality, we may assume
that A has an identity. If this were not so, then I would still
be an ideal in A[e] and x would belong locally to I at all
points of A(A[e]) = A{A) U {T J. Indeed, it is obvious that x
CD
belongs locally to I at each point of A(A), and x belongs
locally to I at T since x belongs locally to I at infinity
CD
and Y(T) = 0, yEA. Thus if the theorem is valid for algebras
CD
with identity, we conclude that x E I. Consequently we may assume
that A has an identity.
Then A(A) is a compact Hausdorff, and so normal, topological
space, and A is a normal Banach algebra by Corollary 7.3.3. Thus

we may apply Lemma 8.4.1 to A C C(A(A)).
Now x belongs locally to I at each point in A(A). Hence,
given T E A(A), there exist an open neighborhood U of T and
T
some YT E I such that x{w) = YT(w), w E UTe Clearly the sets
(U
T
I T E A{A)) form an open covering of A(A), and so there exists
a finite number of the U
T
, call them U
l
,U
2
, .. ,U
n
, such that
A(A) = Uk=IU
k
For the sake of notational simplicity we denote the
YT corresponding to Uk by Y
k
, k = 1,2, ... ,n. Applying Lemma
8.4.1 to the closed set E = A(A), we deduce the existence of x
k
E A,
k = 1,2, .. ,n, such that = 1, T E A(A), and = 0,
T - Uk' k = 1,2, .. ,n. Evidently E I. Moreover,
we claim that
204 8. Ideal Theory
"
X (1') (1' (
Indeed. let l' (
Uk such that l' E Uk.
and let Uk ,Uk .. ,U
k
denote those
" .. 12m.
Now Yk.(w) = x(w). W (Uk.' J = 1,2, ... ,m,
J J
so in particular these equations hold for W = 1'. On the other hand,
" "
x
k
(1') = 0 if k k
j
, j = 1,2, ... ,m, since xk(w) = 0, W E - Uk.
Hence we see that
" n "
= x (1') 1: x
k
. (1')
j=l J
"
= x(1').
Therefore, since A is semisimple, we conclude that x =
belongs to 1.
o
8.5. Ditkin's Theorem. We are now almost in a position to
prove the theorem alluded to at the beginning of the preceding section.
To describe the relation between Ditkin's Theorem and spectral syn-
thesis we first need to make another definition.
Definition 8.5.1. Let A be a commutative Banach algebra.
Then A is said to satisfy Ditkin's condition at l' in if,
whenever x E A and l' E Z(x), there exist a sequence (x
k
) C A
and open neighborhoods Uk of T such that
(i) - xII = o.
"
(ii) xk(w) = 0, wE Uk' k = 1,2,3, ....
If d(A) is noncompact, then A is said to satisfy Ditkin's condi-
tion at infinity if, whenever x E A, there exists a sequence
{xkl C A such that
8.5. Ditkin's Theorem 205
A
(b) x
k
E Cc(A(A)), k = 1,2,3, ..
A
We remind the reader that Z(x) = (T I T E a(A), X(T) = 0)
(Definition 7.1.1). It is apparent that, if A satisfies Ditkin's
condition at infinity, then A is Tauberian.
An immediate corollary of Ditkin's Theorem will show that, if
A is asemisimple regular commutative Banach algebra that satisfies
Ditkin's condition at each point of A(A) and at infinity when A(A)
is noncompact, then a closed set E C A(A) will be a set of spectral
synthesis for A provided that bdy(E) contains no nonempty perfect
set. Similarly it follows that, if I is a closed ideal in such a
Banach algebra, then I is the intersection of the maximal regular
ideals containing I, that is, I = k[h(I)], provided bdy[h(I)]
contains no nonempty perfect set.
For the sake of completeness we recall that, if E is a closed
subset of a topological space X, then bdy(E) = E n cl(X - E) =
E n [X - int(E)], and a set E is perfect if E is closed and
every point of E is a limit point of E, that is, E has no iso-
lated points.
One further lemma is necessary before we can prove Ditkin's
Theorem.
Lemma 8.5.1. Let A be a semisimple regular commutative Banach
algebra, let I be a closed ideal in A, and suppose x E A. Then
x belongs locally to I at each T in a(A) that satisfies either
of the following conditions:
(i) T E int[Z(x)].
(ii) T E a(A) - h(I).
Proof. Suppose T E int[Z(x)]. Then there exists an open
206 8. Ideal Theory

neighborhood U of T such that U c Z(x), and so x(w) = O(w) = O.
w E U. Thus x belongs locally to I at T.
Now suppose T ( A(A) - h(I). Then, since A(A) is locally
compact, there exists an open neighborhood U of T such that
cl(U) is compact and cl(U) C A(A) - h(I). Since cl(U) is compact,
h(l) is closed, and cl(U) n h(l) =" we see from Theorem 7.3.2

that there exists some y E I such that yew) = 1, w E cl(U). Hence

xy E I and (xy) (w) = x(w)y(w) = x(w), w E U. Therefore x belongs
locally to I at T.
o
Theorem 8.5.1 (Ditkints Theorem). Let A be a semisimple
regular commutative Banach algebra that satisfies Ditkin's condition
at each point of A (A) and at infinity when A (A) is noncompact,
let I be a closed ideal in A, and suppose x E A is such that
h(l) C Z(x). Then
(i) If E is the set of T E A(A) such that x does not
belong locally to I at T, then E is a perfect subset of
h(l) n bdy[Z(x)] = bdy[h(I)] n bdy[Z(x)].
(ii) If bdy[h(I)] n bdy[Z{x)] contains no nonempty perfect
set, then x ( I.
Proof. From Lemma 8.5.1 and the fact that Z(x) is closed we
see at once that E C h(l) n - int[Z(x)]). However, since
h(l) C Z(x), we have
bdy[h{I)] nbdy[Z(x)] = (h{I) n (Z{x)
= h(l) n (Z{x)
= h{l) nbdy[Z(x)]
h (I) n [Z (x) n (6 (A) - int [Z (x)])]
= h{l) n (6(A) - int [Z{x)]).
8.5. Ditkin's Theorem 207
Consequently E C h(l) n bdy[Z(x) = bdy[h(I)] n bdy[Z(x). Moreover,
from the definition of local membership in an ideal we see at once
that E is closed. Thus to show that E is perfect we need only
prove that E has no isolated points.
So suppose that T E is isolated. Then there exists some
open neighborhood U of T such that cl(U) is compact and
(cl(U) - (T) n E = ,. Since E C bdy[Z(x)] c Z(x), we see that
T E Z(x). Thus, by Ditkin's condition, there exist a sequence (x
k
)

in A and open neighborhoods Uk of T such that xk(w) = 0,
wE Uk' k = 1,2,3, .. , and limkllxxk - xII = O. Let W be an open
neighborhood of T such that cl(W) C U. This is possible because
A(A) is locally compact. Then cl(W) is compact, A(A) - U is
closed, and cl(W) n [A(A) - U] = " whence, since A is normal,

we deduce the existence of some y A such that yew) = I, w E cl(W),

and yew) = 0, w ~ A(A) - U. We claim that the elements yxx
k
,
k = 1,2,3, , belong locally to I at each point of A(A) and at
infinity.

Indeed, since xk(w) = 0, w E Uk' and T E Uk' we see that
yxx
k
belongs locally to I at T; and if w cl(U) - (T), then
by the definition of E we see that x belongs locally to I at
w, whence yxx
k
belongs locally to I at w, k = 1,2,3, .... Thus

each yxx
k
belongs locally to I on cl(U). Finally, since y
vanishes identically on A(A) - U and cl(U) is compact, we see
at once that yxx
k
' k = 1,2,3, .. , belongs locally to I at each
w E A(A) - cl(U) C A (A) - U and at infinity. Hence yxx
k
belongs
locally to I at each point of A(A) and at infinity, k = 1,2,3, ..
Therefore (yxx
k
) C I by Theorem 8.4.1.
Consequently yx I, as I is closed and limkllyxxk - yxil = O.

But if w W, then (yx) (w) = y(w)x(w) = x(w), as y ~ ) = I,
w E cleW). Since T E W, this says that x belongs locally to I
at T, contradicting the fact that TEE.
Therefore E is perfect and part (i) of the theorem is proved.
208 8. Ideal Theory
To prove part Cii) we suppose that bdy[hCI)] n bdy[Z(x)] con-
tains no nonempty perfect sets. Then from part (i) we see that x
belongs locally to I at each point of a(A) , as E = ~ . If aCA)
is compact, then x E I by Theorem 8.4.1. On the other hand, if
a(A) is noncompact, then, since A satisfies Ditkin's condition

at infinity, there exists a sequence (x
k
) C A such that (xk) is
contained in C
c
(6(A and limklixxk - xII = O. For each k let
Ek be the set of T E a(A) such that xX
k
does not belong locally
to I at T. We claim that Ek = ~ k = 1,2,3, ...
Indeed, since x belongs locally to I at each point of a(A),
we see that, if T (A(A), then there exist an open neighborhood U

of T and some y E I such that x(w) = yew), w (U. Clearly

then we must also have (xx
k
) (w) = (YX
k
) (w), w E U, and YX
k
E I,
k = 1,2,3, ... ; that is, xX
k
belongs locally to I at T,
k = 1,2,3, Hence Ek C E = k = 1,2,3, .
Thus xX
k
belongs locally to I at each point of 6(A),
-
k = 1,2,3, ... Moreover, since (x
k
) C CcCaCA, it is apparent
that xX
k
also belongs locally to I at infinity, k = 1,2,3, ...
Consequently by Theorem 8.4.1 we conclude that (XX
k
) C I, whence
x E I, as I is closed.
It should be noted that the hypothesis that A satisfies Dit-
kin's condition at infinity was only utilized in proving part (ii)
of the theorem. Thus Theorem 8.S.1(i) remains valid under the weak-
er assumption that A satisfies Ditkin's condition only at each
point of a(A).
Some easy consequences of Ditkin's Theorem are the following
corollaries:
Corollary 8.5.1. Let A be a semisimple regular commutative
Banach algebra that satisfies Ditkin's condition at each point of
a(A) and at infinity when a(A) is noncompact. If E C a(A) is a
8.6. LI(G) Satisfies Ditkin's Condition
209
closed set such that bdy(E) contains no nonempty perfect set, then
E is a set of spectral synthesis for A.
Proof. Let I be any closed ideal such that h(I) = E. If
x E k[hel)] = keEl, then obviously E = hel) c Zex), and
bdy(E) = bdy[h(I)] bdy[h(I)] n bdy[Z(x)]. Since bdy(E) contains
no nonempty perfect set, it follows that bdy[h(I)] n bdy[Z(x)] has
the same property, whence we conclude that x E I by Ditkin's Theorem.
Hence I = k[h(I)] = k(E), and E is a set of spectral synthesis'
D
Corollary 8.5.2. Let A be a semisimple regular commutative
Banach algebra that satisfies Ditkin's condition at each point of
a(A) and at infinity when a (A) is noncompact. If I C A is a
closed ideal such that bdy[h(I)] contains no non empty perfect set,
then I is the intersection of the maximal regular ideals contain-
ing I; that is, I = k[h(I)].
It is perhaps worthwhile mentioning explicitly that, if A is
a semisimple regular commutative Banach algebra, then a (A) is com-
pact if and only if A has an identity. This is the content of
Corollary 7.3.2. Thus in the theorems of this and the preceding
section a(A) is noncompact precisely when A is without identity.
In the next section we shall show that LI(G), G being a locally
compact Abelian topological group, satisfies Ditkin's condition, and
so we may apply Ditkin's Theorem to this algebra.
8.6. L1(G) Satisfies Ditkin's Condition. Before we prove
that LI(G), G being a locally compact Abelian topological group,

satisfies Ditkin's condition at each point of a(LI(G)) = G and

at infinity when G is noncompact, we shall establish two lemmas .

As usual, ij will denote Haar measure on the dual gruop G, and

we shall assume that the Haar measures on G and G have been so
chosen that Plancherel's Theorem (Theorem 7.2.1) is valid.
210 8. Ideal Theory
Lemma 8.6.1. Let G be a locally compact Abelian topological

group, let W C G be an open neighborhood of the identity y in
0
G, let 6 > 0, and let E C G be compact. Then there exist a

compact set KeG with nonempty interior and an open symmetric

neighborhood a of Yo in G such that
(i) cl(U) is compact.
(ii) y int{K).
o
(iii) - U) <
(iv) K + U - U C w.
(v) 11 - (t,y)1 < 6, tEE and y E K + U - u.
Proof. As observed at the beginning of the proof of Theorem

4.7.5, the mapping from G x G to r defined by (t,y) = (t,y),

t E G and y E G, is continuous. Thus for each tEE there

exist open neighborhoods V
t
of t in G and W
t
of Yo in G
such that 11 - (s,y)1 < 6, s E V
t
and y E W
t
. Since E is compact
and (V
t
I t E EJ forms an open covering of E, we deduce the
existence 0: t
l
,t
2
, ... ,t
n
in E such that E C Set
WI =.W n (Ok=lW
tk
). Evidently WI is an open neighborhood of Yo
in G, and 11 - (t,y)1 < 6, tEE and y E WI. Moreover, using
the local compactness of G and the argument of Corollary 7.2.2,

we easily deduce the existence of a compact set Kl C G and a sym-

metric open neighborhood U of y in C such that
o
(b) cleU) is compact.
If Tt(K
l
- U) < then we set

the case and G is discrete, then we set
K = K
l
. If this is not
K = (YoJ. Finally, if

G is not discrete, then J) = 0, and we readily deduce from
o
8.6. Ll(G) Satisfies Ditkin's Condition

the regularity of ~ that there exists some compact set KeG
such that K C Kl n U, Yo E int(K), and ~ K - U) ~ 2 ~ U ) . as
o < ~ U ) < m by Proposition 4.7.3.
211
Furthermore, it is apparent that with these choices of
have K + U - U C WI C Wand II - (t,y)l < 6, tEE and
y E K + U - U.
K we
o
Lemma 8.6.2. Let G be a locally compact Abelian topological

group, let w E G, let W be an open neighborhood of w in G,

and let 1 > O. If f E LI(G) is such that few) = 0, then there
exists some k (LI(G) such that

(i) k is identically one on some open neighborhood of w.

(ii) key) = 0, y E G - W.
(iii) IIf * kll} < s.

Proof. In view of the fact that [(.,w)g] (y) = g(y - w),

y (G and g E LI(G), we may assume without loss of generality
that w=Yo. Let 0<6<,/[4I2(1+lIf ll})]. Since fEL
1
(G),
there exists a compact set E C G such that
I
G
-
E
If(t)1 dAft) < 6.
..
Moreover, by Lemma 8.6.1 there exist a compact set KeG and an

open symmetric neighborhood U of y in G such that
o
(i) cl{U) is compact.
(ii) Y (int(K).
o
(iii) ~ K - U) < 2 ~ U ) .
(iv) K + U - U c w.
(v) 11 - (t,y)1 < 6, t E and y E K + U - u.
212 8. Ideal Theory
By Plancherel's Theorem (Theorem 7.2.1) there exist k
1
,k
2
L
2
(G)

such that kl = X
K
_ U and k2 = XuI'Il (U) , where, as usual, X
K
-
U
and Xu denote the characteristic functions of K - U and U, res-
pectively. The transform involved is, of course, the Plancherel
transform. Set k = k
l
k
2
Then k LI(G) and the Fourier trans-

form of k is k = X
K
_
U
* as was seen in the proof of
Theorem 7.2.3. The argument of that theorem also shows immediately

that key) = 1, Y K, and key) = 0, y E G - (K + U - U) ::> G - W.
Thus it remains only to show that IIf * kill < s.
To this end we note first that
f * k(t) = IG f(s)k(t - s) dA(s)
= IG f(s)[k(t - s) - k(t)] dA(s)
(t ( G)

because f(yo) = O. Hence
IIf * kl/l = IGlfG f(s) [k(t - s) - k(t)] dA{s)1 dArt)
If{s)I[I
G
lTs{k){t) - k{t)l dArt)] dA(s)
= I
G
-
E
If{s)IIlTs(k) - kill dA{s)
+ IE If(s)IIITs{k) - kill dA.{s).
Thus, using Plancherel's Theorem (Theorem 7.2.1) and the fact that
- U) < we see that
I
G
-
E
If{s) 1IITs{k) - kill dA(s) < 2I1kIl
I
I
G
_
E
If(s) I dA.(s)
< 2611klk2111
< 2611klli21/
k
2
112
- .
= 2611klil2l1k2112
Xu
=
8.6. LI(G) Satisfies Ditkin's Condition
=
< 2/26.
In order to estimate the second integral we observe that
Ts(k) - k = T
s
(k
1
)T
s
(k
2
) - klk2
213
= k
I
[T
s
(k
2
) - k
2
] + [Ts(k
1
) - k
I
]T
s
(k
2
) (s G).
Furthermore, from Plancherel's Theorem we see that, if s (E, then
= II-=--(s-, - kIII2
= [Ia 11 - (s,y)1
2
x
K
_
u
(Y)
= [I
K
-
u
II - (s,y)12
< _ U)]1/2
because 11 - (s,y)l < 6, sEE and y E K + U - U. Similarly,
we see that for each sEE
II
T
s
(k
2
) - k2112 < 6 1/2.
[ll (U) ]
Combining these estimates with the facts that IIkl1l2 = _ U)]1/2
and IIk2112 = we deduce that
IITs(kJ - kill < IIk
l
[T
s
(k
2
) - k
2
]11
1
+ II[T
s
(k
I
J - k
l
]T
s
(k
2
)lI
l
IIkI1l2I1Ts(k2) - k21i2 + IITs(k
l
) - klIi2IiTs(k2)1I2
< 26[TI(K - UJ]I/2
-
< 2./'26 (s E E) J
whence
214 8. Ideal Theory
Consequently we see at once that
IIf * kill 2/26 + 2/
26
11
f
ll
l
= 2/26(1 + IIflll)
< I,
which completes the proof.
We are now in a position to prove the main theorem of this
section.
o
Theorem 8.6.1. Let G be a locally compact Abelian topological
group. Then Ll(G) satisfies Ditkin's condition at each point of

G and at infinity when G is noncompact.
Proof. The assertion that Ll(G) satisfies Ditkin's condition

at infinity when G is noncompact is precisely the content of

Corollary 8.1.3. So suppose wE G and f E Ll(G) is such that

few) = o. If n is a given positive integer. then again appealing
to Corollary 8.1.3 we see that there exist some v E Ll(G) such
n
that v E C (G) and IIf - f * v III < 1/2n. Clearly f * v Ll (G)
n c. n n
and (f * v ) (w) = O. Next, applying Lemma 8.6.2 to f * v, we
n n
deduce the existence of some k E LI(G) such that k (y) = 1.
n n
y E V. where V is a suitable open neighborhood of w, and such
n n
that Ilf * v * k III < 1/2n. Set g = v - v * k. Then g E Ll (G)
n. n n n. n n n
and gn(Y) = vn(Y) - vn{y)kn(Y) = vn(y) - vn{Y) = 0, Y E Vn" More-
over,
IIf - f .. g III = IIf - f * (v - v * k Jill
n n n n
1 + 1
< 2n 2n
1
= -.
n
8.7. Further Remarks on Ideals
215
Consequently limnllf - f * gnU} = 0, and Ll(G) satisfies
..
Ditkin's condition at wand hence at each point of G.
C!
Thus Ditkin's Theorem (Theorem 8.5.1) and Corollaries 8.S.l and
8.5.2 are valid for Ll (G). These give us the following result:
Corollary 8.6.1. Let G be a locally compact Abelian topolo-
gical group. Then

(i) If E eGis a closed set such that bdy(E) contains no
nonempty perfect set, then E is a set of spectral synthesis for
Ll (G) .
(ii) If I C Ll(G) is a closed ideal such that bdy[h(I))
contains no nonempty perfect set, then I is the intersection of
the maximal regular ideals containing Ij that is, I = k[h(I)).
As some concrete applications of this corollary we observe that,

if E eGis closed and bdy(E) is discrete, then E is a set of
spectral synthesis for Ll(G). Thus, for example, if G = Rand
E = [-1,1], we see that E is a set of spectral synthesis for
LIOR) , as bdy(E) = (-1,1). This partially proves the remark made
concerning L.Schwartz's example discussed after Malliavin's Theorem
(Theorem 8.3.4).
The description of the sets of spectral synthesis for Ll(G),
G being a compact Abelian topological group, given by Theorem 8.3.3
and Corollary 8.3.2,is also an immediate consequence of the preced-
ing remark. Indeed, in this case we have by Corollary 4.7.5 that

G is discrete, and so bdy(E) is empty for every set E C G.
8.7. Some Further Remarks on Ideals. In the preceding sections
of this chapter we have developed a modest portion of the ideal
theory for commutative Banach algebras and examined some of the ab-
stract results in specific Banach algebras. As should be apparent,
the study of ideal structure is a very complicated affair, even for
216
8. Ideal Theory
specific algebras, and we shall not pursue it further here. However,
before leaving the subject we would like to indicate some standard
questions concerning ideals, two of which were mentioned in the in-
troduction to this chapter, and tabulate the results for certain
algebras. The proofs of a portion of these results appear in the
preceding pages.
Definition 8.7.1. Let A be a commutative Banach algebra. A
closed ideal I in A is said to be primary if it is contained in
precisely one maximal regular ideal in A.
Evidently every maximal regular ideal is primary, but the con-
verse may fail. For consider the losed ideal I in
CnCrO,I]), n 1, defined by
I = {f I f (Cn([O,I]), f(O) = f'CO) = 0).
Then the only maximal ideal containing I is
M = {f I f (Cn(rO,I]), fCO) = 0),
while clearly I M.
Obviously a primary ideal is always proper.
Some general questions about ideals in a commutative Banach
algebra A are the following:
1. Is every proper closed ideal I in A contained in some
maximal regular ideal; that is, is hCI)
2. Is every proper closed ideal I in A the intersection of
the maximal regular ideals containing I; that is, is I = k[hCI)]?
3. Is every primary ideal I in A a maximal regular ideal?
4. Is every proper closed ideal I in A the intersection
of the primary ideals containing I?
8.7. Further Remarks on Ideals
217
The answers to these questions for some particular Banach alge-
bras are contained in the table below.
*
** Algebra Ll(G)
C(X)
*
*
G a noncompact X a compact
*
*
Cn([O, 1])
*
locally compact A(D) Hausdorff
*
*
*
Abelian
n > I topological space
*
-
Question
topological group
I Yes Yes Yes Yes
2 No No No Yes
3 Yes No No Yes
4 No Yes No Yes
More generally the answer to question I is affirmative either
when A has an identity or A is a Tauberian semisimple regular
commutative Banach algebra. However, if A is a radical algebra,
then the answer to question 1 is negative.
For further material on ideal structure in addition to that
cited in this chapter the reader is referred to [N,Ri).
CHAPTER 9
BOUNDARIES
9.0. Introduction. Recall that A(D) is the commutative
Banach algebra with identity of all those complex-valued functions
that are defined and continuous on the closed unit disk D and
analytic on int(D). The Maximum Modulus Theorem of analytic func-
tion theory asserts that, if f E A(D) , then II f llCl) = sup, . rl f (C) (.
where r = (e Ie. D, lei = 1), and there exists some , int(D)
such that IIfllCD = If(C)1 only if f is a constant function. A
second central result of analytic function theory, the Cauchy Inte-
gral Formula, tells us that, if f A(D), then
f(C) = S fez) dz
2nl r z - C
(C E int (D) ) .
In this chapter we shall examine the question of obtaining generali-
zations or analogs of these classical results in the context of commu-
tative Banach algebras.
Actually, it is advantageous to shift the investigation away
from commutative Banach algebras or, more precisely, away from the
Gel'fand representations of such algebras to a more general class of
algebras, which we shall call separating function algebras. In the
following sections we shall develop a portion of what can be said
about the Maximum Modulus Theorem and the Cauchy Integral Formula
for such algebras. The development falls roughly into two parts.
In Sections 9.1 through 9.4 we shall concern ourselves primarily with
analogs of the Maximum Modulus Theorem. This leads to the introduc-
tion of the concept of a boundary for a separating function algebra,
which will be seen to be a suitable replacement for r in the study
of A(D). We shall investigate three specific boundaries: the ~ i l o v
218
9.1. Boundaries 219
Bishop, and Choquet boundaries. In Sections 9.6 through 9.8 our
primary, but not our only, concern will be counterparts of the Cauchy
Integral Formula. This development leads to the notion of represent-
ing measure. From the classical situation, as exemplified in A(D),
it should come as no surprise that the notions of boundaries and re-
presenting measures are intimately connected. This connection will
be perhaps most apparent in Sections 9.6 and 9.7. Between the two
portions of the chapter just indicated is a section devoted to sever-
al applications of the material on boundaries developed up to that
point. However, the range of applications of the material introduced
in the chapter is substantial, particularly in the study of function
algebras, and we have made no attempt to provide a general introduc-
tion to these applications. The reader who is interested in such
applications should consult the appropriate references cited in this
chapter.
9.1. Boundaries. Consider the commutative Banach algebra with
identity A(O) consisting of all those functions defined on the
closed unit disk 0 in C that are continuous on 0 and analytic
on int(O). Then the classical theory of analytic functions contains
a number of results concerning the maximum modulus of elements of
A(O). To be precise, the following results are valid. Recall that
r = (C ICE D, Ici = 1).
1. If f E A(D), then there exists some C E r such that
I fCC) I = II f Il
CD
In particular, we see that IIfU. = suP
C
E rl fCC) I
and that If I assumes its maximum on f.
2. If E c r is closed and if for each f E A(D) there exists
some C E E such that If(C)1 = IIfll, then E = r. In particular,
CD
r is the smallest closed subset of D on which If I assumes its
maximum for each f E A(D).
3. If f E A(D) and there exists some C E int(D) such that
If(C)1 = UfU, then f is a constant.
CD
220 9. Boundaries
4. If f (. A(D) is nonconstant and e E int(D) , then for
each p > 0 such that
(z I Iz - el < p] C int(D) there exists
some w,
Iw - el < p,
such that
If(w)1 > If(e)l
In other
if f ( A(D) is nonconstant, then
I f I
cannot have a local maximum
at any point in int(D).
The first result or some variation of it is usually referred to
as the Maximum Modulus Theorem or the Maximum Principle (see, for
example, [A, pp. 133-135; Ru
2
, p. 213]). The fourth result is called
the Local Maximum Modulus Theorem. Our main concern in this and the
following four sections will be to examine some extensions of these
results for A(D) to general commutative Banach algebras. In parti-
cular we shall see that there always exist valid analogs of
1 and 2, but result 3 may very well fail to hold. A valid analog
of result 4 also exists, but a proof would involve considel'nble
machinery which we have not developed. The curious reader is referred
to [Ga, pp. 91-93; S, pp. 89-98; Wm
2
' pp. 52-55] for these generali-
zations of the Local Maximum Modulus Theorem.
There is some advantage in the sequel in working with algebras
of continuous functions rather than an arbitrary commutative Banach
algebra. Moreover, the Gel'fand Representation Theorem (Theorem
3.3.1) asserts that every commutative Banach algebra A is continu-
A
ously homomorphic to a subalgebra A of C (A(A) that separates
o
the points of A(A). This theorem will then allow us to translate
results about such subalgebras of C (X), X being a locally compact
o
Hausdorff topological space, into results about commuative Banach
algebras. With this in mind we make the following definition:
Definition 9.1.1. Let X be a locally compact Hausdorff topo-
logical space and suppose A is a subalgebra of C (X). Then A
o
is said to be a separating function algebra on X if
(i) A separates the points of X.
(ii) ZeAl = {t I t E X, f(t) = 0, f E A] = ,.
9.1. Boundaries 221
Clearly, if X is compact and Ac C(X) is a subalgebra that
separates the points of X and contains the constant functions,
then A is a separating function algebra. Furthermore, if A is

a commutative Banach algebra, then A = Ace (a(A)) is a separating
o
function algebra.
A useful observation about such algebras is contained in the
next proposition. The proof is left to the reader.
Proposition 9.1.1. Let X be a locally compact Hausdorff topo-
logical space. If
the topology on X
A is a separating function algebra on X, then
is the weakest topology on X such that all the
functions in A are continuous.
We conclude this section with several additional definitions.
Definition 9.1.2. Let X be a locally compact Hausdorff topo-
logical space and let A be a separating function algebra on X.
A set E c X is said to be a boundary for A if for each f E A
there exists some tEE such that If(t)1 = IIfllCD = sUPsExlf(s)l.
Thus a boundary for a separating function algebra A is a sub-
set of X on which If I assumes its maximum for each f E A. For
obvious reasons boundaries are often called maximum modulus sets.
If A = A(D), then the Maximum Modulus Theorem for analytic
functions says precisely that r is a boundary for A(D) and that
r is even the smallest closed subset of D that is a boundary for
A(D). We shall see shortly that every separating function algebra
has a smallest closed boundary. On the other hand, a smallest boun-
dary need not exist.
Definition 9.1.3. Let X be a locally compact Hausdorff topo-
logical space and let A be a separating function algebra on X.
A point t E X is said to be a peak point for A if there exists
some f E A such that If(t)1 = IIfll = 1 and If(s)1 < 1, s t.
CD
The set of all peak points for A is called the Bishop boundary for
A and is denoted by pA.
222 9. Boundaries
It is apparent that, if t X is a peak point for A, then
t belongs to every boundary of A, and so pA is the minimal boun-
dary for A. However, there exist separating function algebras A
that have no peak points -- that is, such that pA We shall
see an example in Section 9.3.
Peak points are also called strong boundary points and unique
maximum points (see, for instance, [HOI' p. 61; Ri, p. 141]). The
term "strong boundary point", however, is usually applied to points
that we shall call weak peak points. The reader is warned that there
seems to be no fixed terminology for these concepts and is advised
to check the definitions in the individual references cited.
9.2. The ilov Boundary. The main theorem of this section is
that every separating function algebra possesses a unique minimal
closed boundary, that is, a closed boundary that is contained in
every closed boundary. We shall also show that this is true, in a
suitable any commutative Banach algebra and that, if such
an algebra A is either regular or self-adjoint, then the minimal
closed boundary is precisely A(A). This unique minimal closed
boundary will be called the ilov boundary.
I
Theorem 9.2.1 (ilov). Let X be a locally compact Hausdorff
topological space. If A is a separating function algebra on X,
then A has a unique nonempty minimal closed boundary.
Proof. Let E denote the family of all closed boundaries for
A. Clearly E as X E. We introduce a partial ordering in
E by setting EI > E
2
, E
I
,E
2
E E, if and only if El C E
2
. If
(E) is a linearly ordered subset of E, then obviously E = n E
a 0 a a
is a closed subset of X. We claim that E is also a boundary for
o
A.
Indeed, suppose f E A and f is not identically zero. Define
E
f
C X as E
f
= {t l t E X, If(t) 1 = IIflt'). We claim that E
f
is
a nonempty compact subset of X. Since A c C (X), there exists
o
9.2. The ilov Boundary
223
some compact set K c: X such that I f(t) 1 < IIfll /2. t E X - K. Thus
CD
f restricted to K is continuous, and so Ifl assumes its maximum
modulus on K. Hence E
f
is a nonempty closed subset of K, and
so E
f
is a nonempty compact subset of X. Moreover, since each
Ea is a closed boundary for A, we see that E
f
n Ea is a nonempty
compact subset of K for each a. Since the family {E) is linear-
tl
ly ordered, we see that the family {E
f
n EaJ has the finite-inter-
section property, whence we conclude that na(Ef n Ea1 = E
f
n Eo is
nonempty. In particular, this also shows that E itself is non-
o
empty and that, if t E E
f
n E c: E, then If(t)l = IIfll. Evident-
o 0 CD
ly, if f E A is identically zero. then there exists some tEE
o
such that If(t)l =Ufll. Hence we see that E = n E is a closed
CD 0 aa
boundary for A.
Thus every linearly ordered subset of E has an upper bound,
and so, appealing to Zorn's Lemma [D5
l
p. 6], we deduce the exis-
tence of a maximal element in E. call it E.
Clearly E is a minimal closed boundary for A, and to show
that it is unique it suffices to prove that, if F is any closed
boundary for A, then E c: F. To see suppose there exists
tEE - F. Now, by Proposition 19.1.1, the topology on X is the
weak topology by the elements of Aj that is, a neighbor-
hood base for the topology of X at the point t consists of sets
of the form
where s > 0 and f
l
,f
2
, .. ,f
n
in A are arbitrary. In particular,
since F is closed, there exist some s, 0 < s < 1, and
f
l
,f
2
, . ,f
n
in A such that t E U(tjeifl'f2 . ,fn) = U and
U n F = ,. Moreover, we may assume, without loss of generality. that
IfkCs) - fk(t)1 < 1, k = 1,2, . n; sEX, on the following grounds:
If sUPk=1,2, . ,n[suPsEXlfk(s) - fk(t)il = M> 1, then set
gk = fk/M, k = 1,2, . n. Clearly gk A, Igk(S) - gk(t)1 < 1,
k = 1,2, . ,n; 5 X. and U(t;c/M;gl,g2, . ,gn) c: U(t;c;f
l
,f
2
. ,f
n
).
224
9. Boundaries
Moreover, since E is a minimal closed boundary, there evident-
ly exists some f E A such that IIfli = 1 and If(s)1 < 1,
CD
sEE - U. Note that, if E - U is empty, then the second assertion
is trivially satisfied. Let g E A be a sufficiently high power of
f such that Ig(s) I < I, sEE - U. Clearly one still has IIgli = 1.
CD
Then we see that, for each k = 1,2, .. ,n,
<e
(s E U)
and
< I (s E E - U),
whence Ig(s)fk(S) - g(s)fk(t)1 < I, sEE. But E is a boundary
for A, and so we have 19(s)fk(s) - g(s)fk(t)1 < I, k = 1,2, ... ,n
and sEX. However, since F is a boundary for A, there exists
some w E F such that 19(w) 1 = IIgli = 1. Consequently
CD
{
< I (k = l,2, . ,n),
that is, w C U, contradicting the fact that U n F = ,.
Therefore E C F, and E is the unique minimal closed boun-
dary for A.
o
A proof of this theorem that does not depend on Zorn's Lemma
is also available (see, for examvle, [5, pp. 37-39]). This theorem
provides another illustration of the advantages in requiring the
scalar field to be C, since if one considers separating function
algebras over the real numbers, then a unique minimal closed boundary
need not exist. For further details see [Ri; pp. 134 and 135, 311
and 312].
9.2. The ~ l o v Boundary
225
In view of the existence of a unique minimal closed boundary
for a separating function algebra we make the following definition:
Definition 9.2.1. Let X be a locally compact Hausdorff topo-
logical space and let A be a separating function algebra on X.
The unique minimal closed boundary for A is called the ~ l o v boun-
dary and will be denoted by cA.
It is apparent that pAc cA.
Two corollaries of the proof of Theorem 9.2.1 are worth mention-
ing. The details are left to the reader.
Corollary 9.2.1. Let X be a locally compact Hausdorff topo-
logical space. If A is a separating function algebra on X, then
cA is the intersection of all the closed boundaries for A.
Corollary 9.2.2. Let X be a locally compact Hausdorff topo-
logical space and let A be a separating function algebra on X.
If t E X, then the following are equivalent:
(i) t E cA.
(ii) If U is an open neighborhood of t in X, then there
exists some f E A such that IIfll = I and If(s)1 < 1, sEX - U.
CD
(iii) If U is an open neighborhood of t in X and
o < i < 1, then there exists some f E A such that
If(s)1 < i, sEX - U.
IIfll = I and
CI)
The next result provides us with a sufficient condition that
cA = X.
Theorem 9.2.2. Let X be a locally compact Hausdorff topolo-
gical space and let A be a separating function algebra on X. If
A is norm dense in C (X), then oA = x.
o
Proof. Let t E X and suppose U is an open neighborhood of t.
226 9. Boundaries
Since C (X) is a regular commutative Banach algebra, there exists
o
some g E C (X) such that get) = land g(s) = 0, sEX - U. Let
o
f E A be such that IIf - gil < 1/2. Then it is easily verified that
CD
If(s), < 1/2, sEX - U, and ,f(t), > 1/2, from which it follows,
by Corollary 9.2.2, that t E oA. Hence oA = X.
o
Corollary 9.2.3. Let X be a locally compact Hausdorff topo-
logical space and let A be a separating function algebra on X.
If A is closed under complex conjugation, then oA = X.
If A is a commutative Banach algebra, then, as already noted,
..
the Gel'fand representation A of A is a separating function alge-
bra on ~ A ) . This observation prompts the following definition:
Definition 9.2.2. Let A be a commutative Banach algebra. A
subset E of ~ A ) is said to be a boundary for A if E is a
..
boundary for
for A if T
A, and a point T ~ A(A) is said to be a peak point
..
is a peak point for A.
Theorem 9.2.1 then shows at once that each commutative Banach
algebra A possesses a unique minimal closed boundary, the ilov
.. \
boundary for A. This boundary will also be referred to as the
ilov boundary for A, and we shall denote it by oA, that is,
..
oA = oA. For commutative Banach algebras where the Gel'fand trans-
formation is the identity mapping, for instance,
..
C (X) or
o ..
A(D),
this is a natural notation since C (X) = c (X)
o 0
and A(D) = A(D).
Similarly the set of all the peak points for a commutative Banach
algebra A will be called the Bishop boundary for A, and we shall
..
denote it by pA = pA.
Corollaries 9.2.1,9.2.2, and 9.2.3 and Theorem 9.2.2, of course,
have valid analogs for any commutative Banach algebra. For example,
the first and second portions of the next theorem follow immediately
from the first part of the proof of Theorem 9.2.2 and from Corollary
9.2.3, respectively.
9.3. Examples of Boundaries
227
Theorem 9.2.3. Let A be a commutative Banach algebra. Then
aA = A(A) if either of the following two requirements is fulfilled:
(i) A is regular.
(ii) A is self-adjoint.

Recall that A being self-adjoint means precisely that A is
closed under complex conjugation (Definition 5.1.1).
9.3. Some Examples of Boundaries. In this section we wish to
describe aA and pA for some specific separating function algebras.
Example 9.3.1. Let X be a locally compact Hausdorff topolo-
gical space. Then C (X) is a semisimple regular commutative
o
Banach algebra, and so from Theorem 9.2.3 we see that aCo(X) =
A(Co(X = X. If X is oompact, then we also have pCo(X) = pC(X)
= X, provided that X is metrizable. Indeed, if X is compact
and metrizable, then for each t E X there exists a sequence of
open neighborhoods Uk of t such that U
k
+
l
C Uk and n;=lU
k
= {t).
Since X is a compact Hausdorff topological space, it is a normal
.
topological space [W
2
' p. 83], and so there exist open neighborhoods
W
k
of t in X such that cl(W
k
) C Uk' k = 1,2,3,... [W
2
, p. 49].
Then, by Urysohn's Lemma [W
2
, p. 55], there exist functions fk ( C(X),
k = 1,2,3, , such that
(i)
fk(s) = 1, s E cl(W
k
).
(ii)
fk(s) = 0,
s ( X - Uk.
(iii)
o fk(s) 1, sEX.
Evidently f = tk=lf
k
/2
k
belongs to C(X) and f(t) = 1.
Moreover, if sEX and s t, then there exists some positive
integer no such that s Uk' k > no' Consequently fk(s) = 0,
k>n,
o
228 9. Boundaries
no k no k
f(s) = f
k
(s)/2 < 1/2 = 1
k=l k=l
Therefore t is a peak point for C(X), and so pC(X) = X.
We also see at once from Theorem 9.2.3 that 3A = when
n
A is either C ([a,b]), n 0; L
1
(G), G being a locally compact
Abelian topological group; or AC(f), since all these algebras are
regular.
Example 9.3.2. The classical theory of analytic functions, as
indicated in Section 9.1, shows at once that oA(O) = f. Moreover,
in this case we once again have pA(O) = f on the following grounds:
If C r, then the function fez) = (C + z)/2, z (0, obviously
belongs to A(O), and If I is easily seen to have a unique maximum
at z = C. Thus each point of r is a peak point for A(O), and
since pA(O) C oA(O), we conclude that pA(O) = r.
Example 9.3.3. Next consider the set O
2
C e
2
, defined by
02 = ((z,w) 1 Z,w c, Izl 1, lwl 1), and let P(02) denote
the uniform closure on 02 of the algebra of all polynomials in
two complex variables -- that is, polynomials of the form
n " k
a"k
zJw
(aJ"k C; n = 0,1,2, . ).
j , k=O J
Then it is readily seen that P(02) is a commutative Banach algebra
with identity where the algebra operations are the usual pointwise
ones and the norm is the supremum norm over O
2
. Moreover, it is
evident that P(02) is a finitely generated Banach algebra generated
by the polynomials Pl(z,w) = z and P2(z,w) = w, (z,w) E 02' Thus
from Theorem 4.5.1 we see that can be identified with 02'
and the Gel'fand transformation on P(02) is the identity mapping.
We claim that oP(02) = PP(02) = ((z,w) 1 (z,w) E e
2
, Izl = lwl = I).
Indeed, suppose f E P(02) and suppose (qkJ is a sequence of
polynomials such that limkllf - qkilCl) = o. Then we see that, for
each w E 0,
9.3. Examples of Boundaries
229
f (z) = f(z,w) = lim qk(z,w)
w k
(z ( 0)
and that the convergence is uniform. Since the polynomials qkC,w)
are analytic for each wED, we conclude that f E A(D). Conse-
w
quently from Example 9.3.2 we deduce that If I attains its maximum
w
on (z I z (C, Izi = 1] for each w E o. Similarly, if f (w) =
z
f(z,w), w E 0, then f E A(O) and If I attains its maximum on
z z
(w I w E C, Iwl = 1]. Combining these two observations, we conclude
that, if f E P(02)' then If I attains its maximum on
({z,w) I (z,w) E Izl = Iwl = I};
that is, ap(D
2
) c (z,w) I (z,w) E Izi = lwl = I).
However, if (e,w) E and lei = Iwl = 1, then it is easily
seen that f(z,w) = (C + z)(w + w)/4, (z,w) E D
2
, is an element of
P(D
2
) and that If I has a unique maximum at the point (C,w). Thus
every such point in D2 is a peak point for P(D
2
), and so ap(D
2
)
= PP(02) = ((z,w) I (z,w) (C
2
, Izi = Iwl = IJ.
There are two observations worth making about this example.
First, we see that apeD?) is a proper subset of the topological

boundary of = D2 since
bdy(D
2
) = {{z,w) I (z,w) E e
2
, Izi = 1 or lwl = I}.
Second, there exist nonconstant functions f in P(D
2
) such that
If I assumes its maximum off the boundary. For instance, the
generators for P{D
2
) are examples of two such functions. Thus
the classic analytic function theorem which asserts that, if f E A(D)
and If I attains its maximum on int(D) , then f is a constant
has no valid analog for arbitrary commutative Banach algebTas or
separating function algebras.
Example 9.3.4. For our next example we set
x = ((z,t) I zED, t E [0,1])
230 9. Boundaries
with the topology induced on X as a subset of ~ and let A be
the subalgebra of C(X) consisting of all the f (C(X) such that
f(,O) E A(O) and f(O,.) E C([O,I]). Clearly X is a compact
Hausdorff topological space and A is a separating function algebra
on X that contains the constants. We claim that
3A = ((z,O) I Izl = 1) U ((O,t) I t ( [0,1]),
whereas pA = 3A - ((0,0)).
Indeed, suppose for each g E A(O) we define f (A by
g
fg(z,t) = g(z) for (z,t) E X; t = 0,
f (z,t) = g(O) for (z,t) E X; z = 0; 0 < t < 1,
g
and for each h E C([O,I]) we define fh E A by
fh(z,t) = h(O) for (z,t) E X; t = 0,
fh(z,t) = h(t) for (z,t) ( X; z = 0; < t < 1.
Using the facts that oA(O) = pA(D) = f, oC([O,I]) = pC([O,I]) = [0,1]
and that oA is closed, we see easily that
oA = ((z,O) I Izl = 1) U ((O,t) I t E [0,1])
and that every point in oA - (CO,O)) is a peak point for A. How-
ever, (0,0) is not a peak point for A. If it were, then there
would exist some f E A such that If(O,O)1 = IIfll = 1 and
CD
If(z,t)1 < 1, (z,t) EX, (z,t) ; (0,0). But then g(z) = f(z,O),
z E 0, would be an element of A(O) for which Ig(O)1 = 1 and
Ig(z)! < 1, z E 0, Z ; 0, contradicting the fact that pA(O) = f.
Consequently, pA = oA - ((0,0)).
Example 9.3.5. Finally we wish to give an example of a separa-
ting function algebra that has no peak points -- equivalently, such
that the Bishop boundary is empty. Let A be an uncountable set
9.3. Examples of Boundaries
231
and for each a A let I = [0,1]. Then the topological product
a
space X = fiaE:AIa is a compact Hausdorff topological space and
A = C(X) is a separating function algebra on X. As seen in the
first example of this section, we have oC(X) = X. But we claim
that pC(X) To see this it suffices to show that there exist
two disjoint subsets EO and El of X that are boundaries for
C(X). We shall show that the sets EO and El in X that consist,
respectively, of all those points t = (t ) E X such that t = 0
a a
or t = 1 for all but countably many a E A have the indicated
a
property. Evidently EO n El so we need only show that EO
and El are boundaries for C(X).
To this end, for each a E: A we denote by p the projection
a
of X onto I; that is, p (t) = t , t E X. Clearly each p E C(X),
a a a a
and we let W denote the subalgebra of C(X) generated by the pro-
jections p, a E A, and the constants. That is, W is the subal-
a
gebra of C(X) consisting of all polynomials in the projections
p , a E A. It is easily verified that W is a subalgebra of C(X)
a
that separates the points of X, contains the constant functions,
and is closed under complex conjugation. Hence, by the Stone-Weier-
strass Theorem [L, p. 333], we conclude that W is norm dense in
C(X). Consequently, if f E C(X), then there exists a sequence
(qk) C W such that li,\lIf - qkllao = O. However, each qk is a
polynomial in the projections p, a E A, and so a moment's reflec-
a
tion reveals that qk is completely determined by the Pals for
only a finite number of a E A. That is, for each k there exists
a finite subset of A such that, if s,t E X and p (s) = p (t),
-l( a a
a then qk(s) = qk(t). Let Ao = Then Ao is a
countable subset of A and, if s,t E X and p (s) = p (t), a E A ,
a a 0
then qk(s) = qk(t), k 1,2,3, ... Furthermore, it follows immedi-
ately from this that, if s,t E: X and p (s) = p (t), a E A, then
a a 0
f(s) = f(t), since - qkllao = o.
But now suppose t = (t ) E X is such that If(t)l = IIfli
g
Such a point exists since 3C(X) = X. Then define s = (s) and
a
232
w = {w J in X by
a
and
9. Boundaries
s = t for a ( A ,
ex ex 0
s = 0 for a E A - A ,
a 0
w = t for a E A ,
a a 0
w = I for a E A - A.
a 0
Obviously s E EO and w (E
I
, and p (s) = p (w) = p (t) = t ,
a a a a
a (A, whence If(s)1 = If(w)1 = IIfli. Thus, since f E C(X) is
o e
arbitrary, we see that EO and EI are both boundaries fOT C(X),
and so pC(X) =
9.4. Extreme Points, the ~ i l o v Boundary, and the Choquet Boun-
dary. If A is a separating function algebra on a locally compact
Hausdorff topological space X, then the development of the preced-
ing sections has shown that A always possesses a minimal closed
boundary, but it mayor may not have a minimal boundary. The main
purpose of this section is to introduce a new boundary for separa-
ting function algebras on compact Hausdorff topological spaces that
contain the constants. This boundary. which we shall call the Cho-
quet boundary, always exists and lies between the Bishop and the
~ i l o v boundaries. Moreover, the ~ i l o v boundary will be seen to be
the closure of the Choquet boundary.
The existence of the Choquet boundary will be established by
the first theorem of this section. Since the proof of this theorem
depends on applications of the Krein-Mil 'man Theorem, we wish to
recall to the reader some relevant terminology and notation. More
detailed discussion can be found in [L, pp. 317-326]. Also, before
proving the indicated theorem, we shall establish two lemmas that
will be useful in the course of the proof.
9.4. Extreme Points
233
In general, if V is a normed linear space, we denote by V*
the Banach space of all the continuous linear functionals on V.
The weak* topology on V* will be dentoed by rw*, and V* with
this topology by the pair (V*,rw*). If E C V*, then co(E) de-
notes the convex hull of E; that is, co(E) is all the convex
linear combinations of the elements of E. The closure of co(E)
in (v*,rw*) will be denoted by co(E), and will be called the
closed convex hull of E. If E C V* is convex, then a point
x* E V* is said to be an extreme point of E if, whenever xi,xi ( E
and 0 < a < 1 are such that x* = axi + (1 - a)x
2
, we have
x* = xi = xi The set of extreme points of E will be denoted by
ext(E). The space (v*,rw*) is a locally convex topological linear
space and the Krein-Mil'man Theorem [L, p. 322] implies that, if
E C V* is nonempty, convex, and weak* compact, then ext (E) ; ,
and E = co[ext(E)]. Our applications of the Krein-Mil 'man Theorem
in the theorem below will be to subsets of the spaces V* = C(X)*
and V* = A*, where X is a compact Hausdorff topological space
and A is a separating function algebra.
The first of the two lemmas is the following result. We recall
that, if X is a locally compact Hausdorff topological space, then
M(X) is the Banach space of all bounded regular complex-valued
Borel measures on X. If ~ E M(X) is positive, we shall write
~ ~ o.
Lemma 9.4.1. Let X be a compact Hausdorff topological space.
If ~ (M(X) is such that
then ~ ~ o.
Proof. Since ~ is a regular measure, it clearly suffices to
show that
Ix f(5) ~ s ) > 0
234 9. Boundaries
for each f E C(X) such that 0 f(s) < 1, s X. Let f be such
a function and let
Ix f{s) = a + ib.
where a. b E JR. For each E lR define E C (X) by (s) =
f(s) + Then
whence we have
= If{s)12 +
< 1 +
II
x
< [Ix
(1 +
= 1 +
But we also see that
II
x
= II
x
f{s) +
= la + ib(l +
= a
2
+ b2(1 +
(s X).
22222 .
Thus we have a + b (1 < I + b E JR. from WhICh we de-
duce via some elementary calculations that I - (a
2
+ b
2
).
E lR. However, such an inequal i ty can be val id for all E IR
only if b = 0, and so we conclude that
Ix f(s) = a.
Furthermore, since 111 - fll < 1 and
CD -
we see that
9.4. Extreme Points
11 - al = II
x
- Ix f(s)
III - fllCD Ix (s)
< 1,
whence we conclude that a > O. Therefore > O.
Before we give the second lemma we need to introduce another
definition.
235
o
Definition 9.4.1. Let X be a locally compact Hausdorff topo-
logical space and suppose E M(X). Then is said to vanish on
the open set U c X if
Ix f(s) = 0
for every f E C (X) that has compact support contained in U. The
o
support of is the complement of the largest open set on which
vanishes.
The existence of the support of E M{X) can be established
by a routine application of Zorn's Lemma. If 0, then it is
evident that the support of is a nonempty closed subset of X.
Lemma 9.4.2. Let X be a compact Hausdorff topological space
and let E M(X) be such that
= Ix = 1.
If f E C(X) has the property that
Ix f(s) = IE f(s)
for each Borel set E C X such that 0 < < 1, then f is
equal to a constant almost everywhere with respect to
Proof. From Lemma 9.4.1 it is apparent that > O. Let t E X
236 9. Boundaries
be a point of the support of and let e > O. From the continuity
of the regularity of and the definition of the support of
we See that there exists an open neighborhood U of t such that
o < < I and such that If(t) - f(s)l s U. Thus
If(t) - Ix res) = f(t) - f(s)
.. Iu If(t) .. f(s) I d .. (s)
< s.
Since > 0 is arbitrary, we conclude that
f(t) = J X f(s) dlJo (s)
for each t in the support of Hence f
everywhere with respect to
is a constant almost
o
Now suppose that X is a compact Hausdorff topological space
and that A is a separating fUnction algebra on X that contains
the constants. In this case we shall generally denote the function
in A that is identically one on X by the 1. The context
will make clear whether we mean this function or the number "1". If
t 'X and we set T(t)(f) = TtC) f(t). f 'At then it is appar-
ent that T(t) = 'r
t
E A'" and that Ih't":: 1't (1) 'CO 1. It is easily
seen that the mapping X - A* defined above identifies X with
a subset TeX) of (x* 1 x* E A*, IIx*1I = x*(l) = 1]. Moreover,
we shall see that as this latter set is a nonempty weak* compact con-
vex subset of A*. it has extreme points by the Krein-Millman Theo-
rem, and these points all belong to T(X). The set of t E X
such that "t is an extreme point of (x" I x* E A*, Ux*1J = x* (1) = 1)
is a boundary for A whose closure is aA. This is the main con-
clusion of the following theorem:
Theorem 9.4.1 (Bishop-deLeeuw). Let X be a compact Hausdorff
topological space and let A be a separating function algebra on X
that contains the constants.
9.4. Extreme Points
(i) If T : X A* is defined by T{t)(f) = Tt(f) = f(t),
t (X and f' A, then T is a homeomorphism of X onto T(X)
when the latter is considered as a subset of (A*,rw*).
(ii) CO[T(X)] = {x* 1 x* E. A*, IIx*1I = x*(l) = IJ whe-re the
closure is with respect to the weak* topology rw* on A*.
(iii) ext({x I x* E A*, lix*U = x*(l) = 1)) C T(X).
237
(iv) aA is equal to the closure in (A*,rw*) of ext(co[T(X)]);
equivalently, aA is the closure in X of
{t l t X. T
t
' ext(co[T(X)]l].
Proof. From the definition of the weak* topology on A* and
the fact that A separates the points of X it fOllows a ~ once
that T: X A* is injective and continuous. Since the weak* topo-
logy is Hausdorff and X is compact, we conclude that T is a homeo-
morphism of X onto T(X) C (A*,fW*). This proves part (i) of the
theorem.
To prove part (ii) we first note that IITtU = 'ftCl) = 1, t E X,
and so T(X) C (x
1t
' x* EA."', IJx*1I = x*(l) = 1] = BI. Moreover, it
is easily verified that B ~ is convex and weak* closed. Hence
CO[T(X)] C B ~ . However, if x* E A* and IIx*1J = x*(l) = 1, then
from a consequence of the Hahn-Banach Theorem [L, p. 87], we see
that there exists sOJle y* E CeX)* such that llx*U -= lIy*1I and
y* (f) == x* (f) J f E A.. Furtaermore, Ily*1I = IIx*U == x* (1) = y* (1) = 1.
Consequently, by the Riesz Representation Theorem for C{X)* [L,
p. 109], there exists a unique ~ E M(X) such that
y*(f) = Ix f(5) ~ s ) (f C(X))
and
Thus, by Lemma 9.4.1, ~ > O.
238
9. Boundaries
Now we recall that the extreme points of the closed unit ball
in M(X), that is, of the set (v I v E M(X), IIvll < 1), are pre-
cisely those measures of the form aCt' t (X, where a lal = 1,
and 6
t
E MeX) is the measure with unit mass concentrated at t;
that is, if E C X is a Borel set, then 6
t
(E) = 1 if tEE and
6
t
(E) = 0 if t E. A proof of this fact is available, for instance,
in [L, p. 338]. Since the closed unit ball in M(X) is weak* com-
pact and convex, by the Banach-Alaoglu Theorem [L, p. 254], we deduce,
from the Krein-Miltman Theorem [L, p. 322], that it is the weak*
closed convex hull of its extreme points. Combining these observa-
tions with the fact that > 0 and = 1, one easily verifies
that is a weak* limit of convex linear combinations of the ex-
treme points (6
t
I t E xl. However, since the restrictions to A
of the functionals determined by and 6
t
, t E X, are just x*
and T
t
, t E X, respectively, it follows immediately that x* is
a weak* limit of convex linear combinations of the Tt' t E X. That
- - 1
is, x* E COrTeX)], and so CO[T(X)] = B
1

I
Next we shall show that ext(B
I
) C T(X). We observe that, since
Bi is weak* closed and norm bounded, it must be weak* compact by
the Banach-Alaoglu Theorem, whence we conclude from the Krein-Mil'man
1 I
Theorem that ext(B
l
) If E ext (B
I
), then, by the argument
used in proving part (ii), we know that there exists a measure
E M(X), = 1, 0, such that
= Ix f(s) (f E A).
If E C X is any Borel set such that 0 < < 1, then define
xi,xi on A by
It is
it is
since
(f E A).
1
easily verified that xi and xi belong to B
1
. Moreover,
obvious that x* = (E)x
1
* + [1 - (E)]x
2
*. Consequently,
1 0 0 0
E ext(B
I
), we have = xi = xi
9.4. Extreme Points
239
Thus, in particular, we see that
(f Eo A)
for each Borel set E in X such that 0 < (E) < 1. By Lemma
o
9.4.2 we conclude that each f E A is constant almost everywhere
with respect to Since A separates the points of X, it fol-
o
lows at once that the support of must be a single point, and
o
so, since
Thus
> 0,
0-
there exists some t E X such that = 6
t
.
= Ix f(s) = f(t) = Tt(f)
whence we have = T
t
. Hence C T(X).
(f E A)'
Finally, suppose F A* - C is a linear functional that is
weak* continuous; that is, F is a continuous linear transformation
from (A*,rw*) to We claim that
Note that the supremum is finite, as F is continuous on the weak*
compact set Clearly the left-hand side of the equation is
always greater than or equal to the right-hand side. Suppose that
the inequality is strict, let p > 0 be such that
sup lIF(x*)1 > p > sup I IF(x*)I,
x* E 8
1
x* E ext (B
l
)
and let x* E Bli be such that IF(x*)1 > p. By the Krein-Mil'man
o _ 1 0 1
Theorem, we have that co[ext(B
I
)] = B
l
, and so there exist
E and a
k
a
k
> 0, k = 1,2,3, ... ,n, such that
= 1 and
n - p
IF(x*) - F( E a x*)1 < 2
o k=l k k
Hence, on the one hand,
240
9. Boundaries
>
on the other
n n
IF( E akx,) I < akIF(x,ll
k=l k=I
< p.
This contradiction shows that
for each weak* continuous linear functional F on A*.
it is apparent if F is a weak* continuous
linear functional on then we also have
sup IIF(x*)1 =
x* B
1
sup 1 IF(x*)I,
x* ( c 1 [ext (B
l
)]
where
1
denotes the weak* closure of ext(B
I
). Moreover,
since weak* continuous linear functiona1s separate the points of A*
[L, p. 240], it is easily verified that is the smallest
closed subset E C T(X) C Bl such that
1
sup llF(x*)1 = sup IF(x*)1
x* (B x* (E
1
for each weak* continuous linear functional F.
Now, however, each weak* continuous linear functional F on A*
is of the form F(x*) = x*(f), x* for some f E A [t, p. 238].
Thus, if f E A and F(x*) = x*(f), x* (A*, then
9.4. Extreme Points
IIfli :: sup If (t) I
CD t X
= sup ITt(f)1
T
t

:: sup IF(Tt)1
T
t
T (X)
:: sup 1 IF(Tt)l
T
t
E cl[ext(B
l
)]
:: sup 1 IF(Tt)l
T
t
E ext (B
1
)
= sup 1 If(t)l
T
t
ext (8
1
)
Consequently we see that, if f E A, then
IIfllCD :: sup -1 1 I f(t) 1
t E cl (T [ext (B
1
) ] )
:: _lsuP 1 If(t)l,
t ( T [ext (B
l
)]
241
-1 1
and that cl(T [ext(B
1
)]) is the smallest closed subset of X for
which the identity is valid. Therefore we see that
-1 I -1--
oA = Cl(T [ext(Sl)]):: cl(T
that is, oA is the closure in X of
(t I t E X, T
t
ext(co[T(X)])).
This completes the proof of the theorem.
The last portion of the proof reveals that
is a boundary for A that always exists. With this in mind we
make the following definition:
o
242 9. Boundaries
Definition 9.4.2. Let X be a compact Hausdorff topological
space and let A be a separating function algebra on X that con-
tains the constants. If for each t E X we set Tt(f) = f(t).
f E A. and = (x* I x* E A*. IIx*1I = x*(l) = 1), then
xA = (t I t E X. T
t
E is called the Choquet boundary for A.
The next corollary follows immediately from Theorem 9.4.1.
Corollary 9.4.1. Let X be a compact Hausdorff topological
space. If A is a separating function algebra on X that contains
the constants. then
(i) xA ,.
(ii) XA is a boundary for A.
(iii) cl(XA) = aA.
From our discussion of boundaries for specific algebras in the
preceding section we see at once that XC(X) = X, X being a compact
Hausdorff topological space, xA(D) = f. and xP(D
2
) = pP(D
2
) =
ap(D
2
). The Choquet boundary can, however, be a rather complicated
set. If X is metrizable, then xA is always a G
6
-set, but if
X is not metri:able, then may not even be a Borel set (see,
for instance, [5, pp. S4 and 55, 138 and 139]). In Sections 9.6
and 9.7 we shall discuss a number of equivalent descriptions of the
Choquet boundary in the case that A is norm closed.
In view of Definition 9.2.2 and Theorem 9.4.1. it is evidently

meaningful to consider the Choquet boundary XA = XA whenever A
is a commutative Banach algebra with identity. In particular, the
Choquet boundary for such an algebra always exists.
Before we contiftue an investigation of the Choquet boundary we
wish to look at several consequences of the existence of the ilov
and Choquet boundaries.
9.S. Applications of Boundaries
243
9.S. Some Applications of Boundaries. In this section we wish
to prove some results about separating function algebras and commu-
tative Banach algebras whose proofs utilize the notion of a boundary.
Theorem 9.S.l. Let X be a compact Hausdorff topological
space and let A be a separating function algebra on X that con-
tains the constants. If T : A A is a linear isometry of A onto
A such that T(l) = 1, then T is an algebra isometry of A onto
A.
Proof. Let T*: A* A* denote the adjoint of T, that is,
T*(x*)(f) = x*[T(f)], x* E A* and f E A. It is easily seen that
T* : A.* -- A.* is also a surjective linear isometry. Moreover,
since T(l) = 1 and T*[x*(l)] = x*[T(l)] = x*(l), x*
that T* maps = (x* I x* E A*, \lx*1I = x*(l) = lJ
From this observation it is readily deduced that T*
1
onto ext(B
1
).
A*, we see
onto itself.
1
maps ext (B
l
)
Now, if t E XA, then
s E XA such that T*(T) =
t
T
t
E and so there exists some
1
Ts since ext(B
l
) C T(X). Consequently
for each f,g E A we have
T(fg)(t) = Tt[T(fg)]
= T (fg)
s
= (fg) (5)
= f(s)g(s)
= Ts(f)TS(g)
= T*(T
t
) (f)T*(T
t
) (g)
= T(f) (t)T(g) (t).
244 9. Boundaries
Since t E XA is arbitrary, we conclude that for each f,g E A we
have T(fg)(t) = T(f)(t)T(g)(t), t E xA. whence T(fg)(t) =
T(f)(t)T{g){t), t E X, as XA is a boundary for A. Hence T is
an algebra isometry.
o
Next we wish to prove a result about topological zero divisors
in separating function algebras and in commutative Banach algebras.
Clearly a separating function algebra is a normed algebra, and so the
notion of topological zero divisors is meaningful in this context.
The results on topological zero divisors will be corollaries of the
following theorem:
Theorem 9.5.2. Let X be a locally compact Hausdorff topolo-
gical space and let A be a separating function algebra on X. If
f E A, then
inf If (s) I =
s E: oA
IIfgll
m
inf IIglim
gEA
g ~ 0
Proof. If g E A and g is not identically zero, we let
t E oA be such that 19(t)1 = IIglimo Then
inf If(s)1 < If(t)l
s E: oA -
Since this holds for each g E A, g ~ 0, we conclude that
inf If(s)l < a =
s E: oA
Ii
f
gli
m
inf I' II .
g E A Ig Q)
g ~
If a = 0, the proof is complete, so we may assume that a > O. Let
E = (s I sEX, If(s)1 > aJ. Evidently E is a compact subset of
X, as 0 < a < IIfll. If E were a boundary for A, then E would
- Q)
be a closed boundary for A, and so E ~ oA. From this it would
9.5. Applications of Boundaries
245
follow at once that
a inf ,f(s)I inf If(s),,
s E s E3A
which combined with the preceding estimate would show that
Ufgll
CD
inf If (s) I = inf II II .
s EoA g A g CD

Consequently it remains only to show that E is indeed a boundary
for A. However, if E is not a boundary, then there exists some
g E A, g 0" such that suPsEElg{s)l < Ilg11
CD
" as E is compact.
Moreover" it is easily seen for each positive integer k that
sup = sup[lg(s)l]k,
s ( E IIg II sEE
m
from ""hich it follo\\'s at once that limk[suPsEE(lgk(s)l/llgkUCD)] = O.
Furthermore" given k" there exists some tk E 3A such that
If(tk)gk(t
k
)I = IIfgkU
CD
" whence we deduce that
IIfgkUCD
a <_r-"-
- \lgkHCD
If{tk)gk(t
k
) 1
< k
Ig (t
k
) I
= If (t
k
) 1
Thus tk E E" k = 1,2,3, .. , and so
< IIfll sup
- CD 5 ( E IIg II
CD
(k = 1, 2 3, )
246 9. Boundaries
But these estimates entail that a = 0, contrary to assumption.
Therefore E is a boundary for A, and the proof is complete.
The next corollaries are easily established with the aid of
Theorem 1.6.2. The details are left to the reader.
o
Corollary 9.S.1. Let X be a locally compact Hausdorff topo-
logical space and let A be a separating function algebra on X.
If f E A, then the following are equivalent:
(i) f is a topological zero divisor.
Corollary 9.5.2. Let X be a locally compact Hausdorff topo-
logical space and let A be a separating function algebra on X
such that oA is compact. If f E A, then the following are equi-
valent:
(i) f is a topological zero divisor.
(ii) There exists some t E aA such that f(t) = o.
In particular, Corollary 9.5.2 is valid whenever X is compact.
It may not, however, be valid for noncompact X. For example, let
X = lR and A = C OR). Then aA = lR. Let f E C (lR) be such that
o -t2 0
f(t) 0, t E lEt For instance, let f(t) = e ,t E IR. If
gk E Co (lR) is the tent function defined for each positive integer
k by
gk(t) = t - k for k < t < k + 1,
gk(t) = -t + k + 2 for k + I < t < k + 2,
gk(t) = 0 for t [k,k + 2],
then it is easily seen that IIg
k
ll
m
= 1. k = 1.2.3, and
= o. Thus f is a topological zero divisor in A = Co (lR)
but f never vanishes on aA = lR.
9.5. Applications of Boundaries
247
In Section 1.6 we discussed some results about topological
zero divisors similar to the preceding theorem and corollaries. In
particular, the reader should compare the preceding results with
Theorems 1.6.1 and 1.6.2. Theorem 1.6.1 is helpful in proving the
next corollary.
Corollary 9.5.3. Let X be a compact Hausdorff topological
space and let A be a separating function algebra on X. If f E A,
define E C(oA) by = f(t), t (oA; that is,
is the restriction of f to oAt Then
(i) The mapping A - CCoA) is an isometric algebra iso-
morphism of A into C(oA).
(ii) If f (A, then is regular in CCoA) if and only
if f is not a topological zero divisor in A.
If we shift our attention to commutative Banach algebras, the
preceding results combined with Theorems 1.6.1 and 5.1.1 can be used
to establish the next corollary.
Corollary 9.5.4. Let A be a semisimple commutative Banach
algebra with identity such that there exists a K > 0 for which
IIxll
2
< Kllx
2
11, x E A. If x E A, then the following are equivalent:
(i) x is a topological zero divisor.

(ii) There exists some T E oA such that X(T) = O.
Our final application will be to obtain some results concerning
the extension of maximal ideals. In particular, we shall show that
every maximal ideal in the ilov boundary of a closed subalgebra A
of a commutative Banach algebra B with identity can be extended to
a maximal ideal in B. More precisely, we have the following theorem:
Theorem 9.5.3. Let B be a commutative Banach algebra with
identity e and let A be a closed subalgebra of B that contains e.
If M E oA c then there exists some N E such that N M.
248 9. Boundaries
Proof. Let T denote the complex homomorphism of A such that
-1
M = T (0) and suppose there exists no N E A(B) such that
N M = T-1(O). Then the ideal I in B generated by M must be
all of Bj that is, the set
m
I = ( E X
k
Y
k
1 X
k
M, Yk E B, k = 1,2, .. ,m),
k=l
where m is an arbitrary positive integer, must be all of B. Other-
wise, by Theorem 1.1.3, there would exist some N E A(B) for which
N I M, contrary to assumption. In particular, there exist
x
k
EM and Yk S, k = 1,2, .. ,n, such that e = tk=lx
k
y
k
, where,
without loss of generality, we may assume that
IICD = sup I (w) 1 < 1
wA(A)
(k = 1, 2, ... , n)
Furthermore, we observe that
sup = limllxnll
1/n
= sup (x ( A),
wEArS) n wA(A)
by the Beurling-Gel'fand Theorem (Corollary 3.4.1).
Now let a be so chosen that
a>
=

sup [sup lYk(w)11
k=l,2, ... ,n w(A(B)
and consider the open neighborhood U of T in A(A), defined by
U = (u,

Ixk(w) - Xk(T)1 < 1/2na, k = 1,2, ... ,n)

= (w I Ixk(w)l < l/2na, k = l,2, ... ,n).
Since M = T-l(O)
some yEA such
Clearly y = ye =
E oA, we see from Corollary 9.2.2 that there exists
that IIri! = 1 and I;(u.) I < 1/2na, w A(A) - U.
CI)
y(tk=lxkYk)' from which it follows that
9.6. Representing Measures 249
n
sup ly(w) xk(w)Yk(w)l =
k=l

= lIyll = 1.

However, we also see that
n n...
sup ly(w) xk(W)yk(w)l < sup
wEA(B) k=l k=l
< [sup sup lyk(w)ll
k=l wEA(B) wEA(B)
n
= [sup IY(w)xk(w)I sup lyk(w)ll
k=l w(A(B)

n
< a[ sup ly(w)xk(w)ll
- k=l wEA(A)
1
< 2'
a: < 1/2na, W E u, k = 1,2, ... ,n, and
ly(w)xk(w)l < IIxkll.1Y(w)1 < 1/2na, w A(A) - u, k = 1,2, ... ,n. We
have thus obtained a contradiction, and so there must be some
N E A(B) such that N M = T-I(O).
o
Theorem 9.2.3 immediately yields the next corollary.
Corollary 9.S.S. Let B be a commutative Banach algebra with
identity e and let A be a closed subalgebra of B that contains
e. If A is either regular or self-adjoint, then for each M E A(A)
there exists some N E A(B) such that N M.
It is not difficult to see that analogs of Theorem 9.5.3 and
Corollary 9.5.5 exist for a commutative Banach algebra A with iden-
tity and superalgebras B of A. We leave the formulations of these
results to the reader.
9.6. Representing Measures and the Choquet Boundary. There are
a number of equivalent descriptions of the Choquet boundary of a
250 9. Boundaries
separating function algebra, which we shall discuss in this and the
following section. One of these descriptions involves the notion of
a representing measure, and we now wish to give some insight into the
origin of this concept and at the same time indicate another funda-
mental connection between the theory of analytic and harmonic func-
tions and the study of function algebras.
With this is mind, suppose f E A(O) and CEO, lcl < 1. Then
the Cauchy Integral Formula [LeRd, p. 133] asserts that
Setting z
1 I fez)
f{C) = 2ui r z _ C dz.
it
= e ,-n t u, we see that this becomes
it it
f(C) = .!... In )e dt
2n -n It ,.
e -
it it
= Ir fee ) ),
it it it
where ) = 1[2n(e - C)])dt. Thus we see that for each
, E OJ lc, < 1, there exists some complex-valued measure in
M(r) such that f(C) is obtained by integrating f on r with
respect to Actually, for such C there even exist positive
measures on r of norm one which have the same property. Indeed,
if CEO, Ic' < 1, and z E r, then
z zr
-z = -(
= 1 - Cz
1
Moreover, it is apparent that h(z) = 1/(1 - Cz), z E 0, belongs
to A(O), and so from the preceding observations we see that
f(C) it it
= (fb) (e) = Ir (fh. ) (e ) )
1 - 1,1
it
= .!... In f {e " ( __ d
2
't "t) t
n -n 1 _ eel 1 _ Ce-
l
1 In
= 2n -n 11
it
fee dt.
_ Ce-
lt
l
2
9.6. Representing Measures
Consequently, we see that for each C E D, lcl < 1, we have
f(C) = 1 r1T f(e
it
) 1 - I Cl
2
dt
2n J_
n
11 _ Ce-
1t
1
2
I
it it
= r fee ) dve(e ),
where dVeceit) = [(1 - 1'12)/2ncll - Ce-
it
I
2
)]dt. Evidently ve
is a positive measure in M(r) and
it
IIvell = I r dV
C
(e ) = 1.
251
. it I 12 I -
it
l
2
it
The funct10n P,(e ) = (1 - e )/2n( 1 - Ce ), e E r, is,
of course. just the Poisson kernel [A, pp. 165-167].
Furthermore. we claim that ve is the only positive measure
in Mcr) such that
Cf E A(D)).
Indeed. suppose ~ e E M(r) is another such measure. Then for each
k = 0,1,2 we have
I
ikt it k k
r e d(V, - ~ C ) e ) = C -, =
and
= 0,
as the measure Vc - ~ , is real valued. Thus, if p is any trigo-
nometric polynomial on r, we see that
I
it it
r pee ) d(V, - ~ , ) e ) = 0,
whence it follows, by the Stone-Weierstrass Theorem [L, p. 332], that
(h e(r).
252 9. Boundaries
Hence we either from the regularity of v, - A, or the
fact that the dual space of C(f) is M(f) , that v, - A, = o.
The preceding development can be summarized by saying that for
each ,E D, Icl < 1, there exists a unique positive measure
\lC E: M(r) , IIvcll = 1, such that
f(C) = Ir f(e
it
) dV,(e
it
) (f E A(D)).
Moreover, this result also holds for ,E D such that 1,1 = 1.
If C E r, then it is apparent that, if Vc = 6
C
' the measure
with unit mass concentrated at C, then
it "t
fCC) = Ir fCe ) dV,(e
1
) Cf E A(D)),
and as before, 6, is the only such positive measure with norm one.
Indeed, suppose ,E r and define g(z) = Cl + ,z)/2, z D. It is
easily verified that g E: A(D), IIgll = g(C) = I, and Ig(z)1 < 1,
CID
z D, z C. Consequently it is evident that limkgkcc) = x{C)(z),
z D, where xC,} denotes the characteristic function of the
singleton set {C). Thus, if AC E M"(r) is any positive measure
such that IlA,lI = 1 and for which
(f A(D)) ,
then, by the Lebesgue Dominated Convergence Theorem [Ry, p. 229], we
deduce that
k "I k it it
1 = lim g CC) = 11m r g (e ) dACCe )
k k
it . t
= Ir x{C)(c ) dACce1 )
= A
C
({'))'
Thus we see that for each C D there exists a unique positive
9.6. Representing Measures
253
measure M(r) of norm one such that
it it
f(,) = Ir fee ) dv,(e )
(f A(D)).
Recalling the description of the maximal ideal space of A(D) given
in Section 4.4, we see that the points of D correspond precisely
to the complex homomorphisms of A(D). Thus, denoting by T, the
element of ACA(D)) = D corresponding to " we see that there
exists a unique positive Mcr) such that
(f E A(D))
and for which
Moreover, we observe that in this case r = 3A{D). Hence in a rather
obvious sense the measure represents the complex homomorphism
T" which prompts the following general definition:
Definition 9.6.1. Let X be a compact Hausdorff topological
space and let A be a separating function algebra on X that contains
the constants. If x* is a continuous linear functional on A,
then E M(X) is said to be a representing measure for x if
Ux*1I = and
x*(f) = Ix f(s) (f E A).
The reader should note carefully that, if x* E A* and E M(X)
are such that
x*(f) = Ix f(s)
(f E A),
then is a representing measure for x* only if IIx*1I = If
t E X and T
t
denotes the complex homomorphism of A defined by
Tt(f) = f(t), f A, then we shall frequently refer to a representing
measure for T
t
as a representing measure for t. Evidently = 6
t
,
the measure with unit mass concentrated at t, always is a represent-
254 9. Boundaries
ing measure for t X. However, it may not be the only such measure.
For example, if A = A(O) and , 0, "I < 1, then on setting
it I 12 1 -
it
l2
dv,(e ) = [(1 - , )/2n{ 1 - 'e )]dt we see that v, E N{r)
is a representing measure for , that is different from 6,. Repre-
senting measures for points are always positive. This is the content
of the next proposition.
Proposition 9.6.1. Let X be a eompact Hausdorff topological
space and let A be a separating function algebra on X that con-
tains the constants. If t E X and E MCX) is a representing
measure for t, then o.
Proof. The assertion follows at once from Lemma 9.4.1 on obser-
ving that
Referring once again to the case of A = A(O) , we see that each
point , 0 has a unique representing measure -- namely, v, --
such that the support of v, is a subset of the ilov boundary r
for ACD). Thus we are naturally led to the following questions
about separating function algebras A on compact Hausdorff topolo-
gical spaces that contain the constants: Which points in X have
unique representing measures? Does there exist a representing measure
for t E X such that E NcaA)? We shall see in Section 9.S
that the answer to the second question is always in the affirmative,
while the answer to the first question is that the points of X that
have unique representing measures are precisely the points of the
Choquet boundary for A.
However, before considering these questions in detail, we wish
to use the language of representing measures to prove an important
result about the algebras Ccr) and A(O). We noted in Section 4.4
during our discussion of the maximal ideal space of ACD) that the
algebra A(O) can be identified isometrically with a subalgebra
B(f) of C(I). The mapping that effects this identification is, of
9.6. Representing Measures
course, just the mapping that associates with
tion of f to f. Recall that 6(B(f)) = D.
that B(f) is a maximal subalgebra of C(f).
have the following definLtion and theorem:
255
f E A(D) the restric-
We now wish to prove
More precisely, we
Definition 9.6.2. Let X be a compact Hausdorff topological
space and let A be a separating function algebra on X that con-
tains the constants. If A is a closed subalgebra of C(X), then
A is said to be a uniform algebra.
It should be observed that a uniform algebra is a commutative
8anach algebra with identity.
Theorem 9.6.1 (Wermer's Maximality Theorem). Let A be a uni-
form algebra on f. If 8(I) c A, then either A = B(f) or
A = C(f).
Proof. Since 8(f) c A, it is evident that the function
h(e
it
) = e
it
, e
it
E f, belongs to A. Suppose that T(h) 0,
T E 6(A). Then, by Corollary 3.4.2, h is regular in A, from
- it -1 it -it
which we conclude at once that the function h(e ) = h (e ) = e
e
it
E f, belongs to A. However, an easy application of the Stone-
Weierstrass Theorem [L, p. 332] shows that the algebra generated by
hand h is dense in C(f), whence we deduce that A = Cef), as
A is a closed subalgebra of C(f).
On the other hand, suppose there exists some T E 6(A) such
that T(h) = o. Since 8(f) c A, it is apparent that T E 6(8(f))
= 6(A(D)) = D and. moreover, that T(f) = f(O), f E B(f). Note
that the latter equation is meaningful, as we may identify f E Bef)
with a unique element of A(D). The proof of the equation is as in
Section 4.4. Let E M(r) be a representing measure for T con-
sidered as an element of 6(A). From Proposition 9.6.1 we see that
> O. Moreover, we see that, for each f E B(f),
,
256 9. Boundaries
However, our development at the beginning of this section entails
that ~ = Vo since the representing measures in M(r) for the points
of 0 are unique. Thus we have
f(O)
't it
= T(f) = Ir f(e
1
) dvO(e )
(f A).
In particular, if f E A, then
o = hk(O)f(O) = T(hk)T(f)
= T(hkf)
= ~ n eiktf(eit) dt
2n -n
..
= f(-k) (k = I ~ 2 , 3, .. ) ;
that is, the negative Fourier coefficients of each f A are zero.
However, a classical theorem of Fourier series [E
2
, p. 87] asserts
that the Cesaro means of the Fourier series of a continuous function
f on r converge uniformly to f; that is,
o = limlla (f) - fll
n CD
n
= lim[ sup
n -n<t<n
Thus for each f E A we see that a (f) B(I), n = 0,1,2, ... , as
.. n
h E BCI) and f(-k) = 0, k = 1,2,3, ... , whence it follows that
f BCf), as B(r) is supremum norm closed.
and the theorem is proved.
Therefore A = B(f),
o
This proof of Wermer's Maximality Theorem is due to Hoffman
and Singer. A proof, due to P.J. Cohen, that does not depend on the
use of representing measures is also available. This proof as well
as some further discussion of maximality theorems and their applica-
tions can be found in [Ga, pp. 38-40; S, pp. 297-299, 302-303, 340].
9.6. Representing Measures 257
Finally, we shall establish the description of the Choquet boun-
dary indicated before Wermer's Maximality Theorem: a point t lies
in the Choquet boundary if and only if it has a unique representing
measure, which of course must be 6
t
, the unit mass concentrated
at t.
Theorem 9.6.2. Let X be a compact Hausdorff topological space
and let A be a separating function algebra on X that contains
the constants. If t E X, then the following are equivalent:
(i) t E XA.
(ii) The representing measure for t is unique.
Proof. Before we begin the proof we recall the definition of
xA; namely, if for each t E X we set Tt(f) = f(t), f E A, and
= (x* 1 x* E A*, IIx*1I = x*(l) 1), then
xA = (t 1 t E x, T
t
E
where, as usual, denotes the set of extreme points of
For further details the reader is referred back to Section 9.4.
Suppose t E XA. We must show that, if is a representing
measure for t,
then > 0,
and
then = 6 .
t
However, if is such a measure,
II = J X <4J, (s ) = I,
f(t) = Ix f(s)
(f E A).
Moreover, if E C X is any Borel set such that 0 < < 1, then
xi and x
2
defined by
1
xi(f) = SE f(s)
xi(f) = I
X
-
E
f(s) <4J,(s)
(f A),
258 9. Boundaries
are elements of such that
f(t) = Tt(f) = + [1 -
1
Since T
t
E ext(B
l
), we conclude that
Tt(f) = Ix f(s)
= f(s)
= I
X
-
E
f(s)
(f E A).
(f E A).
Then, appealing to Lemma 9.4.2 and arguing precisely as in the proof
of Theorem 9.4.1, we see that the support of must be a single
point. But since
f(t) = Ix f(s)
(f E A),
the support of must be the singleton set ttl, that is = 6
t
.
An even simpler argument gives the desired conclusion in the
case that for each Borel set E C X we have = 1. Thus part (i)
of the theorem implies part (ii).
Conversely, suppose t has a unique representing measure,
which of course must be 6
t
We need to show that T
t
E
If xi,xi E and 0 < a < I are such that T
t
= axi + {I - a)xi'
and if and are representing measures for xi and xi,
respectively, we see that 6
t
= + (1 - since -"1 + (1 -
is a representing measure for t, and the representing measure for
t is unique. However, as mentioned in the proof of Theorem 9.4.1,
the measure 6
t
is an extreme point of the closed unit ball in M(X),
whence we conclude that 6
t
= = Consequently, T
t
= xi = xi
1
and T
t
E ext{B
I
); that is, t E XA. 0
In view of our discussion of A(D) at the beginning of Section
9.6, we see at once from the preceding theorem that the Choquet boun-
dary of A(D) is precisely f. Thus, since the Bishop boundary of
9.7. Characterizations of the Choquet Boundary 259
A{D) exists and is also equal to r, we have pA{D) = XA{D). We
shall see at the end of the following section that this is always
the case for uniform algebras on compact metric spaces.
9.7. Characterizations of the Choguet Boundary. The main pur-
pose of this section is to establish a number of characterizations
of the Choquet boundary for uniform algebras. The characterizations
will consist of four necessary and sufficient conditions for a point
to lie in the Choquet boundary. Before we can state and prove the
theorem we need a number of preliminary results and definitions.
Definition 9.7.1. Let X be a compact Hausdorff topological
R
space. Then C (X) will denote the Banach space over lR consisting
of all the real-valued continuous functions on X with the supremum
norm. A linear functional F on CR{X) is said to be positive if
R
F{f) > 0 whenever f E c (X) and f{t) ~ 0, t E X.
Lemma 9.7.1. Let X be a compact Hausdorff topological space.
If F is a positive linear functional on CR(X), then F is con-
tinuous and IIFII = F{l).
Proof. Since F is positive, it is apparent that, if f and
g belong to CR(X) and f(t) ~ get), t E X, then F(f) > F(g).
Thus, if f E CR{X) is such that IIfll < 1, then it follows easily
CD -
that -F(l) ~ F(f) ~ F(l), whence IF(f), < F(l). Consequently
F is continuous and UFII = F(l).
o
In the statement and proof of the next lemma we shall use r(x*)
to denote the set of representing measures for a given continuous
linear functional x*.
Lemma 9.7.2. Let X be a compact Hausdorff topological space,
let A be a uniform algebra on X, and let T E 6(A). For each
h CR(X) define
~ T , h ) = sup{Re[T(f)] I f ( A, Re[f(s)] < h(s), s E xl
260 9. Boundaries
and
Q(T,h) = inf{Re[T(f)] I f A, Re[f(s)] h(s), s ( xl.
If (r(T), then
Q(T,h) < Ix h{s) < Q(T,h)
and if h (eR(x) and a (R are such that Q(T,h) < a < Q(T,h),
then there exists some E reT) for which
a = Ix h(s)
Moreover, for each h ( eR(x),
= inf{I
x
h(s) I E reT)}
and
Q(T.h) = sup{!x h(s) I E reT)}.
Proof. Suppose h (CR(X). If E reT), then for each f E A
such that Re[f(s)] h(s), sEX, we have
since > O. Hence
Re[T(f)] = Re[I
x
f(s)
= Ix Re[f(s)]
< Ix h(s)
< Ix h(s)
Similarly we deduce that
Ix h(s) < Q(T,h)
which proves the first assertion of the lemma.
( r(T.
E r(T,
Next we note that Q(T,.) is a positive homogeneous subadditive
9.7. Characterizations of the Choquet Boundary
261
function from CReX) to that is,
Q(T,bh) = bQ(T,h),
Q(T,h + g) < Q(T,h) + Q(T,g)
R
(g,h E C (X); b > 0).
We shall need this easily verified observation in a moment. Now let
h E CRex) and suppose a En is such that Q(T,h) < a < Q(T,h).
Set W = (bh I b E IR). Evidently W is a linear subspace of CR(X).
We define a linear functional F on W by F (bh) = ba, b
o 0
From the restrictions on a we see at once that F (bh) = ba < Q{T,bh),
o -
b E kL Thus, since QeT,) is a positive homogeneous subadditive
function on eReX), we may appeal to the Hahn-Banach Theorem for
real linear spaces [L, p. 83]to deduce the existence of a linear
functional F on CR(X) such that F restricted to W is equal
- R
to Fo and such that Feg) QeT,g), g E C (X). In particular, if
g (CR(X), then Fe-g) Q(T,-g), whence
F(g) = -Fe-g) > -Q(T,g) = Q(T,g).
Moreover, an easy computation using the definition of reveals
that, if g E CReX) and g(s) > 0, sEX, then
o = < peg),
where 0 is used to denote both the number zero and the function in
CRex) that is identically zero. Hence F is a positive linear
functional on eRex), and so, by Lemma 9.7.1, F is a positive
continuous linear functional on CR(X) and IIFIi = Fel). Appealing
to the Riesz Representation Theorem for such functionals [L, p. 109],
we deduce the existence of some E M(X), 0, such that
peg) = Ix g(s)
and for which
= Ix
262
9. Boundaries
Furthermore, if f E A, then Re[f] (CR(X) and =
Q(T,Re[f]) = Re[T(f)]. The details of the latter assertions are
left to the reader. Hence we see that
Re[T(f)] = F(Re[f])
= Ix Re[f(s)]
= Re[i
x
f(s)
(f E: A),
from which it follows at once that
T(f) = Ix f(s)
(f ( A)

Moreover, since T(f) = f(T), we see that
11TH = 1 = T (1) = I x (s) =
and so E reT). Finally we observe that
Ix h(s) = F(h) = Fo(h) = a,
and the proof of the second assertion of the lemma is complete.
The last assertion of the lemma is now apparent.
One more definition is necessary before we can turn to the
main theorem.
o
Definition 9.7.2. Let X be a locally compact Hausdorff topo-
logical space and let A be a separating function algebra on X.
A subset E C X is said to be a peak set for A if there exists
some f E A such that E = (t I t E X, If(t)1 = IIflle] = E
f
. A
point t E X is said to be a weak peak point for A if there exists
a family of peak sets (Ef] for A such that n E
f
= (t). A
a a a
point t E X is said to be a strong boundary point for A if for
each open neighborhood U of t there exists some f E A such that
II f lie = I f (t) I = I and 1 f (s) 1 < I, sEX - U.
9.7. Characterizations of the Choquet Boundary
263
Clearly a peak set E
f
is a nonempty compact subset of X
provided f is not identically zero. The definition of a strong
boundary point should be compared with the characterization of points
in the ilov boundary given in Corollary 9.2.2. Our main result is
the following theorem:
Theorem 9.7.1. Let X be a compact Hausdorff topological space
and let A be a uniform algebra on X. If t E X, then the follow-
ing are equivalent:
(i) Given 0 < a < a < 1, if U is an open neighborhood of
t, then there is some f E A such that IIfll., < 1, If(t)l > a, and
If(s)I < a, sEX - U.
(ii) If U is an open neighborhood of t, then there exists
some f E A such that IIfU., 1, lfet)1 > 3/4, and If(s)l < 1/4,
sEX - U.
(iii) t is a strong boundary point for A.
(iv) t is a weak peak point for A.
(v) t E xA.
Proof. Obviously part (i) implies part (ii), so we assume that
part (ii) holds and let U be an open neighborhood of t. We must
construct some g E A such that Iget)1 = IIgll = 1 and Ig(s)1 < 1,
.,
sEX - U. To accomplish this we shall first define a sequence of
open neighborhoods {Uk] of t and a sequence {gk] C A such that
for each k we have
(a)
Uk ::> U
k
+
1

(b)
gk (t) = 1.
(c)
1I1)t1l., < 4/3.
Cd) Ilk(S)' < 1/3, sEX - Uk.
264 9. Boundaries
sup U Ig(s)1 < 1 + 1/(3.2
k
), j = I,2, ... ,k - l.
s J -
k
(e)
The definition of these sequences will be inductive. Set U
I
= U.
Then there exists some fl E A such that IIfIIICD < 1, Ifl(t)1 > 3/4,
and Ifl(s)1 < 1/4, sEX - U
I
. Moreover, we may assume without loss
of generality that we even have fI(t) > since if this is not
true of f l' it certainly is valid for e -1 arg f 1 (t) fl' The other
properties of fl are clearly unchanged by this convention. Now
define gl = fI/fI(t). Clearly gl E A, gl(t) = 1, /lgIlioo < 4/3,
and Igl(s)1 < 1/3. sEX - U
l
. Conditions (a) through (e) are
vacuously satisfied for gl' Suppose that U
I
,U
2
""'U
n
and
gl,g2, ... ,gn have been defined and satisfy conditions (a) through
(e). Then define U 1 as
n+
n n 1
U +1 = n (s I sEn U., Igk(s)1 < 1 + 1 ).
n k = 1 j = 1 J 3 2
n
+
Since t E and gj(t) = 1, j = I,2, ... ,n, it is apparent
that U
n
+
l
is an open neighborhood of t and U
n
+
l
C Un. Let
fIE A be such that IIf 111 < 1, f I(t) > 3/4, and
n+ n+ CD - n+
If
n
+
1
(s)1 < 1/4, sEX - U
n
+
1
, and define gn+I = fn+I/fn+I(t).
As before, it is evident that gn+I E A satisfies conditions (b)
through (d). Moreover, if s E U l' then from the definition of
n+
U it follows that
n+l
(k = 1,2, ... , n) .
Proceeding in this way, we define inductively a sequence of open
neighborhoods (Uk) of t and a sequence (gk) C A that satisfy
conditions (a) through (e). Now define g as g = Ik=Igk/
2k
.
Since A is a uniform algebra, we see that g E A.
Furthermore, since gk(t) = 1, k = 1,2,3, ... , it is obvious
that get) = 1. If sEX - UJ then sEX - Uk' k = 1,2,3, ... ,
and so
9.7. Characterizations of the Choquet Boundary
265
Finally, suppose s E U.
CD
If s E ~ = I U k then for each j
Ig(s)1 < sup 19.(s)1
J -sEU J
k
< I + 1
3":2k
(k = j + 1,j + 2, ... ),
whence we conclude that Ig.(s)1 < 1, j = 1,2, .. Thus, if
J -
s E Ok=lUk' then Ig(s)1 < 1. On the other hand, if s E U but
s ~ ~ = l U k then there exists some m I such that s E: U
m
but
s ~ U +1. Then sEX - U ., j = 1,2,3, .. , and so conditions
m m+J
(c) through (e) combined entail that
m-l Igk(s)1 Igm(s)1 CD Igk(s)l
Ig(s)1 < ~ + + t k
k=l 2k 2
m
k=m+1 2
I m-1 I 4 I CD 1
< (1 + -) 1: - + - + - ~ -
3.2
m
k=l 2k 3.2
m
3 k=m+l 2k
4 1
+-+--
3.2
m
32
m
=1 __ 1.....",..._
3.2
2m
-
1
<1.
Thus g E A is such that
11&11 = get) = 1 and 19(s) I < 1,
CD
5 ~ X - U, and so we conclude that t is a strong boundary point
for A.
Next suppose that t is a strong boundary point for A and
let (U) be a family of open neighborhoods of t such that
a-
n U = {t}. Since t is a strong boundary point, we see that for
0'0'
each U there exists some f E A such that IIf II = If (t) I = 1
a a rteD a
and If (s)1 < 1, sEX - U. Setting E
f
= {s I sex, IIf II = If (s)ll,
a a a- am a-
we see that t E E
f
C U, that E
f
is a peak set, and n E
f
= (t).
a a a a a
Hence t is a weak peak point.
266 9. Boundaries
Now suppose t is a weak peak point and let ~ be a repre-
senting measure for t. If U is an open neighborhood of t, then,
since t is a weak peak point, there exist some f E A and a closed
set E c: X such that t E: E C U and E = (s I sEX, IIfU = If(s) 11
..
Without loss of generality we may assume that 1 = IIfU = f(t)
..
Then we see at once that
o = 1 - f(t)
= Ix [1 - f(s)] ~ s )
= Ix Re[l - f(s)] ~ s ) ,
as ~ > 0 by Proposition 9.6.1. But since If I assumes its maxi-
mum precisely on E ~ we deduce that the nonnegative function
Re[l - f) cannot be identically zero on X - U. Thus ~ X - U) = o.
Since U is an arbitrary open neighborhood of t and ~ is regular,
it follows at once that the support of ~ is the singleton set (t),
and so ~ = Qt. Hence the representing measure for t is unique,
and t E XA by Theorem 9.6.2.
Finally, suppose that t E XA and 0 < a < ~ < 1. If U is
an open neighborhood of t, then, using Urysohn's Lemma [W
2
' p. 55]
it is easy to see that there exists some h E CR(X) such that
h(t) = 0, h(s) < log a, sEX - U, and h(s) ~ 0, sEX. Conse-
quently, since the only representing measure for t is ~ = 0t by
Theorem 9.6.2, we see from Lemma 9.7.2 that
o = h(t) = Ix h(s) dOtes)
= inf(I
x
h(s) ~ s ) I ~ E r(Ttl]
= Q,(T ,h)
= sup{Re[Tt(f)] I f E A, Re[f(s)] ~ h(s). s E xl
= sup(Re[f(t)] I f E A, Re[f(s)] ~ h(s), s E xl.
9.7. Characterizations of the Choquet Boundary
267
Hence, since log < 0, we conclude that there exists some g A
such that Re[g(s)] < h(s), sEX, and Re[g(t)] > logS. Let f = ego
Since A is a uniform algebra, it is evident that f E A, and some
elementary computations reveal that IIfli. I, If(t) I > (:), and
If(s)1 < a, sEX - U.
This establishes the equivalence of all five parts of the theo-
rem and completes the proof.
o
Theorem 9.7.1, combined with Theorems 9.6.2 and 9.4.1, gives
a rather complete description of the Choquet boundary of a uniform
algebra. Moreover, the results allow us to show quite easily that
the Bishop boundary exists for any uniform algebra A on a compact
metric space. First we require a lemma.
Lemma 9.7.3. Let X be a compact Hausdorff topological space
and let A be a uniform algebra on X. If t E X, then the fol-
lowing are equivalent:
(i) t is a peak point for A.
(ii) t is a strong boundary point for A and it) is a G
6
-set.
Proof. If t is a peak point, then it is clearly a strong
boundary point. Moreover, there exists some f E A such that
IIfll = If(t)1 = 1 and If(s)1 < 1, s t. For each positive integer

n let U = {s 1 sEX, If(s)1 > 1 - lIn). Evidently each U is
n n
open, t E Un' and n:=IU
n
= ttl. Thus ttl is a G
6
-set.
Conversely, suppose t is a strong boundary point and ttl is
a G
6
-set. Then there exists a decreasing sequence {Un) of open
neighborhoods of t such that n:=lU
n
= ttl. For each Un there
is some f E A such that IIf II = f (t) = 1 and If (s)1 < 1,
n n. n n
s X - U. On replacing f by a suitable power of itself, we may
n n
assume without loss of generality that If (5)l < 1/2, sEX - U
n n n
Let g = Clearly g E A and get) = 1. If sEX and
268 9. Boundaries
and s; t,
that s ( U
m
Ig(s)1 < 1/2,
then either s ~ U
1
or there exists some m > 1
and s ~ U 1. In the first instance we have
m+
whereas in the second we see that
mIl CX) 1
< 1: + - E
n=l 2
n
2 n=m+l 2
n
I
= I --
2
m
+
1
<1.
Hence t is a peak point for A.
such
o
Theorem 9.7.2. Let X be a compact Hausdorff topological space
and let A be a uniform algebra on X. If X is metrizable, then
pA = xA.
Proof. Since X is metrizable, it is obvious that every point
of X is a Go-set, and so by c ~ m a 9.7.3 a point t E X is a peak
point if and only if it is a strong boundary point. However, by
Theorem 9.7.1, the Choquet boundary, which always exists, consists
precisely of the strong boundary points, whence pA = XA.
o
As already noted in Example 9.3.5, if X is not metrizable,
then the Bishop boundary may fail to exist. Using Theorem 9.7.ICi)
or Cii), one can show, when X is compact and metric, that the
Choquet boundary of a uniform algebra A on X is a Go-set. As
indicated in the remarks following Corollary 9.4.1, this may not be
the case if X is not metrizable. We state this result as the next
corollary, whose proof is left to the reader.
9.8. Representing Measures and the ~ i l o v Boundary
269
Corollary 9.7.1. Let X be a compact Hausdorff topological
space and let A be a uniform algebra on X. If X is metrizable,
then pA = XA is a G
6
-set.
9.8. Representing Measures and the ~ i l o v Boundary. Returning
again to the algebra A(D), we recall that, if ,E int(D), then
there exists a representing measure ~ for , such that v, E M(l)
= M[aA(D)]. The next theorem shows that an analogous result is valid
for any separating function algebra on a compact Hausdorff topologi-
cal space that contains the constants.
Theorem 9.8.1. Let X be a compact Hausdorff topological space
and let A be a separating function algebra on X that contains
the constants. If t E X, then there exists a representing measure
~ for t such that ~ E M(oA).
Proof. Let B(aA) denote the subalgebra of C(aA) consisting
of the restrictions t6 aA of the elements in A. From the nature
of the ~ i l o v boundary it is apparent that the restriction mapping
is an isometric algebra isomorphism of A onto B(aA) and that aA
is a compact Hausdorff topological space. We define a continuous
linear functional 1t on B(aA) by Tt(fl
oA
) = f(t), f E A, where
flaA denotes the restriction of f to oA. Clearly IITtll =
Tt(llaA) = 1. By a consequence of the Hahn-Banach Theorem [L, p. 87]
we see that T
t
can be extended to a continuous linear functional
on C(oA) without increasing the norm, and so, by the Riesz Repre-
sentation Theorem [L, p. 109], there exists some ~ E M(oA) such
that
(f E A).
Obviously ~ is a representing measure for t, and the proof is
complete.
o
This result leads to the following corollary about commutative
Banach algebras:
270
9. Boundaries
Corollary 9.8.1. Let A be a commutative Banach algebra. If
T E then there exists some E MCoA) such that
Ci) > o.
(ii) (oA) :5. 1.

(iii) T(X) = X(T) = foA lew) x E A.
Proof. If A has an identity, then the conclusion follows at

from Theorem 9.8.1 applied to A = A. Suppose A is without iden-
tityand consider the algebra A[e]. From Theorems 3.2.1 and 3.2.2
we recall that = U (T J. We claim that
Q)
oA c oA[e] c oA U {T J.
Q)
Indeed, since the embedding of A{A) into A(A[e]) is a homeomor-
phism and oA is closed, we see that oA can be considered as a
closed subset of A(A[e]) that is the minimal closed boundary for

A C A[e] . Thus oA c oA[e]. Since the functions ae obviously
attain their maximum absolute value only on {T), we conclude that
CXI
oA[e] c oA U {T J.
CD
Applying Theorem 9.8.1 to A = A[e] , we deduce the existence
of some E M(oA[e]) such that > 0, = 1, and for
which

X(Te) = foA[e] x{w)
(x E A),
where, as usual, T denotes the unique complex homomorphism on
e
A[e] determined by T. If = 0, then we let be v
CD
restricted to oA. while if J) 0, then we set =
CD
- V{{T )6
T
. In this case it is apparent that = 0, and
CD CD CD
so, in either case, restricted to oA defines an element of
M(oA) such that > 0, :5. I, and for which

T(X) = X{T) = faA x(w)
(x E A).o
9.S. Representing Measures and the ilov Boundary
The representing measure given by Theorem 9.8.1, of course,
need not be unique. Indeed, in view of Theorem 9.6.2, it will be
unique precisely when t belongs to the Choquet boundary.
271
An analog of Theorem 9.S.l with the ilov boundary replaced by
the Choquet boundary is also valid. Its proof, however, depends on
a deep result, generally referred to as the Choquet-Bishop-deLeeuw
Theorem, which could not be fully presented without a considerable
digression. Consequently we shall content ourselves with a state-
ment of the theorem and refer the interested reader to [Ph; S, pp.
5S and 59] for further details.
Theorem 9.S.2 (Choquet-Bishop-deLeeuw). Let X be a compact
Hausdorff topological space and let A be a uniform algebra on X.
If t E X, then there exists a representing measure ~ E M(X) for
t such that ~ E ) = 0 whenever E X is a Baire set or a G
6
-set
that is disjoint form xA.
Extensive discussions of the use of representing measures in the
study of polynomial and rational approximation as well as the signi-
ficance of many of the other concepts introduced in this chapter for
the investigation of function algebras can be found in [B, Ga, Lb,
S, wm
l
, WID
2
]. However, we shall not pursue the matter further since
it would take us too far afield from the general theme of this volume.
CHAPTER 10
B*-ALGEBRAS
10.0. Introduction. A B*-algebra is a Banach algebra on which
there exists a certain antilinear idempotent mapping called an invo-
lution such that a particular norm identity is valid with respect to
this mapping. Such algebras are, in general, rather tractable objects
of study. For example, we shall prove below that, if A is a com-
mutative B*-algebra. then the Gel'fand transformation on A is an
isometry of A onto C (A(A that preserves the involution on A.
o
As we shall see shortly, the involution on C (A(A is just complex
o
conjugation. This result, called the Commutative Gel'fand-Naimark
Theorem, should obviously be a valuable tool in investigating commu-
tative B*-algebras, and we shall devote the final five sections of
this chapter to examining its application to some specific B*-algebras.
In Section 10.3 we shall concentrate our attention on the commu-
tative Banach algebra Ao(G) obtained as the closure in L(L
2
CG
of the continuous linear transformations T
f
L(L
2
(G, f E LICG),
defined by TfCg) = f * g, g (L
2
(G). As usual, G is a locally
compact Abelian topological group. The results we shall obtain in
Section 10.3 about A (G),
o
which is a B*-algebra, will enable us
to finally prove Plancherel's Theorem in the succeeding section. In
Section 10.5 we shall establish the Pontryagin Duality Theorem, which
depends directly on the regularity of LICG) and hence indirectly
on Plancherel's Theorem. This theorem asserts, as indicated in

Section 4.7, that CG) is topologically isomorphic to G.
In the last two sections we shall utilize the Commutative Gel'-
fand-Naimark Theorem to obtain a spectral decomposition theorem for
272
10.1. B*-Algebras
273
self-adjoint continuous linear transformations on Hilbert spaces
and to study the commutative Banach algebra of almost periodic func-
tions on a locally compact Abelian topological group G and the so-
called Bohr compactification of G.
10.1. B*-Algebras. In this section we shall define algebras
with involution and B*-algebras, give a number of examples, and es-
tablish some elementary properties of B*-algebras. We make no gen-
eral assumptions of commutativity.
Definition 10.1.1. Let A be a Banach algebra. Then A is
said to be a Banach algebra with involution if there exists a mapping
* : A - A such that for any x,y E A and a ~ we have
(i) (x + y)* = x* + y*.
(ii) (ax)* = ax*.
(iii) (xy) * = y*x*.
(iv) (x*)* = x** = x.
A Banach algebra A with involution is said to be a B*-algebra if
IIx*xll = IIx1l
2
, x ( A.
It is apparent that, if A is a commutative Banach algebra
that is self-adjoint and semisimple, then A is a Banach algebra
-- ..
with involution. Indeed, in this case it happens that x A for
.. ..
each x E A. If Y is the unique element in A such that y = x,
then one sets x* = y. It is easily verified that the mapping
* : A - A is an involution. Thus, for example, if X is a locally
compact Hausdorff topological space, then c (X)
o
is a Banach alge-
bra with involution, where for f Co(X) we set f*(t) = f(t),
t E X; and if G is a locally compact Abelian topological group,
then LI(G) is a Banach algebra with involution provided we define
f* almost everywhere by the formula f*(t) = f(-t), t G. Obvious-
ly Co(X) is also a B*-algebra. However, LI (G) is not generally
274 10. B*-Algebras
a B*-algebra. Although we shall not prove this in detail, it is an
immediate consequence of the Commutative Gel1fand-Naimark Theorem

(Theorem 10.2.1) and the fact that LI(G) is properly contained

in Co (G) unless G is a finite group [Ru
l
, p. 90].
On the other hand, there exist commutative Banach algebras with
involution that are not self-adjoint. One such algebra is A(D).
It is easily seen that this is an algebra with involution provided
we set f*(I) = f(Z), zED, for each f E A(D). It is, however,
not self-adjoint, since, if f E A(D), then f E A{D) only if f
is a constant, where of course fez) = fez), zED. Recall that the
Gel'fand transformation on A(D) is just the identity mapping. That
A(D) is not a B*-algebra will be immediately apparent from the
Commutative Gel'fand-Naimark Theorem.
An important example of a B*-algebra is the Banach algebra
A = L(V) of all the continuous linear transformations from a Hilbert
space V over C to itself. As indicated in Example 1.2.11, L(V)
is a Banach algebra with identity. It is a Banach algebra with invo-
lution since the mapping * : LcY) L(V) defined by T T*,
T E L(V), where T* denotes the Hilbert space adjoint of T, satis-
fies the requirements of an involution as set forth in Definition
10.1.1. Moreover, L(V) is a B*-algebra. Indeed, if T E LcY),
then IITII = IIT*II fL, p. 96], and so, on the one hand, we see that
IIT*TII < IIT*III1TII IITII2, while on the other hand, by a consequence
of the Hahn-Banach Theorem [L, p. 89] and the Riesz Representation
Theorem for Hilbert spaces [L, p. 390], we have
IIT*TII = sup IIT*T(x)1I
IIxll=1
= sup [ sup I(T*T(x),y)l]
IIxll=1 lIyll=1
sup I(T*T(x),x>1
- IIxll=1
10.1. B*-Algebras
275
= sup I<T(x),T(x)l
lIxll=l
= sup IIT(x)1I
2
IIxll=l
=
IITII2,
where <.,.) denotes the inner product in V. Hence IIT*TII = IITII2,
T E L(V). Furthermore, it is evident that, if B is a norm-closed
subalgebra of L(V) with the property that T E B implies T* E B,
then B is also a B*-algebra. Such a subalgebra of Lev) is gener-
ally called a C*-algebra.
One more definition is needed before we can prove some element-
ary properties of B*-algebras.
Definition 10.1.2. Let A be a Banach algebra with involution.
A subset E C A is said to be symmetric if x E E implies x* E E.
An element x E A is said to be self-adjoint if x* = x.
Thus, if V is a Hilbert space over ~ then those T E L(V)
that are self-adjoint as elements of the B*-algebra L(V) are pre-
cisely the elements of L(V) that are self-adjoint as continuous
linear transformations. A C*-algebra is bbviously a norm-closed
symmetric subalgebra of L(V) for some Hilbert space V over C.
Proposition 10.1.1. Let A be a B*-algebra and let x E A.
Then
(i) 0* = O.
(ii) If A has an identity e, then e* = e.
(iii) If A has an identity e and x E A is regular, then
-1 -1
x* is regular and (x*) = (x )*.
and
(iv) If x E A is quasi-regular, then x* is quasi-regular
(x*) = (x )*.
-1 -1
276 10. B*-Algebras
(v) if and only if ,E a(x*).
(vi) IIx*1I = IIxli.
(vii) If x is self-adjoint, then
(viii) If x is self-adjoint, then a(x)
Proof. The proofs of parts (i) through (v) follow almost imme-
diately from the mapping properties of the involution. The details
are left to the reader. Part (vi) is evident if x = 0, and if
x 0, then lixll
2
= IIx*xll < IIx*lIl1xll, whence IIxll < IIx*U. Conse-
quently IIx*1I < IIx**U = IIxll, and so IIx*1I = IIxU.
From the Spectral Radius Formula (Theorem 2.2.2) we know that
IIxlia = limnllxnUl/n for any x A. If x is self-adjoint, then
IIx
2
11 = IIx*xll = IIxU
2
, from which it follows easily that IIx
2n
ll = IIxll2n,
n = 1,2,3, . Thus
n n n
In proving part (viii) of the proposition we may assume, with-
out loss of generality, that A has an identity e. This is possi-
ble because, by Theorem 2.1.1, aA(x) = aA[e] (x). Suppose a + ib
belongs to a(x) and let pt(x) = x + ite and qt(x) = x - ite,
t Ent Evidently a + i(b + t) (a[Pt(x)], whence by part (v) we
see that a - i(b + t) a[Pt(x)*] = a[qt(x)], as x is self-adjoint.
Thus by the Polynomial Spectral Mapping Theorem (Theorem 2.2.1) we
2 2 2 2
deduce that a + (b + t) (a[ptqt(x)] = a(x + t e). Hence from
Theorem 2.1.2 we see that
22222
a + (b + t) = a + b + 2bt + t
< IIx
2
+ t
2
ell
< IIx
2
11 + t
2
j
that is, a
2
+ b
2
+ 2bt < IIx
2
11, t E If\.
10.2. The Gel'fand-Naimark Theorem
277
This estimate, however, obviously leads to a contradiction unless
b = o. Therefore a(x) c IR.
o
10.2. The Gel'fand-Naimark Theorem. We shall now state and
prove the commutative version of the Gel'fand-Naimark Theorem, which
asserts that the Gel'fand transformation on a commutative B*-algebra
A is an isometric isomorphism of A onto C (A(A)) that preserves
o
the involution. We shall also mention briefly the noncommutative
version of this theorem. We choose not to discuss in detail the
noncommutative theorem since our applications of the Gel'fand-Naimark
Theorem in the succeeding sections will all be in a commutative con-
text.
One more definition is required.
Definition 10.2.1. Let A and B be Banach algebras with in-
volutions * and +, respectively. A homomorphism : A - B is
said to be a *-homomorphism if = x E A.
If X is a locally compact Hausdorff topological space, then
we shall consider C (X) as a Banach algebra with involution, as
o
discussed in the preceding section; that is, the involution on Co(X)
is given by complex conjugation.
Theorem 10.2.1 (Commutative Ge1'fand-Naimark Theorem). Let A
be a commuta'ive B*-algebra. Then the Gel'fand transformation on
A is an isometric *-isomorphism of A onto C (A(A)).
o
Proof. If x E A is self-adjoint, then from Proposition 10.1.1
(vi) and the Beurling-Gel'fand Theorem (Corollary 3.4.1) we see that

IIxll = IIxli. Now, if x is any nonzero element of A, then clearly
.
y = x*x is self-adjoint and so Ilyll = lIyll. Thus by Proposition

10.1.I(vi) and the fact that A is a B*-algebra we see that
IIxll
2
= IIx*xll = lIyll

I:
278 10. B*-Algebras
=
CD

< lI(x*) II IIxli
- CD CD

< IIx*lIl1xll
- CD

= IIxllllxll ,
CD

whence we have IIxll < IIxli. Since the reverse inequality is always
- CD
valid, we conclude that IIxli = IIxll , x E A.
CD
Hence the Gel'fand transformation is an isometric isomorphism

of A onto A, and, as always, A separates the points of
.. ..
and ZeAl = (T IT' A(A), X(T) 0, X , AJ ,. Moreover, if
x E A, then x = y + iz, where y,z E A are self-adjoint. Indeed,
y = (x + x*)/2 and z = (x - x*)/2i. Consequently from the proper-
ties of the involution * and Proposition 10.1.I(viii) we deduce
that

.-
(x*) = [(y + iz)*] = y iz = y + iz = x.
Thus we see that the Gel'fand transformation is an *-isomorphism and
..
that A is closed under complex conjugation. Therefore by the
Stone-Weierstrass Theorem [L,
which completes the proof.
.-
p. 332], we conclude that A = C
o
o
Corollary 10.2.1. Let A a commutative B*-algebra. Then
A is a Tauberian, self-adjoint, semisimple, regular, commutative
Banach algebra.
The problem of spectral synthesis is also quite easily solved
in commutative B*-algebras, as seen from the next corollary.
Corollary 10.2.2. Let A be a commutative B*-algebra. Then
(i) If E C is closed, then E is a set of spectral
synthesis.
(ii) If I C A is a closed ideal, then I = k[h(I)].
10.2. The Gel'fand-Naimark Theorem
279
Proof. In view of our discussion of spectral synthesis in Sec-
tion 8.3, it is apparent that we need only prove part (ii). Moreover,
to establish part (ii) it suffices, by Proposition 7.1.I(vii) and
Theorem 8.1.I(ii), to show that I [hell] is dense in k[h(l)]
o
since I [h(I)] C I C k[h(I)] and I is closed. Recall that
o
I [h(I)] consists of all those x (A such that x vanishes ident-
o
icallyon some open set Ox ~ h(l). With this in mind, let
y 'k[h(I)] and let e > O. Obviously the set

U = {T I T E A ) , (y(T)1 < e)
is an open subset of ~ A ) such that U h(I), and X - U is
compact. For each T E A ) let yeT) = p(T)e
i9
(T), where
peT) > 0 and 0 S 9CT) < 2n, and define f on ~ A ) by
f(T) = 0 for T E U,
f(T) = [peT) - e]e
i9
(T) for T E X - u.
It is easily verified that f E C ~ C A ) ) C C ~ A ) ) .
c 0
Hence, by the
Commutative Gel'fand-Naimark Theorem, there exists some x E A such

that x = f. Evidently x E I [h(I)], and again appealing to the
o
preceding theorem, we see that lIy - xII = lIy - XIlCD ~ e. Hence
lo[h(I)] is dense in k[h(I)], and so I = k[h(I)].
o
Thus we see that the closed ideals in a commutative B*-algebra
are precisely the sets I of the form I = keEl for some closed
set E C ~ A ) . These observations provide a new proof of this fact
for the B*-algebra C(X), X being a compact Hausdorff topological
space (Theorem 8.3.2).
As indicated at the beginning of this section, we shall not
discuss the noncommutative version of the Gel'fand-Naimark Theorem
in any detail, but shall only give a precise statement of it. The
reader who is interested in a more detailed treatment is referred
to [HoI' pp. 83-88; N, pp. 309-314; Ri, pp. 239-244]. More detailed
discussions of B*-algebras are also available in these references.
280
10. B*-Algebras
Theorem 10.2.2 (Noncommutative Gel'fand-Naimark Theorem). Let
A be a B*-algebra. Then there exists a Hilbert space V over ~
and a norm-closed symmetric subalgebra B of L(V) such that A
is isometrically *-isomorphic to B.
The invOlution in L(V) is naturally that given by the Hilbert
space adjoint. Thus the theorem says that every B*-algebra is iso-
metrically *-isomorphic to a C*-algebra.
At a number of key points in the preceding chapters we have
utilized Plancherel's Theorem (Theorem 7.2.1). This was the case,
for example, in proving the regularity of Ll(G) (Corollary 7.2.3),
in proving that LI(G) is a Tauberian algebra (Theorem 8.1.2), in
proving that Ll(G) satisfies Ditkin's condition (Theorem 8.6.1),
and generally throughout our discussion of spectral synthesis in
LI(G). We are now in a position to prove this important theorem,
and the next two sections are devoted to this task. Section 10.3
will be rather technical in nature, but it will provide us with the
necessary machinery to establish P1ancherel's Theorem in the subse-
quent section.
10.3. The B*-Algebra ~ G ) . In this section we wish to dis-
cuss a number of detailed results about a certain commutative alge-
bra of continuous linear transformations on the Hilbert space L
2
(G),
G being a locally compact Abelian topological group. We recall
that L
2
(G) is indeed a Hilbert space with the inner product defined
as
{f,g) = J
G
f(t)g(t) dA(t)
This algebra will be instrumental in proving Plancherel's Theorem
in the next section. We begin with a definition.
Definition 10.3.1. Let G be a locally compact Abelian topolo-
gical group. If f (L1(G), then T
f
(L(L
2
(G)) is defined by
Tf(g) = f * g, g ( L
2
(G).
10.3. The B*-Algebra A (G)
o
It is apparent from Example 1.2.8 that T
f
E L(L
2
(G)) and
281
that IITfll < lifll
l
, f E LI(G), since II T
f
(g)1I
2
< II f ll
1
11g11
2
, g E L
2
(G).
This definition of T
f
allows us to embed the algebra L
1
(G)
in the algebra L(L
2
(G)).
Lemma 10.3.1. Let G be a locally compact Abelian topological
group. If LI (G) - L(L
2
(G)) is defined by = T
f
, f E Ll(G),
then is a norm-decreasing *-isomorphism of LI(G) into L(L
2
(G)).
Proof. It is easily verified that is a norm-decreasing
homomorphism of Ll (G) into L(L
2
(G)). To see that is injec-
tive we suppose that f (LI(G) and T
f
= O. If KeG is any
compact set, then X
K
E L
2
(G) n Lm(G), where X
K
denotes the char-
acteristic function of K, and 0 = Tf(X
K
) = f * X
K
. Since by
Proposition 4.7.2(i) f * X
K
is continuous, we in particular,
Hence from the regularity of the measure E M(G), defined by
= f(-s) dA(s), we conclude that f = O. Thus is an iso-
morphism.
Finally, to see that is an *-isomorphism it suffices to
show that (T
f
)* = T
f
*, f L1(G), where f*(t) = f(-t). But,
given f E Ll(G), if g,h E L
2
(G), then
{(Tf)*(g),h) = (g,Tf(h)}
= iG g(t)f * h(t) dA(t)
= iG g(t)[i
G
f(t - s)h(s) dA(s)] dA(t)
= iG h(s)[J
G
f*(s - t)g(t) dA(t)] dA(s)
= (f* * g.h)
282 10. B*-Algebras
By a consequence of the Hahn-Banach Theorem [L, p. 89] and the Riesz
Representation Theorem for Hilbert spaces [L, p. 390] we conclude
that (T
f
)* = T
f
* 0
Thus we see that is a symmetric commutative subalge-
bra of L{L
2
{G that is *-isomorphic to LI{G). We now define
the algebra Ao{G).
Definition 10.3.2. Let G be a locally compact Abelian topo-
logical group. Then Ao(G) is the norm closure in L{L
2
(G of
= {T
f
I f E Ll{G).
Lemma 10.3.2. Let G be a locally compact Abelian topological
group. Then A CG) is a commutative B*-algebra.
o
Proof. Clearly Ao{G) is a norm-closed commutative subalgebra
of L(L
2
(G. It is symmetric, and so a B*-algebra, since
is symmetric and the involution in L{L
2
{G is continuous. 0
From Corollary 10.2.1 we see at once that A (G) is a Tauberian,
o
self-adjoint, semisimple, regular, commutative Banach algebra. Next
we wish to identify the maximal ideal space of A (G).
o
Lemma 10.3.3. Let G be a locally compact Abelian topological

group. Then 6{Ao(G is homeomorphic to G.
Proof. If w E 6{AO(G, then define Tw(f) = w{T
f
)
f E L1{G). Evidently Tw E 6(L
I
(G, and so there exists
Y
w
E G such that
= (w 0 (f),
a unique
Moreover, if w
l
,w
2
E 6(Ao(G are such that wl(T
f
) = w
2
(T
f
),
f LI(G), then, since is norm dense in Ao(G),. we con-
clude that wI = w
2
. Thus the mapping from 6(Ao(G to G, defined
by w - Y , w E 6(A (G, is injective. Since the topologies on
W 0
6(A (G and G are just the relative weak* topologies, it is
o
10.3. The B*-Algebra Ao(G) 283
apparent that this mapping is also continuous. To see that the map-

ping is surjective we must show that, given y E G, there exists
some w E A(A (G)) such that y = y.
o w
With this in mind let w E A(A (G)) be and suppose
.00
y E G. We need to define weT) for each TEA (G). If TEA (G)
o 0
and (f) c: LI(G) is a sequence such that lim IITf - Til = 0, then
n n n
define gn = (.,y )e,-y)f, n = 1,2,3, Clearly (g 1 c: L1(G),
Wo n n
and
w (T ) = IG g Ct)(t,y ) dA(t)
o gn n Wo
= SG fn(t)(t,y)
(n = 1,2,3, ).
some straightforward computation reveals that for each
m and n
sup liT (h) - T (h) 112
lIhll2=1 &m gn
sup lIg * h - g * hll2
IIh1l2=1 m n
=
sup IIr(,y )('ry)f 1 * h - [('YWo)(,-Y)fn1 * hll2
IIh1l2=1 Wo m
=
= sup IIf * [(',-Y )(,Y)h] - f * [(',-Y )(.,y)h1l
2
llh1l2=1 m Wo n Wo
= sup IIf * h - f * hll2
IIh1l2=1 m n
= IITf - T
f
II
m n
Hence {T } is a Cauchy sequence in A (G), and so converges. In
gn 0
particular, the sequence of numbers {w (T ) is convergent, and
o gn
we define weT) = lim w (T ). The preceding argument shows that
n 0 gn
weT) is well defined. Indeed, if C LlCG) is a second sequence
such that lim IITf' - Til = 0 and g' = C",y )(,-y)f', n = 1,2,3, . ,
n n n Wo n
then, as above, we deduce that liT - T ,II = IITf - T
f
,II for each n,
gn gn n n
284 10. B*-Algebras
whence we have lim w (T ) = lim w (T ,). In this way, given
n 0 gn n 0 gn
y E G, we obtain a well-defined linear functional w on A CG).
o
We claim, moreover, that w is a homomorphism on the following
grounds: Suppose T,S ( A (G) and (f J and (f'l are sequences
o n n
in Ll(G) such that lim IITf - Til = 0 and lim IITf' - 511 = o.
n n n n
Then the estimates
~ IITS ... T f 511 + liT f 5 - T f T f' II
n n n n
< IIsliliT - T
f
II + IITf 11115 - T
f
,1I
n n n
reveal that limn"TS - Tfn * f.
1l
= O. Furthermore, if
are defined as before, then itnis readily verified that
(,YWb)C,-y)(f
n
* f ~ , whence
weTS) = lim w (T ,)
n 0 gn * gn
= lim w (T T ,)
n 0 gn gn
= lim w (T ) lim w (T ,)
n 0 gn n 0 gn
= w(T)w(S),
g and g'
n n
g * g' =
n n
as w E a(A (G). Consequently w E A(A (G. Moreover, if
000
f E LI(G) and ~ e take fn = f, n = 1,2,3, ... , then we see, on the
one hand, that
while, on the other hand,
= lim w (T )
n 0 gn
= lim IG fnCt) (t,y) dACt)
n
= IG f(t)Ct,y) dArt),
10.3. The B*-Algebra Ao(G)
285
..
Therefore y = y, and the mapping from (G)) to G is sur-
W 0
jective.
An easy argument, which we shall leave to the reader, shows
that the inverse mapping is continuous, and so (G)) is homeo-
o
..
morphic to G.
o
We now combine the preceding lemmas with the Commutative Gel'fand-
Naimark Theorem (Theorem 10.2.1) to obtain the following
Theorem 10.3.1. Let G be a locally compact Abelian topolo-
gica1 group. Then the Gel'fand
metric *-isomorphism of A (G)
o
transformation on A (G) is an iso-
.. 0
onto Co(G). Moreover, if f Ll(G),
then the Gel'fand transform of T
f
Ao (G) coincides \'1ith the
Fourier transform of f.
Proof. The only point that is perhaps not transparently clear
is the last assertion. But, if f Ll(G), then as seen in the
proof of Lemma 10.3.3, we have
..
(T
f
) (w) = w(T
f
)
= IG f(t) (t,yw) dA(t)
..
= fey )
w
from which the assertion is evident.
(wEIl(A (G)),
o
c
Sinee Ao(G) is semisimple, the last portion of the theorem
also provides a new proof of the semisimplicity of L1(G).
Next we wish to use the B*-algebra Ao(G) to define a Fourier
transform on a certain subspace Qf C(G) and then study some pro-
perties of this transform. Suppose TEA (G). Then there exists
o
a sequence (f 1 C L1(G) such that lim IITf - Til = O. Now in
n n n
some instances the sequence (f) may consist of continuous fune-
n
tions. For instance, if f LI(G) and T = T
f
, then there exists
286 10. B*-Algebras
a sequence (f J c C (G) such that lim IIf - fill = O. Hence.
n c n n
since z. is norm decreasing, we see that limnl/T
fn
- Tfll = O. We
wish to pay special attention to this situation in the case where
we require that the sequence {f) converges uniformly on G. This
n
would occur. for example. if f E C (G). To be more precise. we
c
make the following definition:
Definition 10.3.3. Let G be a locally compact Abelian topo-
logical group. Then CAoCG) is the set of all those f E CCG) for
which there exist some T E Ao(G) and a sequence {fn} C Ll(G) n C(G)
with the following properties:
(i)
(ii)
lim IITf - Til = o.
n n
lim IIf - fll = o.
n n co
Clearly CAo(G) ~ ~ Indeed, if f E Cc(G) and T = T
f
, then
with f = f, n = 1.2,3, . , we see that lim IITf - Til = 0 and
n n n
lim IIf - fll = O. Thus C (G) C CA (G). Moreover, we claim that
n n co c 0
f E CAo(G) is uniquely determined by the T corresponding to it.
To see this, suppose that f,g E CA (G) and there exist some TEA eG)
o 0
and sequences {f} and {g) in Ll(G) n C(G) such that
n n
limllTf - Til = limllT - TI/ = 0
n n n gn
and
limllf - fll = limllg - gil = o.
n m n co
n n
If f ~ g, then obviously we have h = f - g ~ 0 and
limll(f - g ) - hll = o.
n n m
n
Since Cc(G) is norm dense in LI(G) and the dual space of LI(G)
can be identified with Lm(G) [OSI' pp. 289 and 290], it follows
easily that there exists some k E C (G) such that h * k ~ O.
c
Moreover, the estimate
10.3. The B*-Algebra Ao(G)
287
lI(f - g ) * k - h * kll < lI(f - g ) - hll IIklll
n n CD- n n CD
reveals that lim lI(f - g ) * k - h * kll = O. Thus, in particular,
n n n CD
no subsequence of {(f - g ) * kl converges almost everywhere to
n n
zero, as h * k O. However, since k (CcCG) c L
2
(G) , (fnl C Ll(G),
(g 1 C Ll(G). and lim IITf - T II = 0, we see from the estimate
n n n h
IICf
n
- gn) * kll2 = II
T
f
Ck) - Tg (k)1I
2
n n
< IITf - T II Ilk II 2
n gn
that lim IICf - g ) * kll2 = O. Standard results of integration
n n n
theory [DS
l
, pp. 122 and 150] allow us to deduce that the sequence
{(f
n
- gn) * k] has a subsequence that converges almost everywhere
to zero, contradicting the preceding observation. Consequently f = g.
Thus we see that f (Cho(G) is uniquely determined by the
T E Ao(G) corresponding to f. To emphasize this fact and to indi-
cate the correspondence between f E CA CG) and the appropriate
o
T (Ao(G) we shall frequently write f = fT' f E CAo(G). A similar
argument employing the norm density of CcCG) in L
2
(G) shows that.
if f E CAo(G) and f = fT = fS' then T = S. The details are
left to the reader.
It is now an easy matter to define a Fourier transformation on
..
CA (G). Since f has already been used in a different context, we
o
shall use , to denote this transform.
Definition 10.3.4. Let G be a locally compact Abelian topO-
q.
logical group. If f = fT E CAoCG) , then we set t = T, where T
denotes the Gel'fand transform of TEA (G).
o
Clearly 'E C (G) by Theorem 10.3.1.
o
Lemma 10.3.4. Let G be a locally compact Abelian topological
group. Then
Ci) CAoCG) is a translation-invariant linear subspace of C(G).
288 10. B*-Algebras
(ii) C (G) C Ll(G) n C(G) C CA (G).
c 0
(iii) If f = fT (CAo(G) and f*(t) = f(-t), t E G, then
f* ( CA (G) and f* corresponds to T* E A (G).
o 0
(iv) If f = fT E CAo(G) , then
(v) If f E CAo(G) n Ll(G),
the Fourier transform of f.
q.
then t = f, where f denotes
(vi) The aapping from CAo(G) to Co (e) , defined by f - t,

f E CAo(G) , is the unique iinear mapping from CAo(G) to Co (G)
such that
o
(a) f = f, f E CAo(G) n Ll(G).
(b) IItii = IITII, f = fT ( CA (G).
GO 0
Proof. The proofs of parts (i) and (ii) are left to the reader.
If f = fT E CAo(G) , then there exists a sequence {f
n
) C Ll(G) n C(G)
such that lim IITf - Til = 0 and lim IIf - fll = O. From Lemma
n n n n CD
10.3.1 we see that T
f
* = (T
f
)*, and so, by Proposition 10.1.I(vi),
n n
and
limllTf* - T*II = liml/(T
f
)* - T*II
n n n n
= 1imllT
f
- Til
n n
= 0
limllf* - f*1I = limllf - fll = o.
n CD n CD
n n
Thus f* E CAo(G) and f* = f
T
*. Part (iv) is immediate from Theo-
rem 10.3.1.
To prove part (v) suppose f (CAo(G) n L
1
(G). Then, on setting
f = f, n = 1,2,3, . , we see at once that lim IITf - Tfll = 0 and
n n n
lim IIf - fll = 0, whence we conclude that f = fT. Hence, by
n n CD f
10.3. The B*-Algebra Ao(G)
289
Theorem 10.3.1, we have t = (T
f
)- = f.
o
Finally, it is now apparent that the mapping f - f, f ChOCG) ,
is linear and satisfies conditions (a) and (b) of part (vi). Suppose
f -1, f E CA (G) is another such mapping. If f = fT CA (G)
o 0
and (f) C Ll(G) n C(G) is such that lim IITf - Til = 0 and
n n n
lim IIf - fll = 0, then from parts (ii), (iv), and (v) we see that
n n CD
111' - til < 111 - 1 II + IIr - t II + II! - '"
CD- nCD n nCD n CD
= !Ir - T f II + liT f - Til
(n = 1,2,3, ),
n n
from which it follows that 1 = f.
lemma.
This completes the proof of the
o
In particular, the last part of the lemma says that the defini-
o
tion of f, f E CA (G),
o
is the only way we can extend the Fourier
transform on CA (G) n Ll(G) to all of CA (G). In the next sec-
o 0
tion we shall use this Fourier transform on a subspace of
CAo(G) (1 L
2
(G) to define an appropriate Fourier transform on all
of L
2
(G) and thereby obtain Plancherel's Theorem. Before this,
however, we need some rather technical results about CA (G).
o
Lemma 10.3.5. Let G be a locally compact Abelian topological
group. If f E CA (G) and , is real valued, then f(O) is real.
a
o -
Proof. Suppose f = fT. Then f = T, and, since the Gel'!and
transformation on A (G) is an *-isomorphism, we see that (T*) =
W Q q 0 0 _ _
T = f = ~ as f is real valued. Thus (T*) = T, and so T = T*
since A (G) is semisimple. If (f 1 C LlCG) n C(G) is such that
o n
lim IITf - Til = 0 and lim IIf - fll = 0, then we also have
n n n n CD
lim IIT* - Til = lim IITf* - T*II = O. Let g = Cf + f
n
*)/2, where,
n n n D n n
as usual, f *(t) = f (-t), t E G. Then (g] C Ll(G) n C(G), and
n n n
it is evident that lim liT - Til = 0 and lim IIg - (f + f*)/211 = o.
n gn nn CD
However, f = fT is uniquely determined by T, and so we conclude
that f = (f + f*)/2. Thus frO) = [frO) + f(0)]/2 is real.
o
290 10. B-Algebras
Lemma 10.3.S is also valid if we replace real by nonnegative
throughout its statement.
Lemma 10.3.6. Let G be a locally compact Abelian topological
o
group. If f E CAo(G) and fey) > 0, Y E G, then f(O) > o.
Proof. Let f = fr. From Lemma 10.3.S we know that f(O) is
real. Suppose f(O) < O. Then, since f is continuous, there
exists a symmetric open neighborhood U of 0 that has co,pact
closure for which If(t) - f(O)1 < If(0)1/2, t U. Recall that U
symmetric means that U = -U. Let g E Ll(G) n L
2
(G) be such that
(a) g * g* E Cc(G).
(b) The support of g * g* is contained in U.
(c) IG g * g*(s) dA(s) = 1.
Such an element g can be obtained in the following manner: Let W
be an open symmetric neighborhood of the identity in G such that
W + W C U and set gl = Xw' the characteristic function of W.
Then g = gl/[i
G
gl * gies) dl(s)]1/2 has the indicated properties.
We claim that f * g * g* CA (G). Indeed, from Proposition 4.7.2(i)
o
we see that f * g * g* E C(G). Furthermore, let (f
n
) C L
1
(G) n C(G)
be such that lim 1I1f - 111 = 0 and lim IIf - fll = 0, and set
n n n n CD
h
n
= fn * g * g*, n = 1,2,3, .. , and S = TTg * g*. Obviously
(hnl C LI(G) n C(G), while the estimates
- S II = liT - TT II
f * g * g* g * g*
n
< 1I1f - TlIlITg * g*1I
n
(n = 1,2,3, )
and
10.3. The B*-Algebra A (G)
o
show that f * g * g* E CA (G) and corresponds to S = TT *
o g * g
in Ao(G). Moreover, from Lemma lO.3.4(ii) and (v), we see that
o
(f * g * g*) (y) = (TT
g
* g*) (y)
. -
= T(y)(g * g*) (y)
02
= f(y)lg(y)1
291
> 0 (y E G).
Thus, in particular, from Lemma 10.3.5 we see that (f * g * g*)(O)
is real. Thus, on the one hand, the estimate
If * g * g*(O) - f(O)1 = IIG f(s)g * g*(-s) d1(s) - f(O)1
< Iu If(s) - f(O)lg * g*(-s) d1(s)
< If(O)l
- 2
reveals that f * g * g*(O) f(0)/2 < 0, as f * g * g*(O) is
real.
q.
But, on the other hand, try) = T(y) > 0, y E G, and TEe (G)
entail that (T)I/2 is a well-defined real-valued function in Co(G).
Since by Theorem 10.3.1 the Gel'fand transformation on Ao(G) is
surjective, we deduce that there exists some PEA (G) such that
1/2 0 1/2
P = (T) . Moreover, since A (G) is semisimple, (T) is real
o
valued, and the Gel'fand transformation is an *-isomorphism, we see
that P is self-adjoint and p2 = T. Now let (gn) C Ll(G) be such
that lim liT - pil = O. Since P is self-adjoint, we may assume,
n gn
without loss of generality, that g* = g , n = 1,2,3, ... , as was
n n
done in the proof of Lemma 10.3.5. Then for each n we have
292 10. B*-Algebras
whence we conclude at once that lim liT - T *11 = O. In parti-
n gn * gn
cular. we must also have lim IIT - T *11 = O. Finally then,
n n gn * gn
since g * g* (; C (G), g E L
2
(G), and lim IIf - fll = 0, we
c n n m
see that
f * g * g* (0) = iG g * g*(-s)f(s) dA(s)
=
lim iG g * g*(-s)fn(s)
dA(s)
n
=
lim f
* g * g*(O)
n
n
=
lim T
f
(g) * g*(O)
n n
=
lim fG T
f
(g)(s)g(s) dA(s)
n n
=
lim {T f (g),g)
n n
=
lim {T
g
* g*(g),g}
n n n
=
lim g * g* * g * g*(O)
n n n
=
lim (g * g) * (g * g)*(O)
n n
n
= lim fG
(g * g)(s)(g * g)*(-s) dA(s)
n n
n
= lim IG
(gn * g)(s)(gn * g)(s)
n
= lim (lig
n
* g1l2)2
n
= lim rliTg (g)1I
2
1
2
n n
= rll
p
(g)1I
2
1
2
> O.
dA(s)
This, however, contradicts the previous observation that
f * g * g*(O) < o.
Therefore f(O) > o.
[j
10.3. The B*-Algebra A (G)
o
We shall have need of one more result about CA (G).
o
But in
order to prove it we need to introduce a new topology on G and
establish a preliminary result.
293
Lemma 10.3.7. Let G be a locally compact Abelian topological

group. For each > 0 and each compact set KeG define
= {t 1 t E G, 1Ct,y) - 11 y (K).
Then the family of sets where > 0 is arbitrary and

KeG is an arbitrary compact set, forms a neighborhood base at the
identity of G for a topology that is weaker than the given topology
on G.
We shall see subsequently that the topology obtained as in the
lemma actually coincides with the given topology on G. The proof
of the lemma is left to the reader. The argument is essentially
the same mutatis mutandis as the one given in the first half of the
proof of Theorem 4.7.5, the crucial point being that the mapping

from G X G r, defined by (t,y) (t,y), t E G and y EGis
continuous.
Lemma 10.3.8. Let G be a locally compact Abelian topological

group. If KeG is compact, 6 > 0, and a > 0, then there
exists some g E C (G) such that
c
(i) get) > 0, t E G.


(ii) g(y) > 0, y E G.

(iii) Ig(y) - al < 6, y E K.
Proof. From Lemma 10.3.7 we see that N(K,6/a) is an open
neighborhood of the identity in G. Let U c N(K,6/a) be an open
neighborhood of the identity in G that has compact closure and
let W be a symmetric open neighborhood of the identity such that
W + We U. Let g = c(Xw * Xw), where c is so chosen that
294 10. B*-Algebras
IG g(s) dA(s) = a.
Since Xw * Xw is nonnegative, we see that c > O. Evidently
g E Cc(G) , the support of g is contained in W + W, get) ~ 0,
t E G, and

g(v) = c(Xw * Xw) (V)

= c(Xw * XW) (V)
2
= cIXw(v)1

>0 (V E G).
Finally, if y E K, then

19(y) - al = IIG g(s)(s,V) dA(S) - IG g{s) d1{s)1
~ I u g(s)l(s,v) - 11 dA(S)
~ sup I(s,v) - 11 IG g{s) dA(S)
s E N{K,6/a)
< 6.
A couple of additional remarks about this lemma are in order.
Clearly the function g is of the form g = f * f*, where
o
I- 12 f e Ll(G) n L
2
{G) n L.(G) and g(V) = fey) > 0, V E G. Moreover,
it is not difficult to see that we can even demand that f ( C (G).
c
Indeed, in the proof of the lemma we need only replace Xw by
Xwl * X
WI
' where WI is a symmetric open neighborhood of the iden-
tity such that WI + WI C W. We shall have need of these observa-
tions in the next lemma and in the next section.
Lemma 10.3.9. Let G be a locally compact Abelian topological
-
group. If h E C (G) is real valued, then for each e > 0 there
o c
exist h,k E CA (G) such that
o
(i) ~ ~ E C (G).
c
10.3. The B-Algebra A (G)
o
295
o 0
(ii) hand k are real valued.
o 0 ..
(iii) key) < h (y) < hey), y G.
- 0 -
(iv) 0 <h(O) - k(O) < e.
Proof. Let the compact support of h be K and let a = 2
o
and 6 = 1/2. By bemma 10.3.8 there exists some gEe (G) such
o .... 0 c
that g (t) > 0, t e G, g (y) > 0, y E G, and Ig (y) - 21 < 1/2,
o - 0.. 0
y E K. In particular, since g is real valued, we have
.. 0
go(y) > 3/2 > 1, y E K. Applying Lemma 10.3.8 again, this time
with a = I and 6 > 0 arbitrary, we obtain a function gEe (G)
c
such that get) > 0, t G, g(y) 0, y G, and 19(y) - 11 < 6,
..
Y E K. In particular, 1 - 6 < g(y) < 1 + 6, y E K. We shall specify
6 more precisely in a moment.
..
First, however, since h E C (G) and the Gel'fand transfor-
o c
mation on A (G) is a surjective isomorphism by Theorem 10.3.1, we
o ..
deduce the existence of a unique TEA (G) such that T = h. Let
o 0
Tl = TT 6 and T2 = TT . We claim that Tl and T2 deter-
g+ g
mine of CA (G) for any 6. To see this it clearly suffices
o
to show that TT and TT determine elements of CA (G). But
g go 0
suppose {f) c:: Ll(G) is such that lim IITf - Til = o. As observed
n n n
after Lemma 10.3.8, we may assume, without loss of generality, that
g = f f*, f E Ll(G) n L
2
(G) n Lm(G). Consequently (f
n
g)
belongs to Ll(G) n C(G) and, using HHlder's inequality, we find
from the estimates
IIf * g - f gil = lI(f - f ) f * f*1I
n m m n m m
< lI(f
n
- fm) f1l211f*1I2
< IITf - T
f
II Uf1l211f*1I2 (n,m = 1,2,3, ... )
n m
and
IITf g - 11 II < IITf - TIIIIT II
* g - g
n n
(n = 1,2, 3, ... )
296 10. B*-Algebras
that lim IITf - 17 II = 0 and that (f * g) cenverges uniform-
n n * g g n
Iy to some function in C(G). Hence TT determines an element of
g
CAo(G). A similar argument shows that the same is true of TT .
go
Thus TT 6 and TT 6 determine elements of CA (G),
g+ go g- go 9 o.
which we denote by hand k, respectively. Clearly h = T(T +6 )
9 0 g go
= h (g + 6g) and k = h (g - 6g), whence we see that hand
000 00.
k are real-valued functions in C (G). Moreover, if y K, then
c
since g (y) > I and g(y) + 6 > 1, we have
o
o
hey) = h (y)[g(y) + 6g (y)]
o 0
o 0
whereas if y ~ K, then h(y) = 0 = h (y). Thus h(y) > h (y),
0 .0 - 0
y (G. Similarly k(y) < h (y), y E G. Hence parts (i) through (iii)
- 0
of the lemma are established for any 6 > O. However, to obtain
part (iv) we must put some additional restrictions on 6.
Suppose g = f * f*, f E Ll(G) n L (G). Then, by HHlder's
o 000 m
inequality,
Ifn * go(O)l
=
If * f * f ~ O ) 1
n
0
=
ITf (fo>
* f*(O)1
0
n
= lIG T
f
(fo)(s)o(s) dA(s)1
n
< IITf (fo)1I2l\
f
oll2
n
(n = 1,2,3, ).
Since (I\T
f
III is a bounded sequence because lim UTf - Til = 0,
n n n
we conclude that {f * g (0) is a bounded sequence. Set
n 0
x = sup If * g (0)1 and suppose 0 < 6 < e/6x. Now from the proof
n n 0
that TT 6 and TT 6 determine elements of CAoCG) it is
g+ go g- go
immediately apparent that
10.4. Plancherel's Theorem 297
limllf * (g + 6g ) - hI! = 0
n 0 CD
n
and
limllf * (g - 6g ) - kll = o.
n 0 Q)
n
Thus there exists some positive integer n
o
such that, if
then
and
Therefore, for
IIf * (g + 6g ) - hll < 3
n 0 Q)
JJ
f * (g - 6 g ) - k II < 3
n 0 CD
n > n ,
- 0
we have
I h(O) - k{O)1 < Ih(O) - f
*
{g + 6g
o
)(0)1
n
n>n,
- 0
+ If *
(g + 6g ) CO) - f
*
(g - 6g ) (0) I
n 0 n 0
+ If *
n
(g - 6g ) (0)
0
- k(O)1
2
<-+
3
261f *
n
go (0) I
2
26H. <-+
- 3
< .
00.
However, since hey) - key) > 0, y E G, we see from Lemma 10.3.6
that h(O) - k(O) 0, whence 0 <h(O) - k{O) < e.
c
10.4. Plancherel's Theorem. In this section we shall apply
the results developed in the preceding section to prove Plancherel's
Theorem. The first step in this direction is an inversion theorem
that is of considerable interest in its own right since it allows
us to recover certain f E CA (G) from a knowledge of t.
o
298 10. B*-Algebras
Theorem 10.4.1 (Inversion Theorem). Let G be a locally com-
pact Abelian topological group and let A be a given Haar measure

on G. Then a Haar measure on G can be so chosen that, if
f E CAo(G) and t E Cc(G), then
f(t) = fa ley) (t,y) (t E G).
Proof. The crux of the proof is to construct an appropriate

translation invariant nonzero positive linear functional on Cc(G)
that will yield the Haar measure we require. By such a functional

F we mean, of course, a linear mapping from C (G) to C that is
c
not identically zero and such that F(f) > 0 whenever f C (G)
- c.
and f(V).> 0, V E G, and such that = F(f), f E Cc(G)
and V E G. With this in mind, let g E C (G) be real valued.
c
Then from Lemma 10.3.9 we see that
sup{k(O) , k E CA (G), < g(y), V E Gl
o -
= inf{h(O) , h E CAo(G), > g(V), V E Gl.
We denote this common vAlue by F(g). In this way we obtain a map-
R R
ping F: C (G) where C (G) denotes the space of real-valued
c c
continuous functions on G with compact support. It is easily veri-
fied from the definition of F that F is linear and F(g) > 0
R -
whenever g E C (G) and g(V) > 0, V E G. Moreover, if f E CA (G)
oR. c - 0
and f E Cc(G) , then, since in this instance we can take h = k = f
in Lemma 10.3.9, we see that F(!) = f(O). We can extend F to

all of C (G) in the obvious manner: if gEe (G), then
c c
g = Re(g) + iIm(g), and we set F(g) = F[Re(g)] + iF[Im(g)]. Hence
F is a positive linear functional on C (G) such that FC!) = f(O)
c
whenever f E CA (G) and tEe (G). Furthermore, we claim that
o c
F is not identically zero.
Indeed, by Theorem 10.3.1, there exists some T E AO(G) , T 0,

such that TEe (G). Let gEe (G) C L
2
(G) be such that
c c
IIT(g)1I2 > 0 and set f = T(g) * T(g)*, where, as usual, T(g)*(t) =
10.4. Plancherel's Theorem
299
T(g)(-t). From Proposition 4.7.2(ii) we see that fEe (G). Actu-
o
ally, f E CA (G). To see this, let {f) c L
1
(G) be such that
o n
lim IITf - Til = O. Moreover, lim IIT* - T*II = 0 by Lemma 10.3.1.
n n n n
Now (f * f* * g * g*) c L
1
(G) n erG) and for n = 1,2,3, ...
n n
IIf * f* * g * g* - T(g) * T(g)*11
n n m
= lI(f * g) * (f * g)* - T(g) * T(g)*11
n n m
= IITf (g) * T
f
(g)* - T(g) * T(g)*lI
co
n n
< IITf (g) * T
f
(g)* - T
f
(g) * T(g)*lI
co
n n n
+ IITf (g) * T(g)* - T{g) * T{g)*lI
co
n
:s. IITf (g)* - T{g)*1I
2
I1T
f
(g)1I
2
n n
+ IIT(g)*1I
2
i1T
f
(g) - T(g)1I
2
n
= IITf (g) - T(g)1I
2
[IIT
f
(g)1I
2
+ IIT(g)1\2]'
n n
Hence we conclude that
limllf * f* * g * g* - T(g) * T(g)*11 = O.
n n n m
Furthermore, for n = 1,2,3, ...
IITf * f* * g * g*
n n
whence we have
limllT
f
* f* * g * g* - TT*T
g
* g*1I = 0
n n n
by the continuity of multiplication in a Banach algebra. Consequently
f = T(g) * T(g)* CA (G), and f corresponds to TT*T * E A (G).
q ..0. g * g 0
However, f = T(T*) g(g*) E C (G), and so
c
F(t) = f(O) = T(g) * T(g)*(O)
300
= fG T(g) (s)T(g)(s) dA(s)
= [IIT(g)1I
2
]2
> O.
10. B*-Algebras
Hence F is not identically zero.
Finally, we claim that F is translation invariant. Indeed,
by the same sort of argument as we used several times previously,
it is not difficult to show that, if k (CA (G), then (,w)k ( CA (G),
0 0 0 0
w E G, and [(,w)k] = T (k). The details are left to the reader.
-w
Thus, since k(O) = [(.,-w)k](O), w G, we deduce that, given
R
g (C (G), we have for each w G
c
F [T (g)]
=
sup{k(O) IkE CA (G), < T (g)(y), Y E G)
w o - w
sup{k (0)
I
o
=
k (CA (G), T (k)(y) < g(y), y E G}
o -w -
sup{k (0)
I
o
=
k (CA (G), [(.,-w)k] (y) < g(y), y E G)
o -
sup{k (0)
I
o
= k ( CA (G), k (y) < g (y), y E G)
o -
= F(g).
It is now apparent that F is translation invariant.
An appeal to the Riesz Representation Theorem for positive lin-
ear functionals [HR
1
, p. 129] shows that there exists a unique posi-

tive regular Borel measure that is, a Haar G such
that
F(g) = fa g(y)

(g E C (G)).
c
o
Consequently, if f E CA (G) and f E C (G),
o c
then
However, if t E G
f(O) = fa t(y)
and f ( CA (G), then an easy argument reveals
o
10.4. Plancherel's Theorem
301
00.
that T_tCf) E CAo(G) and that [T_tCf)] (y) = (t,y)f(y), y ( G,
o
from which we deduce that, if f E CAoCG) and f E CcCG), then
o
f(t) = T_t(f) (0) = fa [T_t(f)] (y) d ~ y )
o
= fa f(y) (t,y) dftCy) (t E G).
This completes the proof of the inversion theorem.
o
The following corollary is now immediate since, by Lemma 10.3.4
o
(v) ,
f = f whenever f E CAo(G) n Ll(G).
Corollary 10.4.1. Let G be a locally compact Abelian topolo-
gical group and let A be a given Haar measure on G. Then a Haar

G can be so chosen that, if f E CAo(G) n LI(G) measure ~ on

and f E C (G),
c
then

fCt) = fa f(y) (t,y) d ~ y )
(t ( G).
Note that in the statement of the corollary we are tacitly as-
suming that f is a continuous function; that is, we are consider-
ing f as an element of C(G). If f is thought of as an element
of LICG), that is, as an equivalence class, then the formula in
the corollary holds only almost everywhere.
Utilizing this inversion theorem, we can now prove Plancherel's
Theorem.
Theorem 10.4.2 (Plancherel's Theorem). Let G be a locally
compact Abelian topological group and let A be a given Haar measure
on G. If
v = (f I f E CA (G) n L
2
(G) , , ( C (G)],
o c

then a Haar measure ~ on G can be so chosen that
(i) V is a norm-dense linear subspace of L
2
(G).
302 10. B-Algebras
o 0
(ii) V = (f , f E V] is a norm-dense linear subspace of L
2
(G).
(iii) IIfll2 = 11'11
2
, f E V.
(iv) The mapping f ~ I, f E V, from V to ~ can be uniquely

extended to a linear isometry of L
2
{G) onto L
2
{G).
o
Proof. Obviously V and V are linear subspaces of L
2
(G)

and L
2
{G), respectively. To see that V is norm dense in L
2
{G)
let g E L
2
(G) and I > O.
Then there exists some f E Cc{G) such that IIf - gll2 < 1/2.
Moreover, we ~ l a i m that we may assume without loss of generality
that f = fl * f2 * f3' fk E Cc(G), k = 1,2,3. Indeed, from the
proof of Theorem 4.7.2, if (u 1 is a family of symmetric open
a
neighborhoods of the identity in G that have compact closures and
are such that n U = {oj, then we know that {u] = ~ /A(U )1
(t a a 'U
a
(t
is an approximate identity for LICG). It is easily seen that, if
f E C (G), then lim 'If - f * u II = 0 and that (u * u 1 c: C (G)
c a' (tCII a a c
is also an approximate identity. After combining these remarks it
becomes apparent that we can assume f has the indicated form. We
leave the details to the reader.

Now since fl E Co(G) and, by Theorem 10.3.1, the Gel'fand
transformation on A (G) is a surjective isometry, we see that
o
there exists some TEA (G) such that T E C (G) and
o c
Thus
liT - fIll = liT - T f II < 211f : f II .
CII 1 2 3 2
c
<-
2
c
<-
-2
c
<-
2
+ IITf (f2 * f3) - T(f
2
* f
3
)11
2
1
+ IITf - Tllllf2 * f3112
1
c
+ - = c.
2
10.4. Plancherel's Theorem
303
However, T(f
2
* f3) E V. Clearly T(f
2
* f3) E L
2
(G), and if
{h ) C Ll(G) is such that lim 11Th - Til = 0, then it follows that
n n n
lim 11Th f f - IT
f
f II = O. But the estimates
n n * 2 * 3 2 * 3
IIh
n
* f2 * f3 - h
m
* f2 * f311CD< IIh
n
* f2 - h
m
* f211211f3112
= 11Th (f
2
) - Th (f2)1I2
I1f
311 2
n m
~ 11Th - Th IIl1f2112I1f3112'
n m
valid for n,m = 1,2,3, , reveal that (h
n
* f2 * f31 is a Cauchy
sequence in C(G). Thus there exists some h E C(G) such that
limnllhn * f2 * f3 - hU
CD
= O. However, we also have
Inimllhn * f2 * f3 - T(f
2
* f
3
)11
2
= limliTh (f2 * f3) - T(f
2
* f
3
)11
2
n n
= 0,
as limnUTh - Til = O. Consequently, by a result of integration
theory [ S l ~ pp. 122 and ISO], we deduce that (h
n
* f2 * f31 has
a subsequence that converges almost everywhere to T(f
2
* f
3
).
Hence T(f
2
* f3) = h almost everywhere, and so we may assume that
T(f
2
* f3) E L
2
(G) is the continuous function h. Thus T(f
2
* f3)
belongs to CAo(G) n L
2
(G), as (h
n
* f2 * f31 C Ll(G) n C(G).
Moreover, it is apparent from the preceding argument that T(f
2
* f3)
o --
corresponds to TT
f
f E A (G), and so T(f
2
* f3) = TT
f
* f
_ 2* 3 0 23
belongs to Cc(G). Therefore T(f
2
* f3) ~ V and V is norm dense
in L
2
(G).
0--
To see that V is norm dense in L
2
(G) we let g E L
2
(G) and
- -
e > O. Since Cc(G) is dense in L
2
(G) and the Gel'fand transfor-
mation on A (G) is surjective, we see that there exists some o _ _ _
TEA (G) such that TEe (G) and lIg - TII2 < e/2. Let the
o _ c
compact support of T be K and apply Lemma 10.3.8 with a = 1
and 6 = e/[2l1TII 1)(K)1/2] to deduce the existence of some fEe (G:
_ CD _ 1/2 c
for which I fey) - Il < el [211TIl 1)(K) ], y E K. The Haar measure
- CD
1) on G can, at this point, be any such measure. As indicated in
304 10. B*-Algebras
the remarks following Lemma 10.3.8, we may assume that f = f * f*,
o 0
where f (C (G). Repeating the same argument as before, mutatis
o c 0 __
mutandis, we see that T(f) = TCf * f*) E V and TCf) = Tf. Con-
o 0
sequently
liT - TfU2 = [fa IT(y)1
2
cll - fcy)12) dll(y)]1/2
= [f
K
IT(y)12(ll - f(y)12) dll(y)]1/2
~ II
r
ll .. ll(K)I/2[ 11.11 c 1/21
2 T 1}(K)
co
whence
e
= 2'
o -
e e
<-+-
2 2
= e.
Thus V is norm dense in L
2
(G).
- Next we choose Haar measure r. on G so that Theorem 10.4.1
is valid. If f V, then f. f* ( CA CG) and (f * f*)o =
o 0 ~ _ 0
f(f) = ff E C (0). Hence, by the Inversion Theorem, we see that
c
= fG f(s)*(-s) dA(s)
= f * f*(O)
o
= fa (f * f*) (y) d1}(y)
= fa !(y)JCY) dll(y)
= (11,"
2
)2,
o
from which it follows at once that the mapping f - f, f (V, is a
10.4. Plancherel's Theorem
305
o
linear isometry from V onto V.
Thus parts (i) through (iii) are proved, and the proof of part
(iv) is now apparent.
o
It should be remarked that part (ii) of Plancherel's Theorem
follows immediately from parts (i) and (iii).
the direct proof of part (ii) to emphasize the
Nevertheless we gave
fact that the density
o ..
of V in L
2
(G) is independent of the choice of Haar measure on
In accordance with the version of Plancherel's Theorem intro-
duced in Section 7.2 we shall call the unique extension of the map-
o
-
G.
ping f - f, f (V, to all of L
2
CG) the Plancherel transformation
and denote, for the moment, the Plancherel transform of f E L
2
(G)
o
by f. The same proof as that given for Corollary 7.2.1 establishes
Parseval's Formula.
Corollary 10.4.2 (Parseval's Formula). Let G be a locally
compact Abelian topological group and let A be a given Haar measure
..
on G. If Haar measure on G is so chosen that Plancherel's
Theorem is valid, then
Several other useful properties of the Plancherel transformation
are easily proved. We shall prove only the last one stated in the
next corollary. The others are left to the reader.
Corollary 10.4.3. Let G be a locally compact Abelian topolo-
gical group, let A be a given Haar measure on G, and let be
..
a Haar measure on G so chosen that Plancherel's Theorem is valid .
..
If f E L
2
(G), then for almost all y E G
0
Ci)
_ 0
fCy) = fC-y)
Cii)
o 0
Cf*) Cy) = fCy)
Ciii) [C,w)f]oCY) = tCY - w), w E G.
306
10. B*-Algebras
Moreover, if f,g E L
2
(G), then (fg)-(y) = t * g(y), where y E G
- and (fg) denotes the usual Fourier transform.
Proof. If f,g E L
2
(G), then evidently fg E LI(G). Hence
from Parseval's Formula and parts (i) and (iii) of this corollary
we see that
"
(fg) (y) = IG f(t)g(t) (t,y) dAft)
= fG f(t)iCtf(t,y) dAft)
o I
= fa f(w)g(w - y) d ~ w )
= Sa ~ w ) g y - w) d ~ w )
o 0
= f * g(y)
-
(y E G).o
The present version of Plancherel's Theorem is not quite the
same as the one introduced in Section 7.2 since V need not be a
subset of Ll(G) n L
2
(G). However, the next theorem combined with
Theorem 10.4.2 immediately yields the previous form of Plancherel's
"
Theorem (Theorem 7.2.1). Here f will denote the usual Fourier
transform.
Theorem 10.4.3. Let G be a locally compact Abelian topolo-
gical group, let A be a given Haar measure on G, and let \ be
- a Haar measure on G so chosen that Plancherelts Theorem is valid.
If
then
(i) Vo is a norm-dense linear subspace of L
2
(G).
" "
(ii) Vo is a norm-dense linear subspace of L
2
(G).
o "
Proof. We note first, by Lemma 10.3.4(v), that f = f, f V .
o
To prove part (i) it suffices to show that V
2
= (gh I g,h VJ is
10.4. Plancherel's Theorem
307
norm dense in L
2
CG). This is so because, if g,h E V, then
o 0 ..
I.h E CAoCG) n L
2
(G) and g.h E Cc(G). whence. by Lemma 10.3.4(ii)
and (v) and Corollary 10.4.3, we see that gb E CCG) n LICG) n L
2
(G)
o .. 0,.. 2
C Cho(G) n L
2
(G) and (gb) = (gb) = g * h E C (G). Thus V c V .
o 0 2
c
o.. 0
Moreover, since it is apparent that V * V = (V) C V and since
o
tbe Plancberel transformation is an isometry, we see that we can
prove botb parts (i) and (ii) by showing tbat * is norm dense
..
in L
2
(G).
..
So suppose f E L
2
(G) and let e > O. Then, as noted in the
..
proof of Theorem 10.4.2, there exist g,b E C (G) such that
c 0
IIf - g * hll2 < e/3. Furthermore, by Plancberel's Theorem, V is
0
dense in L
2
(G) , and so there exist gl,h
l
E V such that
IIg - gl112 < e/ 3(Uhll
l
+ 1) and lib - hll12 < '/3(lIg
1
I1
l
+ 1). Combin-
ing tbese estimates, we see tbat
IIf - gl * hlll2 IIf - g * hll2 + IIg * h - gl * hil
2
+ IIg
l
* b - gl * hlll2
< ; + IIg - g111211hlll + 1I1111111h - b
l
ll
2
C I I
<-+-+-
333
= I,
00"
whence we conclude that V * V is norm dense in L
2
(G). Thus parts
(i) and (ii) of the theorem are proved.
We restate the previous form of Plancherel's Theorem for the
sake of easy comparison.
o
Theorem 10.4.4 (Plancherel's Theorem). Let G be a locally
compact Abelian topological group and let A be a given Haar measure

on G. Then there exist a Haar measure on G and a linear sub-
space Vo of L
2
(G) such that
308 10. B*-Algebras
(ii) V is norm dense in
L2 (G)
0

(iii) V is nonn dense in
L2 (G) .
0

(iv)
IIfll2 = II
f
1l
2
,
f E V
0

(v) The mapping
f -. f, f E V , from V to V can be
0 0 0

uniquely extended to a linear isometry of L
2
(G) onto
L2 (G).
Naturally, the Haar measure is so chosen that Theorem 10.4.2
is valid and Vo = V n Ll(G). Clearly the extension of the Fourier
transformation given by this theorem coincides with the one given
by Theorem 10.4.2. Actually it is possible to prove that f = t
for any f E LI(G) n L
2
(G) and that one can take Vo = LICG) n L
2
(G)
in the preceding theorem. We shall not, however, take the time to
do this.
As in Section 7.2, it is standard practice to denote the Plan-
cherel transform of f by i, rather than t. For obvious reasons
it was not desirable to do this for the greater part of this section,
but we shall utilize this notation for the remainder of this volume.
There are other proofs of Plancherel's Theorem that do not make
use of B*-algebras or the Commutative Gel'fand-Naimark Theorem. These
proofs generally utilize some sort of inversion theorem and Bochner'S
Theorem on positive definite functions (see, for example, [N, pp.
409 -415; Ru
l
, pp. 17-27]). Other approaches are discussed, for
instance, in [HR
2
, pp. 324-326; We, pp. 111-115]. We chose the pre-
sent approach to Plancherel's Theorem since it is closer to the gen-
eral theme of this volume than other methods. Other discussions of
this approach, originated by J.H. Williamson, can be found in his
original paper [WI] and in [Hy, pp. 148-163].
10.5. The Pontryagin Duality Theorem. We have observed that,

if G is a locally compact Abelian topological group and G
10.5. The Pontryagin Duality Theorem 309
is the set of continuous homomorphisms of G into the unit circle
- r in the complex plane, then G can be provided with algebraic
operations and such a topology that it becomes a locally compact
Abelian topological group, called the dual group of G. Moreover,
we have given certain examples that suggest that the dual group of
- -.
G, denoted by (G), is topologically isomorphic to G. The main
purpose of this section is to prove this assertion, generally
as the Pontryagin Duality Theorem. The results referred to were
discussed in Section 4.7.
The first step in proving the indicated theorem is to observe
-- the existence of a natural isomorphism of G into (G). Indeed,
- given t (G, we define a(t)(y) = (t,y), y (G. It is easily veri-
-
fied that aCt) is a homomorphism of G into f. The continuity
of aCt) follows from the fact that, given e > 0, the set
- U({t),e) = {y lyE G, l(t,y) - 11 <.) is an open neighborhood of
. . -
the identity in G, by Theorem 4.7.5. The mapping a: G - (G)
so defined is clearly a homomorphism. The injectivity of a follows
at once from Corollary 4.7.3, while an elementary argument utilizing
Theorem 4.7.5, Lemma 10.3.7, and the simple observation that for
- each e > 0 and each compact set KeG we have
N(K,e) = (t I t E G, I(t,y) - 11 < s, y K)
= I x ( (G)-, lx(y) - II < s, y ( K)]
reveals that a is continuous. In view of this fact and Lemma 10.3.7
it is apparent that to prove a is a homeomorphism we need only
show that the sets N(K,s) in G form a neighborhood base at the
identity for the topology of G. This strengthening of Lemma 10.3.7
is the second preliminary step to proving the Pontryagin Duality
Theorem.
Theorem 10.5.1. Let G be a locally compact Abelian topologi-
- cal group. For each s > 0 and each compact set KeG define
N(K,s) = {t I t E C, I(t,y) - 11 < S, y E K).
310 10. B-Algebras
Then the family of sets (N{K,e)}, where e > 0 is arbitrary and
..
KeG is an arbitrary compact set, forms a neighborhood base at the
identity of G for the topology of G.
Proof. In view of Lemma 10.3.7 we need only show that, if U
is an open neighborhood of the identity in G, then there exist
..
some e > 0 and some compact set KeG such that N{K,) c U.
To this end let WI eWe G be symmetric open neighborhoods of the
identity with compact closure such that W + W c U and cl{W
l
+ WI)
c W. Let g = c{Xw
I
Xw
1
), where Xwl denotes the characteristic
function of WI and c > 0 is so chosen that IIg1l
2
:: 1. Then it
is readily verified that g E C (G), get) > 0, t G, and the sup-
c -
port of g is contained in W. Furthermore, if s U, then the
supports of g and T (g) are disjoint, whence
s
= 2 > 1.
Consequently, if 5 E G and IIg - T
s
{g)1I
2
1, then s E U.
Next, recalling the proof of Theorem 4.7.2 and the comments
following it, we choose a nonnegative u LI{G), lIulll = 1, such
that I/u. g - gll2 < 1/3. Since the norm in L
2
(G) is translation
invariant, we have
lIu Ts(g) - T
s
(g)1I
2
:: UTs(u g) - T
s
(g)1I
2
= lIu g - g/l2
I
<3
(s E G),
whence we see that, if s E G and /lu. g - u T
s
(g)/l2 < 1/3,
then
IIg - T
s
(g)/l2 < /lg - U gll2 + /lu g - u T
s
(g)/l2
+ lIu Ts(g) - T
s
(g)/l2
10.5. The Pontryagin Duality Theorem
III
<-+-+-
333
= 1.
311
Hence, if s ~ G and lIu. g - u T
s
(g)1I
2
~ 1/3, then s E U.
Moreover, since u ~ Ll(G) can be considered as an element of the
B-algebra A (G), namely, u is identified with T, and the
o u
Gel'fand transform of T and the Fourier transform of u agree,
u
we see from Theorem 10.3.1 that, if s E G were such that
.. .. ..
luCy) - T (u) (y)l < 1/3, y E G, then
s
lIu g - u T
s
(g)1I
2
= lIu g - Ts(u) gll2
= lI[u - Ts(u)] gll2
< lIu - T
s
(u)lIllg11
2
.. ..
= lIu - T (u) II
s CD
Thus, by the previous observation, if s EGis such that
.. .. ..
luCy) - T (u) (y)1 < 1/3, y E G, then s E U.
s
. .. ..
Now U ( C (G), and so there exists some compact set KeG
.. 0 ..
such that luCy)l < 1/6, y E G - K. Then, if s E N(K,1/3), we
see that for y E K we have
. .. .. ..
luCY) - Ts(u) (y)1 = luCy) - (s,y)u(y)I
= 1 ~ Y ) l l l - (s,y)l
< II - (s,y)1
1
< 3'
..
as lIu1l1 = 1; and if y E 6 - K, then
.. .. -
luCy) - TsCu) (y)' = lu(Y)"l - (s,y)l
312 10. B-Algebras

as lu(y)1 < 1/6, y E G - K.
1
< 3'
Therefore, if s E N(K,1/3), then

y (G, whence we conclude that s E U,
which is what we set out to prove.
A
lu(y) - Ts(u) (y)1 < 1/3,
that is, N(K,I/3) c U,
As already indicated, this theorem entails that the mapping
o
A A
a : G - (G) is a topological isomorphism of G onto a(G) C (G)
It is now not too difficult to show that a is surjective.
Theorem 10.S.2 (Pontryagin Duality Theorem). Let G be a lo-
A
cally compact Abelian topological group and let (G) denote the
A A
dual group of G. If a: G - (G) is defined for each t E G by

a(t)(y) = (t,y), y E G, then a is a topological isomorphism of

G onto (G).
Proof. In view of the foregoing discussion it is sufficient
A
to prove that a(G) = (G). To do this we first show that a(G) is
A
a closed subset of (G). Indeed, let K E cl[a(G)] and let
be a net in a(G) that converges to K. Since a is a
topological isomorphism, we see, given a compact set U C G that
has a nonempty interior containing the identity of G, that there
exists some such that - tw E U, > and w In
particular, - E U, a and so a(t
p
) ( a(t
po
) + a(U) =
a(tao.+.U)' a >Po. However, a(t
po
+ U) is compact, and so closed,
in (G), as a is a homeomorphis=". Hence k E a(t
po
+ U) C a(G),
and a(G) is a closed subset of (G) "

Now, if a(G) (G), then a(G) is a proper closed subset of
A
(G) , whence, since LI(G! is a regular commuta!ive Banach algebra,
there exists.some h E Ll(G), h 0, such that h[a(t)] = 0, t G,
the symbol h here of course denoting the Fourier transform of h
A
as an element of Considering LI(G) as a subspace of M(G),
we deduce at once from the Riesz Representation Theorem for continuous
10.5. The Pontryagin Duality Theorem
313

linear functionals on C (G) [L, p. 109] and Theorem 10.3.1 the exis-
o
tence of some TEA (G) such that
o

ja h(y)T(y) o.
Furthermore, a second appeal to Theorem 10.3.1, the definition of
Ao(G), and the fact that Cc(G) is norm dense in Ll(G) allows
us to assert the existence of some f ( C (G) such that
c

Ia h(y)f(y) o.
However, by Fubini's Theorem [Ry, p. 269], we see that

Ia h(y)f(y) = Sa h(y)[I
G
f(t) (t,y) dA(t)]
= IG f(t)[ja h(y)Ct,y) dl(t)
= IG f(t)[ja h(y) (a(t),y) d4(t)

= IG f(t)h[aCt)] dl(t)
= 0,
which contradicts the choice of f.

Therefore aCG) = (G), and the theorem is proved.
o
It may not be apparent to the reader where we have used Plan-
cherel's Theorem (Theorem 10.4.4) to prove the preceding result. The

use occurs in appealing to the regularity of LlCG) (Corollary 7.2.3),
a fact whose validity was established by means of Plancherel's Theo-
rem.
In view of the Pontryagin Duality Theorem it is customary to
identify a locally compact Abelian topological group G with its

second dual group (G). It is this identification that was alluded

to in the discussion of and L in Section 4.7.
A number of interesting results are now easy consequences of the
314 10. B*-Algebras
Pontryagin Duality Theorem. The first of these establishes the con-
verse of Corollary 4.7.5 and is an immediate consequence of it and
the Pontryagin Duality Theorem.
Corollary 10.5.1. Let G be a locally compact Abelian topolo-
gical group. Then the following are equivalent:
(1) G is compact (discrete)

(ii) G is discrete (compact).
If G is a locally compact Abelian topological group and
h E L
2
(G), then we shall denote by ~ the element of L
2
(G) de-
fined almost everywhere by ~ t ) = h(-t), t E G.
Corollary 10.5.2. Let G be a locally compact Abelian topolo-
gical group, let A be a given Haar measure on G, and let ~ be

a Haar measure on G so chosen that Plancherel's Theorem is valid.
A _.. ..
If f E L
2
(G), then [(f)] c f, where denotes the Plancherel

transformation on L
2
(G) and L
2
(G), respectively.
Proof. Since the Plancherel transformation is an isometry, it
suffices to prove the result for f E Vo = V n LI(G) because, by
Theorem 10.4.3, Vo is norm dense in L
2
(G). However, on Vo the
Plancherel and Fourier transforms coincide and the Inversion Theorem
(Theorem 10.4.1) is valid; so that, if f E V, then, on identify-
o
.. A
ing G with (G) , we see that
Thus
A_A
[(f) ] 'Ct) = la f(-y) (t,y) d ~ y )
A
= la fey) (t,y) d ~ y )
= fet)
A_A
[(f) ]1 = f, f E V, and the corollary is proved.
o
(t E G).
o
In particular. we observe that, if G is a locally compact
Abelian topological group. A is a given Haar measure on G, and
10.5. The Pontryagin Duality Theorem
315
..
is Haar measure on G so chosen that Plancherel's Theorem is

valid for L
2
(G) , then Plancherel's Theorem also holds for L
2
(G)
.. .
with the Haar measure on (6) taken to be A. This makes sense
. ..
as we can identify (6) with G. Moreover, the inverse of the
..
Plancherel transformation from L
2
(G) to L
2
(G) is the transfor-
.. r.! .. ..
mation .from L
2
(6) to L
2
(G) defined by g a 19) , g L
2
(6) ,
where denotes the Plancherel transformation from L
2
(G) to
L
2
(G).
Our final corollary entails that M(G), G being a locally
compact Abelian topological group, is semisimple.
Corollary 10.5.3. Let G be a locally compact Abelian topolo-
gical group. If E M(G) and

= fG (t,y) = 0

(y E G),
then = o.
.. .. .
Proof. Identifying (G) with G, we see that, if f E LI(G) ,
then f E C (G) and, by Fubini's Theorem [Ry, p. 269], we have
o
..
fG f(t) = fG[ia f(y) (t,y)
= ia f(y)[jG (t,y)

= ia
= o .
..
However, since L (G) is a self-adjoint commutative Banach algebra,
we know that L
I
(!)- is norm dense in Co(G) (see. for example. the
discussion following Definition 5.1.1). Consequently from the
Riesz Representation Theorem for continuous linear functionals on
CoCG) [L, p. 109] we deduce that = o. 0
The semisimplicity of MeG) is evident from this result because
..
each y E G determines a ca.plex hOllOllOrphism of MeG). However,
316 10. B*-Algebras
as indicated in Section 4.8, the maximal ideal space of M(G) in gen-

eral properly contains C.
10.6. A Spectral Decomposition Theorem for Self-adjoint Con-
tinuous Linear Transformations. A central concern in the theory of
continuous linear transformations on Hilbert spaces is the develop-
ment of ways of expressing a given transformation in terms of a sum
or integral involving a family of orthogonal projections indexed ~ y
the points in the spectrum of the transformation. A variety of such
results, known generally as spectral decomposition theorems, is
known (see, for example, [BaNr, pp. 433-521; DS
2
, pp. 887-902,1141-
1222; He, pp. 251-258,272-287; L, pp. 462-470; N, pp. 248-253; RzNg,
pp. 233 and 234,272-277,280-291,313-320]). In this section we shall
apply the Commutative Gel'fand-Naimark Theorem to obtain such a theo-
rem for self-adjoint continuous linear transformations.
Let V = (V,<.,.}) be a Hilbert space over ~ and suppose
T E L(V) is self-adjoint, that is, T* = T. We shall denote by Ar
the closed commutative subalgebra of L(V) generated by T. In
other words, ~ is the smallest closed subalgebra of LeV) that
contains T and the identity transformation I. Since T is self-
adjoint, it is evident that Ar is a commutative B*-algebra with
identity, and so from the Commutative Gel'fand-Naimark Theorem (Theo-
rem 10.2.1) we see that the Gel'fand transformation on Ar is an
isometric *-isomorphism from ~ onto C(a(Ar)). Moreover, from
our discussion of finitely generated algebras in Section 4.5 we know
that a ~ ) is homeomorppic to the joint spectrum of the generators
of ~ (Theorem 4.5.1), whence we conclude that a CAr) is homeo-
morphic to aATCT). However, aAT(T) = aL(V) (T). Indeed, this fol-
lows at once from Corollary 2.3.1 because aAT(T) Cffi by Proposition
10.1.1(viii). Thus afAr) is homeomorphic to aLcY)(T) = aft).
Combining this discussion with Theorem 4.5.1, we immediately obtain
the following lemma:
10.6. A Spectral Decomposition Theorem
317
Lemma 10.6.1. Let V be a Hilbert space over C and let
T E L(V) be self-adjoint. If is the closed subalgebra of
L(V) generated by T, then the Gel'fand transformation on Ar is
an isometric *-isomorphism from onto C(a(T)). Moreover,
-
TeT) = T, T E aCT).
Now, as indicated above, we wish to show that T and even cer-
tain functions of T can be approximated by sums of orthogonal pro-
jections that is, by sums of self-adjoint P (L(V) for which
2
P = P. In order to accomplish this it is tempting to approximate

T in C(a(T)) by some suitably nice functions and then use the
isometric nature of the Gel'fand to obtain elements
of that approximate T. However, the elements of Ay so ob-
tained need not be orthogonal projections, and, more to the point,
Ay need not contain any orthogonal projections at all. For instance,
if P E.'r were an orthogonal projection, then, since p2 = P, we
-2 -
would have P = P, and so P would take only the values zero and

one. Thus the set (T 1 T E aCT), peT) = 1) and the set

(T 1 T E aCT), peT) = 0) would be, in general, nonempty disjoint
open subsets of aCT), and aCT) would be disconnected. Consequent-
ly, if aCT) is connected, then Ar contains no nontrivial ortho-
gonal projections. Nevertheless, the original idea will work if we
replace C(a(T)) by a somewhat larger algebra and extend the inverse
of the Gel'fand transformation from C(a(T)) to this larger algebra.
The new algebra we require is the algebra of bounded Baire functions
of the first type on aCT); that is, the algebra S(a(T)) of bound-
ed functions on aCT) that are pointwise limits of sequences of
functions in C(a(T)). Clearly S(a(T)) is a normed algebra under
pointwise operations and the supremum norm. Our task is to extend
the inverse Gel'fand transformation from C(a(T)) to S(a(T)) in
such a way that the image of S(a(T)) in L(V) will contain suffi-
ciently many orthogonal projections to allow us to approximate T
in an appropriate manner. We shall make this latter assertion more
precise in Theorem 10.6.1.
318 10. B*-Algebras
With this in mind, let us denote the inverse Gel'fand transfor-
mation by that is, C(a(T - is such a that,
if f C(a(T, then E is such that = f. Clearly
, is an isometric *-isomorphism from C(O(T onto Ar. Let
x,y V and define F (f) = f E C(a(T. Apparently
xy
F is a linear functional on C(a(T, and the estimate
xy
IFxy(f)1 <
IIl1xllllyll
=
(f E C(a(T)
reveals that F is a continuous linear functional on C(O(T.
xy
Hence, by the Riesz Representation Theorem for continuous linear
functionals on C(a(T [L, p. 109], there exists a unique bounded
complex-valued regular Borel measure in M(a(T such that
xy
FXy(f) = JO(T) f(") (f C(o(T)
and II = I (a(T < IIxlillyli. Moreover, it is easily verified
xy xy -
that the measures so constructed are such that
(a) + z) : +
(b) + y)z = +
(c) =
(d) =
(x,y,z Vj a C).
Furthermore, since every Baire subset of aCT) is a Borel set
[Ry, p. 302], we see, given X,y V, that
Ia(T) geT)
exists for each g 8(0(T. Consequently, given x V and
g 8(a(T - C(a(T , the equation
10.6. A Spectral Decomposition Theorem
319
(y V)
defines an antilinear functional on V. The antilinearity of G
x
follows from the preceding conditions (a) and (d). Moreover,
IGx(Y) I <
< IIgllCDUxllllyli
(y V)
shows that G is even continuous and that IIG II < Ugil IIxli. It
x x - CD
then follows easily from the Riesz Representation Theorem for con-
tinuous linear functionals on Hilbert spaces [L, p. 390] that there
exists a unique element V such that G (y) =
x
y E V, and = II G
x
ll < UgIICDllxll. Now, given g belonging
to B(a{T - C(a{T , if we repeat the previous argument for each
x E V, we obviously obtain a mapping ,(g) from V to V that
maps x to x E V. The preceding conditions (b) and (c)
show that is linear, while the previous estimate of
reveals that <peg) E L(V) and II <lIgIlCD' Taking cpeg) ,
g E to be the inverse Gel'fand transform of g, as the
notation suggests, we see that we have extended the inverse Gel'fand
transformation cp to a norm-decreasing linear mapping of 8{a(T
into L(V). Actually cp is a *-homomorphism. as shown by the next
lemma.
Lemma 10.6.2. Let V be a Hilbert space over C and let
T E L(V) be self-adjoint. If Ar is the closed subalgebra of
Lev) generated by T and : Ar is the inverse Gel'fand
transformation from onto Ar' then can be extended
to a norm-decreasing linear mapping of 8(aCT into Lev) such
that
(i) cp is an algebra homomorphism.
(ii) cp(g)* = <p(g*), g E B(aCT, where g*(T) = i(Tf, T E aCT).
320 10. B*-Algebras
Proof. We shall content ourselves with proving part (i). A
similar argument establishes part (ii), the details being left to
the reader. We define : B(a(T)) L{V) as in the preceding dis-
cussion. To show that is an algebra homomorphism we need only
prove that = g,h (B(a{T)). To this end we assume
first that g,h ( B(a(T)) are nonnegative. Then there exist mono-
tone increasing sequences (g) and (h) of nonnegative functions
n n
in CrafT)) that converge pointwise to g and h, respectively.
Consequently for each k the sequence (gnhk) C CrofT)) is a mono-
tone increasing sequence that converges pointwise to gh
k
, whence
we deduce, via an application of the Lebesgue Dominated Convergence
Theorem [Ry, p. 229], that
= faCT) ghk(T)
= lim JO(T) gnhk(T)
n
= lim
n
= lim Sa(T) gn(T) ) (x)y(T)
n k
= Ia(T) geT) ) (x)y(T)
k
=
=
(X,y V),
as the inverse Gel'fand transformation on crafT)) is an algebra
isomorphism. A similar argument mutatis mutandis combined with the
preceding result then reveals that
= Ia(T) gh(T)
= lim Ja(T) ghk(T)
k
= lim
k
10.6. A Spectral Decomposition Theorem
= lim
k
= lim
k
= IO(T) hk(T)
= Ia(T) h(T)
=
321
(x,y E V).
However, any g E 8(a(T)) can be written as g = gl - g2 + i(g3 - g4)'
where gk E 8(0(T)) and gk is nonnegative, k = 1,2,3,4. Thus
the foregoing argument shows that for any g,h E 8(0(T)) we have
= (x,yEV),
from which we conclude at once that = g,h B(a(T)).
Therefore is an algebra homomorphism.
o
We are now in a position to see how to approximate a self-adjoint
T in L(V) by orthogonal projections. If T L(V) is self-adjoint,
then, as noted before, aCT) Furthermore, by Theorem 2.1.2, we
have oCT) C [-IITII.IITII]. For each t E aCT) we set
gt = Xu(T) n [-IITII,t]'
where X
E
, as usual, denotes the characteristic function of E.
Clearly gt can be considered as an element of B(o(T)), and so,
by Lemma 10.6.2, = P
t
E L(V). Moreover, since , is a
*-homomorphism and g2 = g = g*, we see that p2 = P = P*; that
t t t t t t
is, P
t
is an orthogonal projection. In this way we obtain a family
of orthogonal projections {P
t
I t E oCT)). Furthermore, if
h E C(aCT)) is such that hCT) = T, T o(T) , that is, h = T,
then, given e > 0, a straightforward argument using the uniform
continuity of h on aCT) reveals that there exist points tk o(T) ,
322 10. B-Algebras
k = 0,1,2, ... ,n, such that -IITII to < tl < ... < tn IITII and
for which
n
lh(T) - t tk[gt (T) - gt (T)]l < I
k=1 k k-l
(T E a(T.
The details are left to the reader. However, since by Lemma 10.6.2
is a norm-decreasing homomorphism from B(a(T to Lev), we
conclude that
n
liT - t tk (P t - P t ) II < e,
k=l k k-l
as = P
t
' k = O,I,2, . ,n, and = T. Obviously this
valid if we add to the original set of t
k
,
k = O,I,2, ,n, and so we see that the limit of the approximating
sums of orthogonal projections exists and is equal to T as the
maximum of the differences tk - t
k
_
l
, k = 1,2, ,n, tends to zero.
From the analogous procedure for the definition of the Riemann-Stielt-
jes integral, we also agree to denote this limit as an integral;
that is, we write
T = IO(T) t dP
t
.
Naturally, the integral notation is just a symbol for the previously
described limiting process.
Clearly the same argument is valid if we replace h by any
f C(a(T since each such f is uniformly continuous on aCT).
We summarize this development in the next theorem.
Theorem 10.6.1. Let V be a Hilbert space over C and let
T E LCV) be self-adjoint. Then there exists a family of orthogonal
projections {P
t
I t E a(T)J contained in L(V) such that
f(T) = I
oCT
) f(t) dP
t
(f C(a(T).
A spectral decomposition theorem for T Lev) that are normal,
that is, such that 11* = T*T, can also be obtained by means of the
Commutative Gel'fand-Naimark Theorem. For further details the reader
is referred to [HOI' pp. 96-100].
10.7. Almost Periodic Functions
323
10.7. Almost Periodic Functions. The notion of an almost per-
iodic function was introduced in Section 1.2. We remind the reader
that, if G is a locally compact Abelian topological group, then
f E C(G) is said to be almost periodic if the set {Ts(f) I s E G)
has compact closure in C{G), where, as usual, Ts(f)(t) = f(t - s),
t E G. Equivalently, f is almost periodic if {Ts(f) I s E G] is
a totally bounded subset of C(G); that is, given e > 0, there
exist sl,s2, ,sn in G such that for each s E G we have
liT (f) - T (flU < e for some k = 1,2, .. ,n. As alreafly indicated,
s sk CD
the set of almost periodic functions on a locally compact Abelian
topological group G, denoted by AP(G), is a closed subalgebra
of C(G) that contains the constant functions and so is a commuta-
tive Banach algebra with identity. Moreover, it is easily seen that,
if f E AP(G), then so also is f, where f(t) = f(t), t G. The
verifications of these assertions are left to the reader. It is now
evident that AP(G) is a commutative B-algebra with identity, the
involution being the obvious one of complex conjugation. Our goal
in this final section is to use the Commutative Gel'fand-Naimark
Theorem to obtain several results about the algebra AP(G) and its
compact maximal ideal space A(AP(G)). In particular, we shall see
that AP(G) is precisely the closure in C(G) of the algebra of
trigonometric polynomials on G, that is, of the algebra of finite
linear combinations of the continuous characters on G, and that
A(AP(G)) is a compactification of G that can be viewed as a com-
pact Abelian topological group. In the interests of completeness
we make the following definition:
Definition 10.7.1. Let X be a topological space. Then a
compact Hausdorff topological space Y is said to be a compactifi-
cation of X if there exists a continuous injective mapping
: X -Y such that t(X) is dense in Y.
The reader should note that this is not the usual topological
definition of a compactification (see, for example, [W
2
, pp. 151
and 152]), but it is the one appropriate to the discussion of this
section.
324 10. B*-Algebras
The one-point compactification of a locally compact Hausdorff
topological space, which we have had occasion to use at various points
in the preceding sections, is a simple example of a compactification.
We shall shortly examine one that is considerably more complicated.
Our introductory remarks about AP(G) and the Commutative
Gel'fand-Naimark Theorem (Theorem 10.2.1) combine to yield the next
theorem.
Theorem 10.7.1. Let G be a locally compact Abelian topolo-
gical group. Then the Gel'fand transformation on AP(G) is an
isometric *-isomorphism of AP(G) onto
Now in the usual manner we see that each point t (G defines
a complex homomorphism of AP(G), namely, the complex homomorphism
T
t
defined by TtCf) = f(t), f (AP(G). Clearly the mapping
t(t) = T
t
, t (G, maps G into 6(AP(G)). We claim that is
actually continuous and injective and t(G) is dense in 6(AP(G)),
whence we can conclude that is a compactification of G.
Before we can establish this, however, we need to prove a
lemma. We shall denote by T(G) the algebra of all trigonometric
polynomials on a locally compact Abelian topological group G.
Lemma 10.7.1. Let G be a locally compact Abelian topological
group. Then T(G) C AP(G).
Proof. It obviously suffices to prove that (.,y) AP(G) for

each y E G. But if s (G, then Ts[('Y)] = (s,y)(,y), and so
{Ts[("Y)] , s E G] is a subset of {aC,Y) , a E fl. The latter
set, however, is compact in C(G), being the continuous image of f
under the mapping a a(,y), a f. Thus (.,y) is almost peri-

odic for each y E G, and consequently T(G) C AP(G).
o
Theorem 10.7.2. Let G be a locally compact Abelian topologi-
cal group. Then is a compactification of G.
10.7. Almost Periodic Functions
325
Proof. We need on}y show that the mapping : G - a(AP(G))
defined above is a continuous injective mapping such that V(G) is
dense in aCAPCG)). The continuity of follows at once from the
definition of V and the fact that the topology on a(AP(G)) is
the relative weak* topology. If t,s E G and vet) = ,(5), then,
since T(G).C AP(G), we see that (t,y) = Tt!Co,y)] = Ts[('Y)] =
(s,y), y (G. However, by Corollary 4.7.3, G separates the points
of G, whence we conclude that t = s. Hence is injective.
Finally, if ,(G) is not dense in a(AP(G)), then cl[V(G)]
is a proper closed subset of a(AP(G)). Thus, since C[a(AP(G))]
is a regular commutative Banach algebra (or, equivalently, by Ury-
sohn's Lemma [W
2
, p. 55]), there exists some h E C[a(AP(G))] such
that h 0 but h[V(t)] = 0, t E G. But by Theorem 10.7.1 there

exists some f AP(G) such that f = h. Moreover,
(t E G),

from which we see immediately that h = f = 0, contrary to the
choice of h. Thus ,(G) is dense in a(AP(G)). and a(AP(G))
a compactification of G.
It is important to note that the mapping : G - a(AP(G))
is
o
is
not in general a homeomorphism unless G is compact. In the latter
case the assertion is a well-known fact of general topology [W
Z
'
p. 83].
Actually a good deal more can be said about a(AP(G)) than
just that it is a compactification of G. Indeed, one can extend
the group structure of G to all of in such a way that
a(AP(G)) becomes a compact Abelian topological group. It is appar-
ent how to make t(G) into a group. Namely, one defines
yet) + '(5) = ,Ct + 5), t,s (G. It is easlily verified that this
definition does make t(G) into an Abelian group, and it also, in-
cidentally, makes a group isomorphism. It is not quite so clear,
however, how one should define T + W for arbitrary E a(AP(G)).
326 10. B*-Algebras
To see how this can be done we let T,w E 6(AP(G)) and let
and (W) be families of open neighborhoods of T and w,
a
tively, which are directed by inclusion and such that n u =
aa
and n w = (w). For each a let
aa
E = (,(t) + t(w) I t,w E G, t(t) E U , ,(w) E W ).
a a a
(U )
a
respec-
(T)
It is easily seen that this family of sets has the finite-intersec-
tion property, and hence, since 6(AP(G)) is compact, we see that
n cl(E ) ~ ~ . We claim that n cl(E) is a single point. Indeed,
a a a a
suppose K
I
,K
2
E nacl(E
a
). Since the topology on ~ A P G ) ) is the
relative weak* topology, to show that KI = K2 it suffices to prove
that KI(f) = K
2
(f), f E AP(G). So let f E AP(G) be given and
suppose > o. Since f is almost periodic, there exist
sl,s2, ... ,sn in G such that for any s E G we have
liT (f) - T (f) Ii < /18 for some k = 1,2, ... ,n. Let U(T) =
s sk Q)
U(T;S/36;f
O
,f
l
, .. ,f
n
) be the weak* open neighborhood of T deter-
mined by fO = f, fk = TSk(f), k = 1,2, ... ,n, and /36. That is,
v E UeT) if and only if Iv(f
k
) - T(fk)l < 6/36, k = 0,1,2, ... ,n.
Similarly we set U(w) = U(w;6/36;f
o
,f
l
, .. ,f
n
). Clearly, if t,s E G
and either 'Ct),,(s) E U(T) or t(t),t(s) E U(w), then
Ck = 0,1,2, ... ,n).
Furthermore, we claim that, if t(t),,(s) E U(T) and t(w),,(v) E U(w),
then
Irtct) + tCw)] (f) - r,Cs) + tCv)](f)1 = If(t + w) - f(s + v)l
To see this let Sj and sk be such that
and IIT_vCf) - T (f) II < /18. Then
sk Q)
IIT_t(f) - T (f) Ii < s/18
Sj CD
If(t + w) - f(s + v)l < If(t + w) - f(t + v)1
10.7. Almost Periodic Functions
+ If(t + v) - f(s + v)1
= IT_
t
(f)(w) - T_t(f) (v) 1
+ 1 T (f) (t) - T (f) (s) I
-v -v
< IT_t(f)(w) - Ts. (f) (w)1
J
+ IT (f)(w) - T (f) (v) 1
s. s.
J J
+ ITs. (f) (v) - T_t(f) (v)l
J
+ 1 T (f) (t) - T (f) (t) I
-v sk
+ IT (f) (t) - T (f) (s) 1
Sk Sk
+ IT (f)(s) - T (f)(s)1
sk -v
< IIT_
t
(f) - T
5
. (f) 1\0) + lfj(w) - f
j
(v)1
J
+ IITs. (f) - T_
t
(f)II
CD
J
+ liT (f) - T (f) II
-v sk CD
327
+ lfk(t) - fk(S)l + liT (f) - T (f) \I
sk -v CD
e
<-
3
by the previous estimates.
Consequently frcm the definition of the E we see that there
ex
exist t(t),'(s) E U(T) and t(w),.(v) E U(w) such that
and
whence we deduce that
328 10. B*-Algebras
IKl(f) - K
2
(f)1 IKl(f) - [y(t) + ,(w)] (f)1
+ 1. (t) + t (w)] (f) - rt (s) + ,(v)] (f) I
+ Ir,(s) + ;(v)] (f) - K
2
(f)1
e e e
<-+-+-
333
= e.
Since e > 0 is arbitrary, we have KI(f) = K
2
(f), and so KI = K
2
.
Hence n cl(E) is a single point, call it K. We define
Ot a
T + w = x. The foregoing arguments show that T + w is uniquely
defined, and they also entail that the operation of addition so de-
fined is continuous in the Gel'fand topology on A(AP(G)). A simi-
lar development shows that inverses exist and that the operation of
inversion is also continuous with respect to the Gel'fand topology.
The details are left to the reader. In this way one extends the
group structure of G to all of A(AP(G)), and with this group
structure on A(AP(G)) we see that A(AP(G)) is a compact Abelian
topological group.
We include the results of the preceding develppment in the next
theorem.
Theorem 10.7.3. Let G be a locally compact Abelian topologi-
cal group. Then A(AP(G)) is a compactification of G, and the
group structure of G can be extended to A(AP(G)) in such a way
that A(AP(G)) becomes a compact Abelian topological group. More-
over, suppose H is any compact Abelian topological group and
: G - H is a continuous isomorphism such that
(i) H is a compactification of G.
(ii) The mapping defined by = t G
and h (C(B), is an algebra isomorphism of C(H) onto AP(G).
Then H is topologically isomorphic to A(AP(G)J.
10.7. Almost Periodic Functions
329
Proof. We need only prove the last assertion of the theorem.
Suppose H and : G - H are as described. The mapping
t : G - A(AP{G defined above has the same properties as and
so we see that is an algebra isomorphism of C[A(AP(G]
onto C(H). Consequently, by Theorem 4.1.4, Hand A(AP{G are
homeomorphic. Then, if : H - A(AP(G is a homeomorphism of H
onto A(AP(G, we see at once that is an extension of the
group isomorphism 0 : - ,(G). Since and t(G)
are dense in H and A(AP(G, respectively, we conclude that
is a group isomorphism. Therefore Hand A(AP(G are topologi-
cally isomorphic.
o
Considered as a compact Abelian topological group, A(AP(G
is generally called the Bohr or almost periodic compactification of
G, and it is often denoted by bG. We shall do this in the remain-
der of this section in order to emphasize that bG is a topological
group rather than just the maximal ideal space of AP(G). In view
of the remark following Theorem 10.7.2 we see that, if G is a
compact Abelian topological group, then bG is topologically iso-
morphic to G.
The following lemma is now easily proved:
Lemma 10.7.2. Let G be a locally compact Abelian topological
..
group. Then reG) = r(bG).

Proof. Obviously it suffices to show that, if y (G, then
..
(.,y) is a continuous character of bG and that every such char-
..
acter of bG is obtained in this fashion. Of course, (.,y) de-

notes the Gel'fand transform of (.,y). But if Y E G and t,s E G,
then

( y) [,(t) + '(5)] = (t + s,y)
= (t,y)(s,y)

= (.,y) rt(t)](.,y) ['(5)],
330 10. B*-Algebras
where , : G - bG is as before. Since ,(G) is dense in bG and

(.,y) is continuous, we conclude that (.,y) is a continuous
character of bG. Conversely, suppose h E C(bG) is a continuous
character of bG. By Theorem 10.7.1 there exists a unique f E AP(G)

such that f = h. Moreover, if t,s (G, then
f(t + 5) = y(t + 5) (f)

= f[,(t) + ,(5)]
= h[i(t) + ,(5)]
= h[,Ct)]h[,(s)]
= f(t)f(s),
whence we see at once that f is a continuous character of G. Thus

there exists some y E G such that (.,y) = f and (.,y) = h. C
Since, by Corollary 4.7.3, the dual of a locally compact Abelian
topological group G separates the points of G, an application
of the Stone-Weierstrass Theorem [L, p. 333] reveals that reG) is
norm dense in C(G) whenever G is a compact Abelian topological
group. Combining this observation with Lemma 10.7.2 and the fact
that the Gel'fand transformation on AP(G) is an isometry, we obtain
the following theorem and corollary:
Theorem 10.7.4. Let G be a locally compact Abelian topologi-
cal group. Then reG) is norm dense in AP(G).
Corollary 10.7.1. Let G be a compact Abelian topological
group. Then AP(G) = C(G).
Another application of Lemma 10.7.2 yields the next result.
Theorem 10.7.S. Let G be a locally compact Abelian topologi-

cal group and let G
d
deno!e the algebraic group G with the.dis-
crete topology. Then (bG) is topologically isomorphic to G
d

10.7. Almost Periodic Functions 331

Proof. From Corollary 10.5.1 we know that (bG) is a dis-
crete group, so to prove the theorem we need only show that there

exists a group isomorphism of G
d
onto (bG) The proof of Lemma
10.7.2, however, shows that the Gel'fand transformation on AP(G)
provides such an isomorphism.
o

Replacing G by G in this theorem and applying the Pontryagin
Duality Theorem (Theorem 10.5.2), we obtain the next corollary, the
details being left to the reader.
Corollary 10.7.2.
gical group and let G
d
discrete topology. Then
Let G be a locally compact Abelian topolo-
denote the algebraic group G with the

bG is topologically isomorphic to (G
d
) .
A moment's reflection reveals that this corollary has the fol-
lowing significance: If G is a locally compact Abelian topological

group, and G is the group of the continuous characters of G,

then bG is the group of all the characters of G, that is, of
all the bounded functions h on G such that h(t + s) = h(t)h(s),

s,t E G. For example, if G = R, then, since R = ~ we see that
HR is the group of characters of it This observation leads to the
following approximation result for characters:
Corollary 10.7.3. Let G be a locally compact Abelian topolo-
gical group and let h be a character of G. If s > 0 and

t
l
.t
2
... ,t
n
are in G, then there exists some y E G such that
Ih(t
k
) - (tk,y)l < I, k = 1,2, ... ,n.
Proof. By Corollary 10.7.2 we can consider h as an element

of bG. Since each tk (G defines the almost periodic function

(t
k
,) E AP(G), k = 1,2, ... ,n, we deduce, on recalling that bG =

~ A P G and the definition of the Gel'fand topology on A(AP(G).
that

(g I g E bG, Ih(t
k
) - g(tk)1 < s, k = 1.2 .. ,n)
332
10. B*-Algebras

defines an open neighborhood of h in bG. Consequently, on iden-

tifying G with tCG) C bG, we see that there is some y E G for
which Ih{t
k
) - (tktY)I < 't


k = 1,2, .. ,n, because G is dense
in bG.
o
This corollary is actually an abstract version of the classical
Kronecker approximation theorem, but we shall not pursue this obser-
vation since it requires more detailed information about the Bohr
compactification of specific groups than we have at our disposal.
For this and other general discussions of almost periodic functions
and the Bohr compactification we refer the reader to [Bo; BR
I
, pp.
245-261, 430-438; HR
2
, pp. 310-312; Lo, pp. 165-173; Ru
1
, pp. 30-32;
We, pp. 130-139].
[A]
[Ba]
[BaNr]
[Bo]
[B]
(OkRa]
[Ga]

[Go]
[He]
REFERENCES
L.V. Ahlfors, Complex Analysis, 2nd ed., McGraw-Hill, New
York, 1966.
G. Bachman, Elements of Abstract Harmonic Analysis, Academic
Press, New York, 1964:-
G. Bachman and L. Narici, Functional Analysis, Academic
Press, New York, 1966.
H. Bohr, Almost Periodic Functions, Chelsea, New York, 1951.
A. Browder, Introduction to Function Algebras, Benjamin,
New York, 1969.
N. Dunford and J. T. Schwartz, Linear Operators, Part !:
General Theory, Interscience, New York, 1958.
N. Dunford and J. T. Schwartz, Linear Operators, Part II:
Spectral Theory, Interscience, New York, 1963.
C.F. Dunkl and O. E. Ramirez, Topics in Harmonic Analysis,
Appleton-Century-Crofts, New York, 1971.
R.E. Edwards, Functional Analysis:Theory and Applications,
Holt, Rinehart, and Winston, New York, 1965.
R.E. Edwards, Fourier Series: A Modern Introduction, Vol. I,
Holt, Rinehart, and Winston, New York, 1967.
T. Gamelin, Uniform Algebras, Prentice-Hall, Englewood Cliffs,
N.J., 1969.
I.M. Gel'fand, O.A. Raikov, and G.E. "Commutative
Normed Rings", American Mathematical Society Translations
115-220(1957).
R.L. Goodstein, Complex Functions, McGraw-Hill, New York,
1965.
G. Helmberg. Introduction Spectral Theory in Hilbert
Space, North Holland, Amsterdam, 1969.
333
334
[H]
[Hy]
[HIP]
[Hr]
[J]
[Ka]
[L]
[Lb]
[LeRd]
[Lo]
[Nb]
[N]
References
E. Hewitt, "A survey of abstract harmonic analysis", Surveys
in Applied Mathematics: Some Aspects of Analysis and Prob-
ability. Vol. !. Wiley. New York. 1958. pp. 105-168.
E. Hewitt and K.A. Ross, Abstract Harmonic Analysis, Vol. !,
Springer-Verlag, Berlin-GHttingen-Heidelberg, 1963.
E. Hewitt and K.A. Ross, Abstract Harmonic Analysis, Vol. II,
Springer-Verlag, New York-Heidelberg-Berlin, 1970.
H. Heyer, Dualitlt lokalkompakter Gruppen, Lecture Notes in
Mathematics, No. 150, Springer-Verlag, Berlin-Heidelberg-
New York, 1970.
E. Hille and R.S. Phillips, Functional Analysis and Semi-
groups, American Mathematical Society Colloquium Publication
XXXI, American Mathematical Society, Providence, R.I., 1957.
K. Hoffman, Fundamentals of Banach Algebras, Instituto de
Matematica da Universidade-do Parana, Curitiba, Brazil, 1962.
K. Hoffman, Banach sraces of Analytic Functions, Prentice-
Hall, Englewood C l i ~ s N.J., 1962.
L. HHrmander, An Introduction to Complex Analysis in Several
Variables, D. Van Nostrand, Princeton, N.J., 1966.
B.E. Johnson, "The uniqueness of the (complete) norm topology",
Bulletin of the American Mathematical Society,73,537-539(1967).
Y. Katznelson, An Introduction to Harmonic Analysis, Wiley,
New York, 1968.--- ---
R. Larsen, Functional Analysis: An Introduction, Dekker,
New York, 1973.
G.M. Leibowitz, Lectures on Complex FURction Algebras, Scott,
Foresman, Glenview, Ill., 1970.
N. Levinson and R.M. Redheffer, Complex Variables, Holden-
Day, San Francisco, 1970.
L. Loomis, An Introduction to Abstract Harmonic Analysis,
D. Van Nostrand, Princeton,-W.J., 1953.
L. Nachbin, The Haar Integral, D. Van Nostrand, Princeton,
N.J. J 1965.
M.A. Naimark, Normed Rings, Noordhoff, Groningen, The Nether-
lands, 1964.
References
335
[Ph]
[Pi]
[Po]
[Ri]
[RzNg]
[Ry]
[Ru
l
]
[5]
[Wa]
[We]
[Wr
l
]
R.R. Phelps, Lectures ~ Choquet's Theorem, D. Van Nostrand,
Princeton, N.J., 1966.
H.R. Pitt, Tauberian Theorems, Oxford University Press,
London, 1958.
L. Pontryagin, Topological Groups, Princeton University
Press, Princeton, N.J., 1958.
C.E. Rickart, General Theory of Banach Algebras, D. Van Nos-
trand, Princeton, N.J., 1960.
F. Riesz and B. Sz-Nagy, Functional Analysis, Ungar, New York,
1955.
H. Royden, Real Analysis, 2nd ed., Macmillan, New York, 1966.
W. Rudin, Fourier Analysis ~ Groups, Interscience. New York,
1962.
W. Rudin, Real and Complex Analysis, McGraw-Hill, New York,
1966.
E.L. Stout, The Theory of Uniform Algebras, Bogden and
Quigley, Tarrytown-on-Hudson, N.Y., 1971.
J.K. Wang, Lectures ~ Banach Algebras, Lecture Notes, De-
partment of Mathematics, Yale University, New Haven, Conn.,
1965.
A. Weil, L'integration dans les groupes topologiques et ~
applications, Actualites scientifiques et Industrielles
8690-1145, Hermann, Paris, 1951.
J. Wermer, "Banach Algebras and Analytic Functions", Advances
in Mathematics, Vol. 1, fasc. 1, Academic Press, New York, 1961.
J. Wermer, Banach Algebras and Several Complex Variables,
Markham, Chicago, 1971.
N. Wiener, The Fourier Integral and Certain of Its Applica-
tions, Dover, New York, 1958.
N. Wiener, Generalized Harmonic Analysis and Tauberian Theo-
~ M.I.T. Press, Cambridge, Mass., 1964.
A. Wilansky, Functional Analysis, Blaisdell. New York, 1964.
A. Wilansky, TOpology for Analysis, Ginn, Waltham, Mass., 1970.
J.H. Williamson,"Remarks on the Plancherel and Pontryagin
theorems", Topology, 1,73-80(1967).
INDEX
..
A, 74
Ar, 316
A[e], 9
A(K), 17
AP(G), 18
A(x), 150
A
A ~ (0), assumed after E!&! 4
A_I' 29
A-I, 29
connected component of the
identity in, 156,158
AC(r), 17
Gel'fand representation of, 102
of absolutely convergent
Fourier series, 22
radical, 82
separating function, 220
uniform, 255
almost periodic compactification,
329
almost periodic function. 18,323
apPlications of boundaries,
243-249
approximate identity, 109
in LI(G), 109, 188
Arens' Theorem, 48
Arens-Royden Theorem, 158
B
maximal ideal space of, 102 bG, 329
~ l o v boundary for, 228
A(D), 17
Bishop boundary for, 228
Gel'fand representation of, 97
is not regular, 167-168
maximal ideal space of, 97
~ l o v boundary for, 228
adjoining an identity, 10
algebra, 3
Banach, 4
commutative, 3
division, 35
normed, 4
bdy(E), 51
8(a(T)), 317
B*-algebra, 273
examples of, 273-275,282,323
Banach algebra, 4
337
commutative semisimple, 81
conditions to have an identity,
153-154,178
finitely generated, 98
normal commutative. 176
regular commutative, 166
self-adjoint commutative, 132
338
Tauberian commutative, 183
with involution, 273
belong locally to I, at a
point, 201
at infinity, 201
Beurling-Gel'fand Theorem, 81
Bishop boundary, example of non-
existence of, 230-232
existence for a uniform alge-
bra, 268
for A(D), 228
for C(X), X compact metric,
227-228
for a commutative Banach alge-
bra, 226
for a separating function alge-
bra, 221
Bishop-deLeeuw Theorem, 236-237
Bohr compactification, 329
uniqueness of, 328
boundary, applications of,
243-249
Bishop, 221
Choquet, 242
existence of, 242
C(V) , 23
XA, 242
X
E
' 93-94
C(X), 16
Index
Bishop boundary when X is
compact metric, 227-228
Choquet boundary for, 242
closed ideals in, 195
example of pC(X) = 230-232
Gel'fand representation of,
87-88
maximal ideal space of, 86,90
is regular, 167
sets of spectral synthesis for,
196
ilov boundary for, 227
C (X), 16
o Gel'fand representation of, 88
is regular, 167
maximal ideal space of, 88
~ i l o v boundary for, 227
C
n
( [ a , b] ), 17
Gel'fand representation of, 92
maximal ideal space of, 92
~ i l o v boundary for, 228
for a commutative Banach alge- C*-algebra, 275
bra, 226
for a separating function alge- character, 331
bra, 221 continuous, 113
~ i l o v 225
existence of, 222
C, 3
Cc(X), 16
CAo(G), 286
C
cl(I), cl(E), 34,162
co(E), 233
co(E), 233
characterization of singular ele-
ments, in C(X), 43
in commutative Banach alge-
bras, 81
in self-adjoint semisimple
commutative Banach algebras,
141
Choquet-Bishop-deLeeuw Theorem, 271
Choquet boundary, characterization
of, 257,263
existence of, 242
for A(D), 242
for C(X), 242
Index 339
for separating function alge- Oitkin's Theorem, 206
bras, 242
closed convex hull, 233
closed ideals in C(X), 195
in L
1
(G), G compact, 197
closure operation, 162
commutative algebra, 3
Commutative Gel'fand-Naimark
Theorem, 277
compactification, 323
division algebra, 35
dual group, 127
E
E, 162
1}, 168
1}(x), 46
almost periodic, 329 exp, 156
Bohr, 329
90 ext(E), 233
complex homomorphism, 67 extreme point, 233
connected component,
of the identity in A , 156,158 F
continuity of inversion, 31
of quasi-inversion, 29
continuous character, 113
contour, regular, 143
spectral, 144
convex hull, 233
closed, 233
convolution, 19-20
D
0, 17
()A, 225
A(A), 69
assumed non empty after E!&! 78
discrete topological group, 106
Ditkin's condition, 204
1, 105
o
f, 287

f, 21,114,169,307-308

f(k), 21
f * g, 19
IIfllAc' 22
IIflln, 18
IIfllp' 19
IIfIlCD, 16,18
finitely generated Banach algebra,
98
finitely generated commutative
Banach algebra, maximal ideal
space of, 99
Gel'fand representation of, 99
340
Fourier coefficient, 21
transformation, 21,115
Fourier-Stieltjes transform, 128
homomorphism, complex, 67
*-, 277
homomorphism space, 69
function, almost periodic, 18,323 hull, 160
slowly oscillating, 190
functions which operate, 151
G
..
G, 113
r, 17,21

f, 116
y , 121
o
Gel'fand-Mazur Theorem, 35
Gel'fand representation, 76
of AC(f), 102
of A(D), 97
of C(X), 87-88
of Co(X), 88
of Cn([a,b]), 92
of a finitely generated commu-
tative Banach algebra, 99
of L1(G), 114
of L X , S , ~ ) , 93
CD
Ge1'fand Representation Theorem,
74
Gel'fand topology, 71
Ge1'fand transform, ;6
hull-kernel closure, 162
topology, 163
I , 11
e
10 (E), 182
int(K), 17
ideal, 4
left, 4
maximal, 4
modular, 7
primary, 216
proper, 4
regular, 7
right, 4
two-sided, 4
identity modulo
inverse, 5
left, 5
right, 5
I
I, 7
Inversion Theorem, 298,301
invertible, 5
involution, 273
Index
Gel'fand transformation, 76 isomorphism, topological, 131
H J
hel), 160 Jo(E), 182
Haar measure, 18 joint spectrum, 98
Index
K
k (E), 160
kernel, 160
Kronecker approximation theorem,
332
Lp (G), 19
L(V) , 22
A (G), 282
o
A, 18
L
Ll(G), closed ideals when G is
compact, 197
closed translation-invariant
linear subspace of, 186
condition to have an identity,
107
contains an approximate iden-
tity, 110,188
failure of spectral synthesis
when G is noncompact, 199
Gel'fand representation of, 114
is regular, 173
is semisimple, 117
is Tauberian, 184
maximal ideal space of, 114,126
satisfies Ditkin's condition,
214
sets of spectral synthesis, 215
sets of spectral synthesis when
G is compact, 198
boundary for, 228
L (X,S ,toLL 18
Gel'fand representation of, 93
maximal ideal space of, 93
left ideal, 4
inverse, 5
modulus of integrity, 46
341
quasi-inverse, 13
topological zero divisor, 40
zero divisor, 40
linear functional, multiplicative,
67
positive, 259
Local Maximum Modulus Theorem, 220
M , 12
e

128
20
20
* \I, 20
M(G), 20
is semisimple, 315
maximal ideal space of, 128
Malliavin's Theorem, 199
maximal ideal, 4
maximal ideal space, 69
of AC(r), 102
of A(D), 97
of C(X), 86,90
of Co(X), 88
of cn([a,b), 92
of a finitely generated commu-
tative Banach algebra, 99
of LI(G), 114,126
of L 93
of MTG), 128
maximum modulus set, 221
Maximum Modulus Theorem, 218-219
Mergelyan's Theorem, 101
modular ideal, 7
342 Index
modulus of integrity, 46
left, 46
Polynomial Spectral Mapping Theo-
rem, 57
right, 46
Pontryagin Duality Theorem, 312
multiplicative linear functional,
67 positive linear functional, 259
primary ideal, 216
N
N(K,e), 293,309-310
nilpotent, 33
topological, 33
proper ideal, 4
Q
quasi-inverse, 13
Noncommutative Gel'fand-Naimark left, 13
Theorem, 280 right, 13
normal commutative Banach algebra,quasi-invertible, 13
176
quasi-regular, 13
normed algebra, 4
quasi-singular, 13
p
P(K), 17
Parseval's Formula, 170,305

115-116
peak point, for a commutative
Banach algebra, 226 R(K), 17
for a separating function alge-
bra, 221 R(f), 55
weak, for a separating func-
tion algebra, 262 Rad(A), 82
peak set, for a separating func- pA, 221
tion algebra, 262
radical, 82
perfect set, 205
R
radical algebra, 82
Plancherel's Theorem, 169,301,
307-308 regular commutative Banach algebra,
166
P1anchere1 transform, 169,305 contour, 143
curve, 143
Plancherel transformation, 169,305 simple, 143
simple closed, 143
Index
element, 5
ideal, 7
representing measure, 253
and the Choquet boundary, 257
and the ilov boundary, 269-270
Riemann-Lebesgue Lemma, 114
right ideal, 4
inverse, 5
modulus of integrity, 46
quasi-inverse, 13
topological zero divisor, 40
zero divisor, 40
s
a(x), 54
*-homomorphism, 277
343
for C(X), 227
for Co(X), 227
for CO((a,b]), 228
for L (G), 228
for a !ommutative Banach alge-
bra, 226
for a separating function alge-
bra, 225
ilov-Arens-Calderon Theorem, 151
simple closed regular curve, 143
regular curve, 143
singular element, 5
in C(X), 43
in a commutative Banach alge-
bra, 81
in a self-adjoint semisimple
commutative Banach algebra,
141
slowly oscillating function, 190
self-adjoint, 132,275 smooth are, 143
commutative Banach algebra, 132
element in a B-algebra, 275 Spectral Mapping Theorem, 152
semisimple, 81 spectral radius, 58
separating function algebra, 220 Spectral Radius Formula, 58
set of spectral synthesis, 194
for C (X), 196
spectral radius norm, 130
for L
l
(G), G compact, 198 spectral synthesis, failure in
for L (G), 215 Ll(G) when G is noncompact,
for a regular commu- 199
tative Banach algebra that in C(X), 196
satisfies Ditkin's condi- in L1(G), G compact, 198
tion,208-209 in L
l
(G),
set, of spectral synthesis, 194
perfect, 205
ilov boundary, 225
and representing measures,
269-270
characterization of,
existence of, 222
for AC (n, 228
for A(D), 228
225
in a commutatIve B-algebra,
278
in a semisimple regular commu-
tative Banach algebra that
satisfies Ditkin's condi-
tion, 209
problem of, 194
set of, 194
spectrum, 54,69
joint, 98
344
compactification, 90
strong boundary point, 222,262
structure space, 69
suba1gebra, 4
superalgebra, 48
support of 235
symmetric set, 275
IITII, 22
T , 46
x
TX, 46
Tf' 280
T (f), 18,104
s
T(G) , 324
T, 66
T , 69-70
e
T , 70
CD
T
t
, 86
(t,y), 114
T
Tauberian commutative Banach
algebra, 183
Tauber's Theorem, 189
theorem of Arens, 48
Arens and Royden, 158
Beurling and Gel'fand, 81
Bishop and deLeeuw, 236-237
Choquet, Bishop, and deLeeuw,
271
Index
Ditkin, 206
R.E. Edwards, 38
Gel'fand, 23,74
Ge1'fand and Mazur, 35
Gel'fand and Naimark, 277,280
Ma11iavin, 199
Mazur, 40
Mergelyan, 101
Pitt, 191
Plancherel, 169,301,307-308
Pontryagin, 312
Riemann and Lebesgue, 114
61,222
Arens, and Calderon, 151
Tauber, 189
Wermer, 255
Wiener, 104,190
Wiener and 154
topological isomorphism, 131
. zero divisor, 40-41
in C(X), 43
in a separating function al-
gebra, 246-247
topologically nilpotent, 33
totally disconnected, 95
transform, Fourier, 21,115
Ge1'fand, 76
P1ancherel, 169,305
transformation, Fourier, 21,115
Gel'fand, 76
order preserving, 93
P1anchere1, 169,305
translation invariant linear sub-
space, 186
translation operator, 18,104
trigonometric polynomial, 187
two-sided ideal, 4
topological zero divisor, 40
zero divisor, 40
Index
U
U(K,c), 123
U(T;C;x
l
,x
2
, . ,x
n
), 71
uniform algebra, 255
unique maximum point, 222
v
vanish on an open set, 235
w
weak peak point, 262
Wermer's Maximality Theorem, 255
Wiener's Tauberian Theorem, 190
Wiener's Theorem, 104
Wiener-L:vy Theorem, 154
x
-1
x ,6

X, 74
o
x, 174
x*, 139,273
IIxlla, 58

lIxll., 134
x 0 y, 13
~ x ) , 46
Z, 3

Z, 116
Z(x), 160
Z (I), 160
z
zero-dimensional, 93
zero divisor, 40-41
left J 40
left topological, 40
right, 40
right topological, 40
topological, 40-41
two-sided, 40
two-sided topological, 40
zero set of an element, 160
of an ideal, 160
345
about the book . . .
This textbook designed to provide an introduction to the theory of commutative
Banach algebras. The book not only deals with abstract theory, but illustrates the
usefulness of Banach algebras in the study of harmoni c analysis and function al-
gehras as we I!.
The first half of the book is primarily devoted to the general theory of Banach
algebras. Many specifi c exampl es are included in this part to illustrate the prin-
ci pies discussed. The second half of -the book considers more speciali zed '.opics,
among which are: Wiener's Tauberian Theorem; the problem of spectral synthes is ;
the Bishop. 'Choquet , and Silov boundari es; and Wermer's Maximality Theorem.
The second part also examines the Commutative Gel'fand-Naimark Theorem,
Plancherel' s Theorem, the Pontryagin Duality Theorem, almost periodi c func-
li ons. and the Bohr compacti fi Ci_lti on.
Bana('h Algebras a st imul at ing text ror graduate student s with a background
in abstract algebra. ge neral w[',,:ogy, real and compl ex analysis , and functional
analys is.
about the author .. .
RONALD L ARSEN has been Associate Professor of Mathemati cs at Wcsicyan Universit y
in Middletown, Connecticut. since 1970. He previously taught at the University of Cali-
fornia al Sanla Cruz (1965- 70).anJ Yale University (1963- 65). He received hi s B.S.
(1957) and M.S. ( 1959) from Michigan State University and hi s Ph. D. (1964) from
Stanford University.
Professor Larsen' s rcsC!arl:h interests center on harmoni c analysis and Banach algebras.
He is the author of thre,e books. The Multiplier Problem, Allllllroductioll to the Theory of
Multipliers, and Functional Allal)'sis: All Iwroc/uction (Marcel D=kker, Inc.), and
numerous articles.
The author the recipient of a Ful bright-Hays Act Advanced Research Grant to
the University of Oslo ( 1968-69) . He is a member of the Ameri can M.Hhematical
Society. the Mathemat ical Assoc iation of Ameri ca, and the Norsk Malematisk Forening.
Prin(ed ill rltl' Ullired of AIII('rica
ISBN: 0- 8247- 6078- 6
MARCEL DEKKER, INC., NEW YORK

You might also like