Edwards
Edwards
Edwards
64
Editorial Board
F. W. Gehring
P.R. Halmos
Managing Editor
c.c. Moore
R.E. Edwards
Fourier Series
A Modem Introduction
Volume 1
Second Edition
Springer-Verlag
;fC f ~ ..- f. }f~ .",. iJ
R.E. Edwards
Institute for Advanced Studies
The Australian National University
Canberra, A.C.T. 2600
Australia
Editorial Board
Edwards, Robert E
Fourier series, a modem introduction.
This reprint has been authorized by Springer-Verlag (8erlinfHeidelberg/New York) for sale in
the People's Republic of China only and not for export therefrom.
Reprinted in China by Beijing World Publishing Corporation, 2003
PREFACE TO REVISED EDITION
(OF VOLUMED
v
PREFACE
The principal aim in writing this book has been to provide an intro-
duction, barely more, to some aspects of Fourier series and related topics
in which a liberal use is made of modem techniques and which guides the
reader toward some of the problems of current interest in harmonic analysis
generally. The use of modem concepts and techniques is, in fact, as wide-
spread as is deemed to be compatible with the desire that the book shall
be useful to senior undergraduates and beginning graduate students, for
whom it may perhaps serve as preparation for Rudin's Harmonic Analysis
on Groups and the promised second volume of Hewitt and Ross's Abstract
Harmonic Analysis.
The emphasis on modem techniques and outlook has affected not only
the type of arguments favored, but also to a considerable extent the choice
of material. Above all, it has led to a minimal treatment of pointwise con-
vergence and summability: as is argued in Chapter 1, Fourier series are not
necessarily seen in their best or most natural role through pointwise-tinted
spectacles. Moreover, the famous treatises by Zygmund and by Baryon
trigonometric series cover these aspects in great detail, wl:tile leaving some
gaps in the presentation of the modern viewpoint; the same is true of the
more elementary account given by Tolstov. Likewise, and again for reasons
discussed in Chapter 1, trigonometric series in general form no part of the
program attempted.
A considerable amount of space has been devoted to matters that cannot
in a book of this size and scope receive detailed treatment. Among such
material, much of which appears in small print, appear comments on diverse
specialized topics (capacity, spectral synthesis sets, Helson sets, and 80
forth), as well as remarks on extensions of results to more general groups.
The object in including such material is, in the first case, to say enough for
the reader to gain some idea of the meaning and significance of the problems
involved, and to provide a guide to further reading; and in the second case,
to provide some sort of "cultural" background stressing a unity that
underlies apparently diverse fields. It cannot be over-emphasized that the
book is perforce introductory in all such matters.
The demands made in terms of the reader's active cooperation increase
vii
viii PREFACE
fairly steadily with the chapter numbers, and although the book is surely
best regarded as a whole, Volume I is self-contained, is easier than Volume II,
and might be used as the basis of a short course. In such a short course, it
would be feasible to omit Chapter 9 and Section 10.6.
As to specific requirements made of the reader, the primary and essential
item is a fair degree of familiarity with Lebesgue integration to at least
the extent described in Williamson's introductory book Lebesgue Integration.
Occasionally somewhat more is needed, in which case reference is made to
Appendix C, to Hewitt and Stromberg's Real and Abstract Analysis, or to
Asplund and Bungart's A First Course in Integration. In addition, the
reader needs to know what metric spaces and normed linear spaces are, and
to have some knowledge of the rudiments of point-set topology. The remain-
ing results in functional analysis (category arguments, uniform boundedness
principles, the closed graph, open mapping, and Hahn-Banach theorems)
are dealt with in Appendixes A and B. The basic terminology of linear
algebra is used, but no result of any depth is assumed.
Exercises appear at the end of each chapter, the more difficult ones being
provided with hints to their solutions.
The bibliography, which refers to both book and periodical literature,
contains many suggestions for further reading in almost all relevant directions
and also a sample of relevant research papers that have appeared since the
publication of the works by Zygmund, Bary, and Rudin already cited.
Occasionally, the text contains references to reviews of periodical literature.
My first acknowledgment is to thank Professors Hanna Neumann and
Edwin Hewitt for encouragement to begin the book, the former also for the
opportunity to tryout early drafts of Volume I on undergraduate students
in the School of General Studies of the Australian National University, and
the latter also for continued encouragement and advice. My thanks are due
also to the aforesaid students for corrections to the early drafts.
In respect to the technical side of composition, I am extremely grateful
to my colleague, Dr. Garth Gaudry, who read the entire typescript (apart
from last-minute changes) with meticulous care, made innumerable valuable
suggestions and vital corrections, and frequently dragged me from the
brink of disaster. Beside this, the compilation of Sections 13.7 and 13.8
and Subsection 13.9.1 is due entirely to him. Since, however, we did not
always agree on minor points of presentation, I alone must take the blame
for shortcomings of this nature. To him I extend my warmest thanks.
My thanks are offered to Mrs. Avis Debnam, Mrs. K. Sumeghy, and Mrs.
Gail Liddell for their joint labors on the typescript.
Finally, I am deeply in debt to my wife for all her help with the proof-
reading and her unfailing encouragement.
R.E.E.
CA.NBERRA, 1967
CONTENTS
Exercises 180
xii CONTENTS
Bibliography 207
Research Publications 213
Symbols 21a
Index 221
CHAPTER 1
Y2a o + (1.1.1)
,,=1
(1.1.1*)
a" = -1
1T
f"
_"
f(x) cos nx dx, b" = -1 fn f(x)osin nx dx,
1T -n
(1.1.2)
c" =~
21T
fn
_II
f(x)e-'''X dx, (1.1.2*)
the a" and b" being now universally known as the "real," and the ell as the
"complex," Fourier coefficients of the function f (which is tacitly assumed to
be integrable over (-1T, 1T. The formulae (1.1.2) were, however, known
earlier to Euler and Lagrange.
1
2 TRIGONOMETRIC SERIES AND FOURIER SERIES
converges everywhere; but, as will be seen in Exercise 7.7 and again in 10.1.6,
it is not the Fourier series of any function that is (Lebesgue-)integrable over
( -1T, 1T).
The theory of trigonometric series in general has come to involve itself
with many questions that simply do not arise for Fourier series. For the
express purpose of attacking such questions, many techniques have been
evolved which are largely irrelevant to the study of Fourier series. It thus
comes about that Fourier series may in fact be studied quite effectively
without reference to general trigonometric series, and this is the course to be
adopted in this book.
The remaining sections of this chapter are devoted to showing that, while
Fourier series have their limitations, general trigonometric series have others
no less serious; and that there a.re well-defined senses a.nd contexts in which
Fourier series are the natural and distinguished tools for representing functions
in useful ways. Any reader who is prepared to accept without question the
restriction of attention to Fourier series can pass from 1.1.3 to the exercises
at the end of this chapter.
[1.2] POINTWISE REPRESENTATION OF FUNCTIONS 3
2~ L =
(m # n, m ~ 0, n 0)
~ {~
~
LL = ~ {r.
(m # n, m ~ 0, n ~ 0)
'in ,in nx dx (m=n>O),
(m = n = 0)
(1.1.3)
-2
1
1T
I"_"
cos mx sin nx dx = 0,
-1 I"_" e
21T
1m % e -1,,% dx = {O1 (m # n)
(m = n);
in these relations m and n denote integers, and the interval [ -1T, 1T] may be
replaced by any other interval of length 21T.
Thus we shall say that the series (1.1.1) is summable at a point x to the
function f if and only if
lim aN(x) = f(x).
N_oo
in 1916, there exist trigonometric series which converge to zero almost every-
where and which nevertheless have at least one nonvanishing coefficient;
see 12.12.8. (That this can happen came as a considerable surprise to the
mathematical community.)
lim
N-",
J"
_"
IVNI dx = 0;
1 + IVNI
and the circuitous phrasing is necessary because f may take infinite values
on a set of positive measure.) This sense of representation is weaker than
pointwise representation. For more details see [Ba2]' Chapter XV.
These theorems of Men'shov and Bary lie very deep and represent enormous
achievements. However, as has been indicated at the end of 1.2.2, the
representations whose existence they postulate are by no means unique.
Cantor succeeded in showing that a representation at all points by a
convergent trigonometric series is necessarily unique, if it exists at all.
Unfortunately, only relatively few functions f admit such a representation:
for instance, there are continuous periodic functions f that admit no such
representation. (This follows on combining a theorem due to du Bois-Reymond
and Lebesgue, which appears on p. 202 of [Bad, with results about Fourier
series dealt with in Chapter 10 of this book.) It is indeed the case that, in a
sense, "most" continuous functions admit no representation of this sort.
[1.3] NEW IDEAS ABOUT REPRESENTATION 7
The relationships referred to are (compare 6.1.1, 6.2.6, 12.5.3, and 12.10.1):
(B) lim
N-CX)
J"
-n:
If(x) - SN(X)jP dx = 0;
Each of the relations (A) to (D) can, therefore, be used to characterize the
Fourier series of f under the stated conditions, and each provides some
justification for singling out the Fourier series for study. (There are, by the
way, numerous other relationships that might be added to the list.)
It turns out that the weakest relationship (D) is suggestive of fruitful
generalizations of the concept of Fourier series of such a type that the
distinction between Fourier series and trigonometric series largely disap-
pears. It suggests in fact the introduction of so-called distributions or
generalized functions in the manner first done by L. Schwartz [Sl.2]' It will
then appear that any trigonometric series in which Cn = O(lnl") for some
k may be regarded as the Fourier series of a distribution, to which this
series is distributionally convergent. These matters will be dealt with in
Chapter 12.
1.3.4. Summary. The substance of Section 1.2 and 1.3.3 summarizes the
justification for subsequent concentration of attention on Fourier series in
particular, at least insofar as reference is restricted to harmonic analysis in
its classical setting. We shall soon em bark on a program that will include at
appropriate points a verification of each of the unproved statements upon
which this justification is based. As for trigonometric series in general, we
shall do no more than pause occasionally to mention a few of the simpler
results that demand no special techniques.
A bird's-eye view of-many of the topics to be discussed at some length in
this book is provided by the survey article G. Weiss [1].
1.3.5. Fourier Series and General Groups. There are still other reasons
in favor of the chosen policy which are based upon recent trends in analysis.
Harmonic analysis has not remained tied to the study of Fourier series of
periodic functions of a real variable; in particular it is now quite clear tha.t
Fourier-series theory has its analogue for functions defined on compact Abelian
groups (and even, to some extent, on still more genera.l groups); see, for
example, [HR], [Re], [E l ]. While the level at which this book is written
precludes a detailed treatment of such extensions, we shall make frequent
reference to modern developments. However regrettable it may seem, it is a
fact that these developments cluster arOlind the extension of precisely those
portions of the cla.ssical theory which do not depend upon the deeper properties
of p~intwise convergence and summability, and that a detailed treatment of
the analogue for compact groups of the theory of general trigonometric series
appears to lie in the future. Moreover, the portions of the classical theory that
have so far been extended appear to be those most natural for handling those
problems which are currently the center of attention in general harmonic
analysis. Of course, these prevailing features may well change with the pa.ssa.ge
of time. While they prevail, however, they add support to the view that it is
reasonable to accord some autonomy to a theory in which the modes of
representation mentioned in 1.3.2 take precedence over that of pointwise
representation.
10 TRIGONOMETRIC SERIES AND FOURIER SERIES
EXERCISES
( ) -_ ~
D NX "" 170% _ sin (N + %)x
e - . u '
1701.; N SIn /2X
FN(x) = (N + 1)-1 [Do(x) + ... + DN(x)]
= (N + 1)-1 [Sin ~(N + 1)x]2
sm%x
for N ~ 0 an integer and x ~ 0 modulo 217, where the equality signs im-
mediately following DN(x) and F N(X) are intended as definitions for all real x.
1.2. Prove that if p and q are integers and p < q, and if x ~ 0 modulo 217,
then
I L:
P.; .. .;q
eln:rl ~ Icosec %xl
By using partial summation (see 7.1.2 and [HJ, p. 97 if.) deduce that if
cp ~ CP+l ~ ~ Cq ~ 0, then, for x ~ 0 modulo 217,
L:
p.;n.;q
cneln%1 ~ cp Icosec %xl
1.3. Assume that Cn ~ Cn + 1 and lim .. _<Xl Cn = O. Show that the series
1L:
N
Cn sin nxl ~ A(lT + 1).
70=1
Hints: One may assume 0 < x < 17. Put m = min (N, [IT/XJ) and split
the sum into 2:i' + 2:~ + 1> an empty sum being counted zero. Estimate the
partial sums separately, using Exercise 1.2 for 2~+l'
1.5. Assume that the Cn are as in Exercise 1.4. Show that the series
2::'= 1 c.. sin nx is boundedly convergent, and that the sum function is con-
tinuous, except perhaps at the points x =
0 modulo 217'. (More general
results will appear in Chapter 7.)
EXERCISES 11
f(x) = 2: cneilll:,
Inl .. N
vanishes for all but a finite number of integers nand thatf(x) = Lnezj(n)einl:.
Show also that
1
271' J"
_II If(x)i2 dx = .
nfz If(nW
whenever f is a trigonometric polynomial. (This is a special case of Parseval's
formula, to which we shall return in Chapter 8 and Section 10.5; see also
Remark 6.2.7.)
A trigonometric polynomial f such that j(n) = 0 for Inl > N is said to be
of degree at most N.
(2) Verify that the set TN of trigonometric polynomials of degree at most
N forms a complex linear space of dimension 2N + 1 with respect to point-
wise operations, and that if f E TN' then also Re f E TN and 1m f E TN'
(3) Show that if f E TN' f =ft 0, then f admits at most 2N zeros (counted
according to multiplicity) in the interval [0,271') (or in any interval congruent
modulo 271' to this one).
1.8. (Steckin's lemma) Suppose f E TN is real-valued, and that
Prove that
f(xo + y) ~ M cos Ny for Iyl ~ ; .
Hints: Put g(y) = f(x o + y) - M cos Ny. Assuming the assertion false,
we choose Yo so that IYol < 71'/N and g(yo) < O. We assume 0 < YO<71'/N;
otherwise the subsequent argument proceeds with the interval [-271',0) in
12 TRIGONOMETRIC SERIES AND FOURIER SERIES
Notes: Many other proofs are known; the above, due to Steckin, is perhaps
the simplest. For a proof based upon interpolation methods, see [Z2]' p. 11.
More general results, also due to Bernstein, apply to entire functions of order
one and exponential type; see [Z2], p. 277.
See also the approach in [Kz], p. 17; W. R. Bloom [1], [2]; MR 51 # 1239;
52 ## 6288, 11446; 53 # 11289; 54 # 829.
The inequality has also been extended in an entirely different way by
Privalov, who showed that if J = (a', b') and J = (a, b) are any two sub-
intervals of [-11,11] satisfying a < a' < b' < b, then there exists a number
c(I, J) such that
sup II'(x) 1~ c(J, J)N . sup 11(x)1
.%61 xeJ
for any 1 E TN' It is furthermore established that similarly (but perhaps with a
different value for c(I, J)) one has
for any 1 E TN and any p satisfying 1 ~ P < 00. Both inequalities are also valid
when I = J = [ -11,11] and c(J, J) = 1, the first reducing to that of Bernstein
and the second being in this case due to Zygmund. For more details, see [B~],
pp. 458-462. See also [L2]' Chapter 3.
The aim of the first two sections of this chapter is to show how and to
what extent the topological group structure of the set R of real numbers,
and of some of its subgroups and quotient groups, lead to the study of
periodic functions, the complex exponential functions, and the problem of
expansions in trigonometric series in general and Fourier expansions in
particular. In the remaining sections of this chapter we shall begin the study
of Fourier coefficients in some detail.
In pursuing the aims of Sections 2.1 and 2.2 we are led to refer to fairly
general topological groups and to constructs related to them. It is hoped
thus to convey a very rough idea of how the classical theory of Fourier series
fits into contemporary developments in parts of analysis, and to prepare the
reader for a later perception of genuine structural unity underlying obvious
similarities. It is of course not expected, nor is it necessary for an under-
standing of subsequent developments in this book, that the reader should
at this stage stop to gather the details concerning topological groups and the
related concepts to be spoken of (duality, invariant integrals, and so on);
this writer will indeed venture the opinion that the return to a detailed study
of generalizations is best made after some familiarity with special cases has
been attained. On the other hand the reader will, it is hoped, gain from the
realization that the classical theory is tributary to a broader stream, and will
in due course want to try his hand at exploring the latter with the help of the
references cited in this chapter.
2.1.1. The Groups Rand T. The set R of real numbers, taken with
addition as its law of composition and with its usual metric topology, is an
example of an Abelian topological group. This means that it is first an
Abelian group, and second a topological space, and that moreover the
algebra and the topology are so related that the mapping (x, y) - x - y is
continuous from R x R into R. If one drops the demand that the group
structure be Abelian, one has here the concept of a topological group in
general; see [B], pp. 98 ff., and/or [HR], Chapter II. Hereinafter the term
"group" will always mean "locally compact group whose topology satisfies
the Hausdorff separation axiom." This particular topological group R is
locally compact but noncompact. We wish to focus attention, not so much on
R, as on quotient groups thereof.
It is a simple matter to show that the only closed subgroups of R, other
than {O} and R itself, are those consisting of all integer multiples of some
nonzero positive number (see Exercise 2.1). Which of these is selected is
largely immaterial: we choose that one which is formed of all integer multiples
of 21T and which is hereinafter denoted by 21TZ (Z denoting the additive
group of integers).
Let us form the quotient group Rj21TZ = T and denote by p the natural
projection of R onto T, which assigns to x E R the coset x = x + 21TZ
containing x. The group T is made into a topological group by endowing it
with the so-called quotient topology. In concrete terms, this means that the
open sets in T are precisely the sets p( U) where U is open in R. Even more
concretely put, the quotient topology on T is that defined by the metric
d(x, iI) = inf {Ix - y + 2n1Tl: n E Z}.
Another way of looking at T is to recognize that the mapping x _ exp(ix)
is an isomorphism of T onto the multiplicative group of complex numbers
having unit absolute value. In this isomorphism, the quotient topology
corresponds to that induced on the unit circumference in the complex plane
by the usual metric topology on the latter. In view of this, the group T is
often referred to as the circle group or the one-dimensional torus group.
Perhaps the most essential difference between Rand T is that the latter is
compact. Were we to attempt to apply to R the subsequent considerations
concerning T, we should be led to Fourier integrals in place of Fourier series;
almost a.ll the additional difficulties thereby encountered would stem from
the fact that R is noncompact.
!'S correspond to continuous j's. It will in fact be the case that this corre.
spondence preserves every structure relevant to our purpose, and we shall
before long ask the reader to make a mental identification of / and f.
It is also perfectly legitimate to regard functions on the circle group as
functions of the complex variable z = el:r having unit absolute value, but we
shall make no systematic use of this notation.
2.1.3. Role of the Group Structure. As we shall see in Section 2.2, the
topological group structure of T is inextricably bound up with the genesis
and study of Fourier series. Indeed, it will slowly emerge that many of the
most fundamental aspects of this study depend almost exclusively on the
fact that T is a compact Abelian group. It will be seen, too, that the Lebesgue
integral itself is determined (up to a nonzero constant factor of proportionality)
by the topological group structure.
To this basic ingredient may be added, for the sake of richness and
refinement, more specialized structures and concepts-the concepts of
bounded variation and differentiability for functions, for example. In line
with the remarks in 2.1.2, we say that a function / on T is of class Ck (=k
times continuously differentiable, or indefinitely differentiable if k = CX),
or is of bounded variation, if and only if the function / 0 p on R has the
corresponding property on some one (and therefore every) interval in R of
length 27T.
To = I, (2.2.1)
where I denotes the identity automorphism of C.
In general, and certainly for the groups Rand T here considered, the space
C is infinite-dimensional and the problem of analyzing the behavior of the
operators T a on C is a complicated one. However, elementary linear algebra
(and, even more so, suitable forms of the simultaneous spectral resolution
theorem) encourage one to hope for simplification if one can "reduce" the
problem by finding linear subspaces V of C which are invariant in the sense
that T a(V) c V for all group elements a. For brevity we term such a V an
invariant 8ub8pace. The hope would lie in decomposing C into some sort of
(possibly infinite) direct sum of invariant subspaces V1 , V2 , " , each VI
being as small as possible. The T a could then be examined on each VI
separately.
In this way one is led to consider the existence of minimal invariant sub-
spaces V of C, "minimal" meaning that V contains properly no invariant sub-
space other than {O}. Now it is evident that a onedimensional invariant
subspace V (if any such there be) is certainly minimal; and that such a
subspace V is generated or spanned by a function f which is a simultaneous
eigenvector of the Ta (if any such functions exist). So, without more ado, we
seek such functions. (For nonAbelian groups in general there would not
exist anyone-dimensional invariant subspaces-one would have to be content
with seeking finite-dimensional ones, which in fact exist in abundance for
compact groups; for noncompact, non-Abelian groups, the situation is even
more complicated.)
Given fEe, denote by V, the smallest invariant subspace containing f,
that is, the set of all finite linear combinations of translates Taf of f. We
seek functions f such that dim V, = l. Clearly, therefore, f "# 0 and to each
group element a corresponds a complex scalar x( -a) such that
Taf = x( -a)f.
This signifies that
f(x - a) = x( -a) f(x) (2.2.2)
for all pairs (a, x). If x = 0, f( -a) = f(O) x( -a), which shows in particular
18 GROUP STRUCTURE AND FOURIER SERIES
that X is continuous and that X -# O. On the other hand, (2.2.1) and (2.2.2)
yield the functional equation
Since X is continuous and X(O) = 1, h may be chosen and fixed so that the
factor
J: X(b) db
is nonzero. Moreover,
Jo
r x(a + b) db = Jara+ll X(e) de.
ll
1 x(h) - x(O) ()
=lm
1l .... 0
h 'x a ,
X' = ik X, (2.2.4)
[2.2] TRANSLATES OF FUNCTIONS 19
where k = -iX'(O). The only solution of (2.2.4) taking the value 1 at the
origin is
x(x) = etkx (2.2.5)
Evidently, whatever the complex number k, (2.2.5) defines a continuous
character of R. This character is bounded, if and onJy if k is real.
To determine the continuous characters of T, it is merely necessary to add
the demand that X have period 27T. This signifies that k E Z.
To sum up, we find that
(1) The continuous (and so necessarily bounded) characters of T are in
one-to-one correspondence with Z, the character corresponding to 11, E Z
being (derived by passage to the quotient from) the function
(2.2.6)
Corresponding to 11, = 0 is the character eo, which is the constant function 1;
this is usually termed the principal character.
(2) The one-dimensional invariant subspaces of CrT) are precisely the
subspaces V" = {"e,,: " a scalar}, where n ranges over Z.
(3) The problem of harmonic analysis in respect to CrT) (and similarly in
respect of other function spaces) may be suggestively but perhaps oversimply
described as that of expressing C( T) as some sort of direct sum of the
subspaces V" (11, E Z). This task falls into two parts:
(a) GivenfE CrT), it is required to determine the corresponding "com-
ponents " off lying in the various subspaces V". This is, strictly speaking, the
problem of harmonic (or spectral) analysis and is, in the case of compact
Abelian groups anyway, relatively simple. The said components are just the
functions J(n)e", where
It will appear in Chapter 11 that the component j(n)e n is nonzero, if and only
if Vn n V of: {O}, where V is the closed invariant subspace generated by f.
(b) The study of the formula
f = L J(n)e
nEZ
n,
which it is hoped will reconstitute f from its harmonic components. This may
be described as the problem of the harmonic (or spectral) synthesis of f. It
presents what is by far the more difficult part of the program and embraces,
of course, the question of representing f by a trigonometric series. It must
be stressed that such a series representation is indeed generally impossible in
C, if one dema.nds pointwise convergence. The study of the sense in which
the synthesis is valid (which will vary from one function space to another)
20 GROUP STRUCTURE AND FOURIER SERIES
is an essential part of the problem before us; see the remarks in 10.3.6 and
Section 16.8.
In connection with (1) above, it is interesting to obscrve that the group
structure of Z corresponds, when Z is used to label the characters en, to point-
wise multiplication of characters. Moreover, thc corresponding "dual topology"
on Z is that for which the relation n -- no signifies that
en(x) ->- eno (x)
uniformly for x E T, and turns out to be just the discrete topology on Z (having
a base of neighborhoods of 0 E Z comprising the one set {O}). This is a general
feature: the bounded continuous characters of any given group may them-
selves be formed into a group under pointwise multiplication, termed the dual
or character group of the given group, and topologized in such a way that
(speaking informally) a sequence or net (xtl of characters converges to the
character X if and only if lim, X,(x) = X(x) uniformly for x E K, and this for
each compact subset K of the original group. Up to this point, everything is
largely a matter of observation and definition. The interesting and decidedly
nontrivial fact is that, by way of justification of the term "duality," the
character group of the character group turns out to be (isomorphic with) the
original group. This duality is profound and is fundamental in general harmonic
analysis, but to develop the ideas in any generality would take us much too
far afield. Suffice it to say that locally compact Abelian groups run around in
mutually dual pairs-such as (R, R) and (T, Z)-either member of such 'l pair
being isomorphic with the dual of the other: this is the so-called Pontryagin
duality law, for more details of which the reader is referred to [B), Chapter 11,
and [RR], Chapter VI. Our main concern will always be harmonic analysis on
the group T, but we shall from time to time cast fleeting glances at the dual
problems concerning harmonic analysis on the group Z, which is always to be
thought of as being endowed with its discrete topology. To the reader we issue
a standing invitation to reflect on the possible analogues for Z of results
established in the text for T. As a start, he might verify that, in conformity
with the Pontryagin duality law, the character group of Z can be identified
with T in the manner suggested by (2.2.6); that is, to each bounded (necessarily
continuous) character ~ of Z corresponds exactly one x E T such that ~(n) =
en (x) for n E Z and that the initial topology on T corresponds exactly to its
dual topology under the association x ~ ~:r (see Exercise 2.3).
Studies of harmonic analysis on each of the groups T and Z form, when
taken together, a useful forerunner to that of general harmonic analysis. This
is partly because they illustrate separately a number of the difficulties that
one encounters in an intermingled state when one moves along to harmonic
analysis on general groups. Actually, the next degree of complexity is repre-
sented by the group R (the additive group of real numbers with its usual
topology). In T, R, and Z one has, so to speak, the natural building bricks
from which quite general locally compact Abelian groups may be built up_ It
is known, for example, that any such compactly generated group is isomorphic
with a product Ra x Zb X F, where a and b are nonnegative integers and F is
[2.2] TRANSLATES OF FUNCTIONS 21
for example: [HR], Chapters III and IV; [HS], Chapter 3; [B], Chapters
8-10; [E], Chapter 4. However, an intelligent reading of the present book
will demand no more than a knowledge of the results of this extension for
functions of one real variable; it is of little importance which of several
possible approaches to the Lebesgue integral has been followed. More details
about what we shall need to assume appear in 2.2.4.
The choice of Cc(G) in place of C(G) as the initial domain of definition of I
comes about in the following way. It is quite easy to see at the outset that,
whenever Gis noncompact, there cannot be any invariant integral I for which
IU) is finite-valued for all nonnegative real-valued 1 E C(G) (or even for all
nonnegative real-valued 1 E C(G) which tend to zero at infinity). In other
words, the "integrability" of a function will demand quite severe restrictions
on the "average smallness" of the function at infinity. One v~ry simple and,
as it turns out, entirely effective way of imposing a priori such a restriction
on 1 is to demand that it shall vanish outside some compact subset of G.
(Of course, it turns out ultimately that this condition is not necessary for
integrability.)
It is not too much to say that the inauguration of modern harmonic
analysis on groups had to await the discovery, by Haar in 1933, of the
existence of a left invariant integral on any locally com pact group G satisfying
the second countability axiom. Subsequent developments, including the
removal of all countaobility restrictions on G, have been due to Weil, Kakutani,
H. Cartan, von Neumann, and many others. The interested reader may also
wish to consult a recent note by Bredon [1]. See also MR 39 # 7066.
On considering some familiar groups, old friends appear in a new light.
For example, if G = Rn, the characteristic properties (1) and (2) show that an
invariant integral is
functions on R (see 2.1.2). Let us first note that any invariant integral I has
the property that
II(!)I ~ I(l} . sup If I ; (2.2.8)
this follows from property (1) and the linearity of I. Now, iffis continuously
differentiable,
lim Taf - f = -1'
a-O a
holds uniformly. Consequently (2.2.8) and property (2) combine to show that
l(j') = 0 for any continuously differentiable f. Next, if g is continuous and
periodic and satisfies
[2"
Jo g(x) dx = 0,
then g = 1', where
f(x) = 50" g(t) dt
is continuously differentiable and periodic. Thus I(g) = 0 for such g. Finally,
choose any nonnegative continuous periodic ho such that
fIt ho(x) dx = 1.
Given any continuous periodic h, we apply what precedes to the function g
defined by
g(x) = h(x) - ho(x) Jo[2" h(t) dt,
which is continuous and periodic, and satisfies
f" g(x) dx = O.
showing that I differs from the expression appearing on the right-hand side
of (2.2.7) by the oonstant faotor 217I(h o) > O. This oompletes the verification
of the essential uniqueness of the invariant integral on T.
Armed with this uniqueness property, it is simple to deduce other invariance
properties of the integral. The elementary properties of the Riemann integral
show that, if f is a oontinuous periodic funotion on R, and if k E Z and
k =1= O,then
1 [211 1 [2"
217 Jo f(kx) dx = 217 Jo f(x) dx. (2.2.9)
24 GROUP STRUCTURE AND FOURIER SERIES
The dual situation. Let us turn momentarily from the circle group T
to the dual group Z. There is no lasting mystery about the invariant integral
on Z; apart from a disposable constant of proportionality, it must be ex-
pressed by summation:
I(t/ = 2 t/>(n),
nsZ
(2.2.7)
L cp(n)
Inl .. N
when N _ 00. A yet more general interpretation which will playa funda-
mental role in the sequel lies in interpreting the right. hand side of (2.2.7*) as
the limit as N -00, when it exists, of the arithmetic means of the first N + 1
symmetric partial sums. This arithmetic mean is expressible in the form
L (1 - ).n+1 1) >(n),
Inl .. N
if Xl = X2
(2.2.12)
otherwise.
Le
nEZ
lnx eln~ =1 or 0
according as the real numbers x and yare, or are not, congruent modulo 21T.
There is, however, no way of making sense of this relation which is based upon
applying a summability method to the series on the left for individual values
of x and y. On using concepts to be introduced in Chapter 12, it is nevertheless
the case that, for a fixed y, the series converges distributionally to a certain
distribution (or generalized function) known as the Dirac measure placed at the
point y. This latter entity does, in a sense, vanish on the open set of points
x ;f:. y modulo 21T, but there is no reasonable way of attaching to it a numerical
value at points x == y modulo 21T.
There is therefore a residual and irreducible asymmetry separating the
mutually dual situations; this is, in the last analysis, because of the profound
topological differences between the "smooth" compact group G and the
discrete noncompact group Z.
2.2.4. V' and Other Function Spaces. It has been remarked in 2.2.2
that the invariant integral can in all cases be extended to functions more
general than those in Cc(G). For G = T, in which case the invariant integral
has been identified in (2.2.7), the extension involved is that from the Riemann
to the Lebesgue integral; for the dual group Z, several stages in the extension
have already been mentioned at the end of 2.2.2. It is essential for a smooth
and satisfactory development of Fourier theory that advantage be taken of
this extension. Broadly and figuratively speaking, the Lebesgue theory of
integration is that which is necessary and sufficient for the major portion of
contemporary analysis; integration theories for functions on more general
sets and spaces almost invariably share the characteristic basic properties of
the Lebesgue theory. However, in certain special connections involving
functions of a real variable, more elaborate theories have proved useful. We
shall have neither occasion nor space for more than a passing reference in
12.8.2(3) to some such theories. (Others, mainly designed to handle integration
strictly as an antiderivation process, will receive no mention at all in this
book.)
We shall therefore assume that the reader is familiar with the definition
and basic properties of the Lebesgue integral of a function of one real
variable. With but relatively few exceptions, some of which are dealt with in
Appendix C, all the results we shall need will be found in the brief account in
[W]. For the exceptional points the reader is referred to [HS], [AB], or [E],
Chapter 4, or to anyone of the several excellent accounts of integration
theory now available. In making use of these sources of results about the
Lebesgue integral of functions of one real variable, it will be agreed that a
[2.2] TRANSLATES OF FUNCTIONS 27
(2.2.13)
is finite, the integral being extended over any interval of length 21T; compare
[W], p. 68, [AB], p. 215, or [HS], p. 188. In addition, L'" = L"'(T) denotes the
set of essentially bounded periodic complex-valued measurable functions,
that is, of periodic complex-valued measurable functions f for which
IIfll", == ess. sup If(x)1 (2.2.14)
is finite, the essential supremum being taken relative to any interval of x-
values of length 21T.
To be perfectly accurate, we shall frequently use V' (0 < p ~ 00) to
denote the set of equivalence cla8se8 of the appropriate type, two functions
going into the same class if and only if they agree almost everywhere (a.e.).
Since we shall not always signal which viewpoint is being adopted, the reader
is warned to be on his guard and to be prepared to devote a little thought to
deciding which interpretation is appropriate. The Fourier series of a function
depends only on the class determined by that function.
Each of Ck (k an integer ~O, or 00) and V' (0 < p ~ 00) is a linear space;
in view of preceding remarks, the reader should check the truth of this when
LP is regarded as a set of equivalence classes.
When 1 ~ p ~ 00, I lip is a norm on LP if the latter is considered as a. set
of equivalence classes of functions (but only a seminorm if LP is viewed as a
set of individual functions); see Appendix B.1.2 for an explanation of the
terminology. This statement is virtually the content of Minkowski 's inequality,
which asserts that f + g E LP and
(2.2.15)
whenever 1 ~ p ~ 00 and f, g E LP. For a proof of Minkowski's inequality,
see p. 68 of[W], or p. 146 of [HLP], or Section 4.11 of [E], or [AB], p. 218, or
28 GROUP STRUCTURE AND FOURIER SERIES
[HS], pp. 191-192. The assertion is false if 0 < p < 1 (see [HLP], loco cit.),
but it is then true that Ilf - gil: is a metric on Lp qua set of equivalence
classes (or a semimetric if 1P is considered as a set of individual functions).
For 0 < p ~ 00, 1P is complete for the metric Ilf - gllp if p ~ 1, or for the
metric Ilf - gil: if 0 < p < 1; the former case is dealt with in [W], Theorem
4.5a, and the same argument adapts readily to the case 0 < p < 1; alterna-
tively, see [HS], p. 192, or [AB], p. 220.
To complete the picture, on C" (k an integer ~ 0) we introduce the norm
Ilfll(,,) = sup II DhJll", , (2.2.16)
O"i h"i "
Despite the notation, Ilfll(",) is not a norm. Then Ck is complete for the metric
Ilf - gll(k) whenever k = 0,1,2, . ',00 (the reader should supply a proof of
this).
With their appropriate metric topologies, all these spaces are topological
linear spaces (see Appendix B.l.1), that is (compare 2.l.1 in relation to
topological groups), the linear space operations (j, g) -+ f - g and (A, f)
-+ A f (A a complex scalar) are continuous. Further details concerning C'"
appear in Section 12.l.
There will be constant use for one or more links in the chain of inclusion
relations
C'" c ... C C"+l C C" c ... c Co = C c L'" c LP C Lq, (2.2.18)
where k is an integer ~O and where 00 > p > q > O. What is more, each
inclusion map of one term of this sequence into any other lying to its right
is continuous. The only nontrivial portion of this last assertion depends on
the inequality
O<q<p~oo, (2.2.19)
The estimate (2.2.19) is itself a consequence of Holder's inequality, which
asserts that if 1 ~ p ~ 00, and if p' denotes the conjugate exponent (or index)
defined by IIp + IIp' = 1 (supplemented by the convention that p' = 00
if p = 1 and p' = 1 if P = 00), then f . g E Ll and
(2.2.20)
whenever f E Lp and g E Lp. A proof of Holder's inequality will be found on
pp. 72-73 of [Ka], the assumption there made concerning continuity of f
and g being unnecessary; see also Section 4.11 of [E], or [HS], pp. 190-191, or
[AB], p. 217. An extended discussion of both the Minkowski and Holder
[2.2] TRANSLATES OF FUNCTIONS 29
11"'111' == { 2: 1",(n)JPp'"
neZ
ifO<p<oo,
or
11"'11", == sup
nFZ
''''(n)1 if p = 00,
is finite.
In addition to these, we occasionally wish to refer to the subspace Co = co(Z)
of I""(Z) formed of those", for which
lim "'(n) = O.
Inl-""
Each of Co and II' (1 ::;; p ::;; 00) is a Banach space; if 0 < p < 1, II' is a
complete metric space.
In lieu of (2.2.18) and (2.2.19) one has the relations
(2.2.18*)
and
(2.2.19*)
for 0 < q < p < 00. (Notice that 11"'l q ::;; 1 implies 1",(n)1 ::;; 1 for all nEZ,
hence 1"'(n)IP ::;; 1",(n)lq for all nEZ, hence L 1""1' ::;; L ''''Iq ::;; 1.)
30 GROUP STRUCTURE AND FOURIER SERIES
The Holder and Minkowski inequalities suffer no change in form other than
the obvious replacement of integrals by the appropriate sums. Proofs for
the case of finite sums appear on pp. 67-72 of [Ka]; for our purposes, which
involve infinite sums, transparent limiting processes constitute the final step;
see [HS], p. 194. A much more elaborate account appears in Chapter II of
[HLP].
Concerning notation, we shall sometimes denote a function cP on Z in the
sequential form: (CPn)nez; this is sometimes a convenience and is in any case
in accord with tradition. There is, however, a nonvanishing chance of con-
fusion with the convention according to which (CPn)"e z might also denote a
two-way infinite sequence of functions on Z. The context will in all cases
dispel initial doubts on this score.
ment with the traditional one referred to here. For any complex-valued
function I, J denotes the complex-conjugate function. For any function I
defined on any group (on R, T, or Z in particular), j denotes the function
t---+/(-t), andf* the function t---+/(-t); thusf* = (J)~ = (j)-. Accord-
ingly D and I ---+ I are linear, whereas I ---+ J and I ---+ f* are conjugate-linear.
for all n E Z.
Note: In 3.1.1 and 4.1.2 we shall add some most important complements
to the first assertion in 2.3.1.
.
I/(n)1 1
~ 21T f I/(x)e-I""'1 dx 1
= 21T f I/(x)1 dx
Illk
==
Note: If we write II/II", = sup {1/(n)1 : nEZ}, 2.3.2 is equivalent to the
inequality II/II", ~ 1IIIk This estimate, although well-nigh trivial, is the best
possible in the sense that /(0) = IIIIII whenever I is real and nonnegative.
On the other hand, for general real- or complex-valued functions I the
relationship between II/II", and 11/111 is complicated; see Exercise 8.8 and
Subsections 11.3.1 and 11.4.14. In particular, there exist functions IE LI
for which IIIIII > 0 and the ratio 11/11",/11/111 is arbitrarily small. Hosts of
examples of this phenomenon can be constructed by using the results of
Chapter 15. A simpler example is provided by the so-called Dirichlet kernel
( ) -_ L
D NX '" I"" _ sin (N + 1/2)x.
e -. ,
I"I"N smx/2
it has been shown (D. J. Newman [2]) that for each positive integer N there
exists a trigonometric polynomial
2: elle
N
f(x) = 1nz
II~O
f f
obtain
.
f(n) = 217
1 g(x) dx = 217
1 T aU(x) dx
= elno. 2~ f To.f(x) e- In % dx
= elno. (To.ft'(n) ,
which is equivalent to the stated result.
Note: On being asked for a proof of 2.3.3, the reader's first reaction might
be to apply the usual formula for change of variable in the integrals involved:
this procedure is, of course, perfectly legitimate. But we prefer to phrase the
device in terms of the characteristic in variance property of the integral
(see 2.2.2).
= /(n) in,
which completes the proof.
[2.3] FOURIER COEFFICIENTS, THEIR ELEMENTARY PROPERTIES 33
Remarks. All that is required ofDj is that the partial integration formula
shall hold for all periodic, indefinitely differentiable functions u. This means
that the preceding interpretation of Dj for absolutely continuous functions f
will accord with the generalized concept of derivation introduced in Chapter
12. The reader is reminded, however, that the result is not generally true for
functions j possessing almost everywhere an integrable derivative: it is in
addition necessary that j be equal to the indefinite integral of this derivatwe,
which is ensured by (and indeed equivalent to) absolute continuity.
L = [f(27T) - L [f(X/C+l) -
m-l
l(xl)JIl(O) - l(x/c)JIl(x/c).
/C=l
By continuity (and periodicity) of I, the first summand on the right will not
exceed e in absolute value, provided the partition is sufficiently fine. Thus
1 m-l
1/(n)1 ~ e + 2: + 27T /C~l II(x/C+l) - I(x/c) I IIl(x/c) I
~ (1 + 27T1) e
1
+ 27T V(f) Tni'
1
since IIl(x)1 ~ l/Inl. Letting e ~ 0, we obtain
1/(n)1 ~ V(f)
27Tlnl
for n :F 0, which is equivalent to the stated result.
Suppose finally that I is merely of bounded variation. We shall obtain the
desired result in this case by approximating! by a suitable sequence of
continuous functions Ir of bounded variation. Perhaps the simplest choice is
I,(x) = f
r:r
:r+(1/')
I(t) dt = r
(lir
Jo I(x + t) dt.
i
Whatever the increasing finite sequence (x/c) of points of [0, 27T], one has
ll,
L/c I!r(X/c) - Ir(x/C-l)1 ~ r
0
L/c I!(x/c + t) - !(X/C-l + t)1 dt,
which, since the integrand never exceeds V(f), is majorized by V(f). Thus
V(f,) ~ V(f) for all r. By what is already established, therefore, we have
Ii, (n)1 ~ V(fr) ~ V(f) (2.3.2)
r 27Tl n l '" 27Tl n l
for n :F 0 and all r.
[2.3] FOURIER COEFFICIENTS, THEIR ELEMENTARY PROPERTIES 35
J, n
A ( )
= exp (in) ) sin (n/2r)
- . fA( n'
r 2r n/2r
so that
lim Ilr(n)! = Il(n)!.
r-oo
l",
lim N ~ Inf(n)!
A
= 0;
N_oo I"I"N
i 2l1(k+llll1l- 1
2kl1l"l-l
e-I":t dx = 0,
it follows that
where V" is the total variation off on the interval [ak _ 1> akJ. Since V 1 + ...
+ V" ~ V(f) and a" - ak-l = 27Tlnl-1,
27Tlj(nli ~ V(f).27Tl nl- 1
and hence
Inj(nli ~ V(f).
See also M. and S.I. Izumi [1].
2.3.7. Define the mean modulus of continuity of f with exponent (or index)
1 by
Then, if fELl,
.
If(n)1 1 wd (TT)
~ 2 n (n E Z, n - 0).
Proof. By definition
.
f(n) = 2TT
1 If(x)e-1,,;r dx,
and by 2.3.3
Proof. This follows immediately from 2.3.7 and the fact ([W], Theorem
4.3c) that wd(a) -+ 0 as a -+ O.
Remarks. (1) The Riemann-Lebesgue lemma is so fundamental that it
is worth pointing out another method of proof (which indeed lies behind the
proof of Williamson's Theorem 4.3c just cited and used). Suppose we denote
by E the set of integrable functions f for which the statement of the lemma
holds. Then 2.3.2 shows that E is a closed subset ofLl (relative to the topology
defined by the norm II Ill)' It is otherwise evident that E is a linear subspace
of Ll. To prove the lemma it therefor~ suffices to show that E contains a.
set of functions, say S, the finite linear combinations of which are dense in U.
There are many such sets S which may be indicated. Examples are: (i) the
[2.3] FOURIER COEFFICIENTS, THEIR ELEMENTARY PROPERTIES 37
set of characteristic functions of intervals [a, b] (0 < a < b < 271"), extended
by periodicity. The finite linear combinations of these are dense in Ll (as is
shown in [W], Theorem 4.3a); and each such function is directly verifiable
to have a Fourier transform satisfying the lemma (assurance on this point
also comes from 2.3.6). (ii) the set C"'; see [W], Theorem 4.3b and 2.3.4.
(2) It is worth pointing out that 2.3.4 to 2.3.8 are all essentially concerned
with restrictions on the rate of decay ofi(n) as Inl ~OO. The indications are
clearly that the smoother the function j, the more rapid this decay. This
conclusion will receive further reinforcement as we progress; some extreme
instances are covered by Exercises 2.7 and 2.8.
2.3.9. Introduction of A(Z). The preceding results and remarks might raise
hopes that the membership ofjto various function spaces (such as C or V' for
various values of p) might be decidable solely by inspection of the rate of
decay ofi, or at any rate by examining Iii. However, while there are many
criteria of this sort that are either sufficient or necessary, with the sole
exception of the case of L2 (dealt with in Chapter 8), there are no known
necessary and sufficient conditions of this type. Moreover, it will appear in
Chapters 12 and 14 that there definitely cannot be any such complete
characterization involving only the values of Iil. The few necessary and
sufficient conditions that are known are of a much more complicated sort
and are unfortunately extremely difficult to apply in specific instances; see
2.3.10. Much remains to be discovered in this direction.
To make things more specific, let us consider Ll itself. If we denote by
co( Z) the linear space of complex- valued functions (two-way infinite sequences)
<fo on Z for which lim,n, ... ",<fo(n) = 0 and equip it with the norm
for
Then
j(x) =
k=l
for n = N" (k = 1,2, .). Furthermore we shall see in Section 7.4 that
(again for any assigned q, E co(Z a function IE Ll can be chosen so
that
l(n) ~ Iq,(n)l
for all n E Z.
Again, although the sequence q, defined by
for Inl ~ 2
otherwise
belongs to A(Z), the sequence q,l defined by
for Inl ~ 2
otherwise
has not this property (see Exercise 7.7 and 10.1.6). This shows that an orderly,
and therefore seemingly harmless, change of sign can destroy membership
of A(Z).
This (or any other similar) example shows incidentally that A(Z) is a
proper subset of co(Z); it also shows that 1<p1 may belong to A(Z) while <p fails
to do so. There is an entirely different, and tYPically modern, approach to
the proper inclusion relation A( Z) ~ co( Z) which shows a little more,
namely, that A(Z) is in fact a meager (that is, first category; see Appendix
A.l) subset of co(Z).
To see this, we must observe that Ll and co(Z) are Banach spaces when
endowed with the norms defined in (2.2.13) and (2.3.3), respectively, and
that T: 1-1 is a continuous linear operator mapping Ll into co(Z) whose
range is A(Z) (see 2.3.1 and 2.3.2). If, contrary to our assertion, A(Z) were
nonmeager in co(Z), the open mapping theorem (Appendix B.3.2) would
entail that T is an open map of Ll onto co(Z). Assuming for the moment the
uniqueness theorem (2.4.1), this would imply the existence of a number
B > 0 such that
11/111 ~ B' 11111.., (2.3.4)
for eachl ELl. However, (2.3.4) can be negatived, the resulting contradiction
thus establishing our assertion. For example, if I = D N , as in Exercise 1.1, a
direct computation, which will be carried out in detail in 5.1.1, shows that
lim 2: 4>nUT(n)
r-+oo nez
= 0
lim IluTllp = o.
T- '"
for all trigonometric polynomials u; see R. E. Edwards [1] and Ryan [1].
The case p = 2 of this result is due to Sall'm; sec [Bad, pp; 239-240.
(2) In order that 4> shall be the sequence of Fourier coefficients of a con-
tinuous function, it is necessary and sufficient that to each e > 0 shall corre-
spond a number k(e) ;;il: 0 and a finite subset F. of Z such that
(2.3.7)
40 GROUP STRUCTURE AND FOURIER SERIES
p' being defined by lip + lip' = 1. It is known (Rooney [1]) that if cfo satisfies
(2.3.7), then it is the sequence of Fourier coefficients of some function in
V(l < p E> 2) if and only if
suP. (v + l)P-1 2:
maO
IM . m(4)W < 00, (2.3.8)
the supremum being taken as 1/ ranges over all nonnegative integers, where
and
an,Y.m = YCm i1o
tm(l- t)-m e n1nl dt
then
f(O) = O. (2.4.2)
We will in fact show that the negation of {2.4.2} implies the negation of(2.4.1}.
H (2.4.2) is false, we may (by changing f into - f if necessary) assume that
f(O} = c > 0 and then choose 8 > 0 so that
f(x} ~ Yzc for Ixl ~ 8 (2.4.3)
[2.4] THE UNIQUENESS THEOREM, THE DENSITY OF POLYNOMIALS 41
2117 fit dx ~ 21 Jr
17 Ixl-' 6
It dx - 21 11/11> (217 - 28)
17
~ Yz c
12 r
17 Jlxl-.6
t dx - 11111>
~ 4c
17
r
Jlxl-. %6
t dx - 11/11>
C
~ 4TT. 8 qN - 11/11>,
which is positive provided we choose Nlarger than log (417 11/11",/c8)/log q, thus
negating (2.4.1). This proves statement (1).
Now assume that I is as in statement (2). Define
dominated convergence, g ELl. By the same token, one has for any m E Z,
.
g(m) = -If g(x)e- lmx dx = L.
"".f(n) - If e1nx e- lmx dx
217 n"Z 217
=j(m),
Proof. (1) First take the case of C. Given e > 0, first choose g E C2 such
that
Ilf - gil"" ::;; Y2e. (2.4.6)
gl(X) = a- 1 r+ 1l
f(y) dy, g(x)=a-1J%
(%+11
gl(y)dy.
the series being convergent in C (that is, uniformly convergent). One may
therefore choose N so large that
IIg - 2:
Inl .. N
g(n)enll", ::;; Y2e. (2.4.7)
[2.5] REMARKS ON THE DUAL PROBLEMS 43
t = L
l"I<iN
g(n)e".
(2) This case follows from (I), the fact that C is everywhere dense in V'
(compare [W], Theorem 4.4e), and the inequality Ilk II" ~ Ilkllco. Thus, given
! E V' and e > 0, first choose gEe so that II! - gil" ~ Y2e, and then [by (I)] a
t E T so that Ilg - tll co ~ Y2e. Then, a fortiori, Ilg - til" ~ Y2e and so
in comparing this with (2.3.1), the reader will observe a change from e- inx to
ei"x, which is made purely on the grounds of subsequent convenience.
Although (2.5.1) makes excellent sense whenever <p E (l(Z), in which case ~ is
evidently a continuous function on T satisfying
(2.5.2)
(compare 2.3.2), it is plain that complications arise if, for example, <p is
known merely to belong to t"( Z) for some p > 1. (In the case of the group T
no analogous complications appeared, because of the compactness of T.)
One has in fact to contemplate conditional convergence and summability,
44 GROUP STRUCTURE AND FOURIER SERIES
perhaps merely for almost all x, as was heralded by the remarks at the end of
2.2.2 concerning the interpretation of the invariant integral on Z.
As a matter of fact, and as will appear in the course of Chapters 8, 12, and
13, it is often more effective to replace considerations of pointwise convergence
(everywhere or almost everywhere) of the series on the right of (2.5.1) by
that of convergence, either in one of the spaces Lq or distributionally, of the
symmetric partial sums
~N(X) == L .p(n)e1n %.
Inl"N
On the other hand, for functions .p which are structurally special, the results
of Chapter 7 yield pointwise convergence (at least almost everywhere) of the
series defining ~, although even here there is no assurance that the function ~
so defined almost everywhere will belong to Ll. Further very special results
of this sort, applying to cases in which .p is known to be of the form J for
some fELl, are contained in Chapters 5, 6, and 10.
2.5.3. The Space A. Dual to A(Z) is A = A(T): this is the linear space of
functions on T of the form ~ obtained when .p ranges over {l(Z). Equiva-
lently, A consists precisely of those continuous functions f on T such that
JE {l. As in the case of A(Z), so with A: there is no complete solution to the
problem of characterizing directly in terms of their functional values the
elements of A. We shall return to the consideration of A in Sections 10.6 and
12.11, where partial results will be obtained and used.
45
EXERCISES
2.1. Let S be a closed subgroup of R distinct from {OJ and R. Show that
there exists a number d > 0 such that S consists precisely of all integer
multiples of d.
Hint: Consider the infimum of all positive members of S.
2.2. Let x be a real number such that X/7T is irrational. Show that the set
{e lnz : n E Z} is everywhere dense in the unit circumference in the complex
plane.
Remarks. A stronger result will appear in Exercise 2.15. The stated result is
a special case of Kronecker's theorem, for which see [HW], 'Chapter XXIII,
especially p. 370. There is a general grouptheoretic formulation of this theorem
which is discussed in [HR], pp. 431-432, 435-436, and which asserts in particular
that any character X of R/21TZ, continuous or not, can be approximated
arbitrarily closely on any preassigned finite subset of R/21TZ by a suitably
chosen continuous character en'
Hint: Show tha.t el .,,: =1= el u if m, n E Z and m =1= n. Deduce that the said
set has I as a limiting point.
2.3. Let, be a bounded character of the group Z. Show that there exists
exactly one x E T such that {{n) = e lnz == {z{n) for all n E Z.
Assume that Z is endowed with its discrete topology; verify that the dual
topology on T (when it is regarded as the dual of Z) is identical with its initial
topology (as defined in 2.1.1).
46 GROUP STRUCTURE AND FOURIER SERIES
Hint: Make use of Stirling's formula describing the behavior of r(z) for
large positive values of z.
Remarks. Functions f for which an inequality of the type (1) holds
(the constant possibly depending uponj) form the simplest types of what are
termed q'Ua8ianalytic cla8se8 of functions; the case a = 1 corresponds to the
analytic functions. The relationships between such classes and those defined by
means of inequalities involving the Fourier coefficients-like (2). for example---
EXERCISES 47
have been studied in great detail. See, for example, [M], especially pp. 78-79,
138-139.
2.9. Give a proof of 2.3.8 based solely on 2.4.4 and 2.3.2.
2.10. Derive 2.4.1 directly from that part of 2.4.4 which refers to the
space C.
2.11. Deduce from 2.4.1 the following uniqueness theorem: If! is defined
and integrable over (-7T, 7T), and if
L L
~ ~
ex).
nal
Hints: Assume without loss of generality that the an and bn are rea1-
valued and put an = Tn cos 8n , bn = Tn sin 8n , where r n ~ 0 and 8n is real.
Use Egorov's theorem to justify termwise integration of the series
2:
~
Note: For a simple generalization of this result, see [KS], p. 84, TMoreme II.
The Lusin-Denjoy theorem has prompted numerous more elaborate investiga-
tions of the absolute convergence of trigonometric series: see [Zl], Chapter VI;
[Ba:!], Chapter IX; [KS], Chapitre VII.
exists finitely for that value of x. Show that if this is true for each point x
belonging to a measurable set E having positive Lebesgue measure, then
liml"I_'" e" = O.
Hints: As in the hints for the preceding exercise, reduce the problem to
the case in which L:r" cos (nx - 8,,) is uniformly convergent for x E Eo, where
m(Eo) > O. Were the assertion to be false, there would exist integers
n l < n2 < . . . so that
cos (n,.x - 8"k) - 0
uniformly for x E Eo. Consider the integrals
L
00
(logn)-l.cosn(x -loglogn);
.. -3
Deduce that if I is a subinterval of [0, 2fT), then N-l times the number of
points x o, 2xo, ... , Nxo which belong modulo 2fT to I converges, as N _ 00,
to (2fT) -1 times the length of I, that is, that the points nxo (n = I, 2, ... )
are equidistributed modulo 2fT. Observe that this result implies that of
Exercise 2.2. See also [Ba 2 ], p. 473.
Hints: First prove (I) for continuous periodic g by using 2.4.4.
EXERCISES 49
2.16. (Fejer's lemma) Suppose that 1 ~ P ~ 00, that IEV', and that
g E V", where lip + lip' = 1. Prove that
f
2Irr I(x)g(nx) dx -* !(O)g(O)
for n E Z and.lnl-*oo.
Hints: Assume first that p > 1 and use 2.4.4 to approximate g in Lp' by
trigonometric polynomials; then use 2.3.8. If p = 1, so that p' = 00, approxi-
mate I in L1 by continuous functions.
Note: The result actually remains true if the restriction n E Z is replaced
by n E R.
2:
co
Enk exp (in "x) ,
k=1
Convolutions of Functions
1* g(x) = 211'1' f
I(x - y)g(y) dy, (3.1.1)
then it appears that em * en has as its Fourier transform the pointwise product
of the Fourier transforms em and ~n' Since each of 1* g and J. g is evidently
bilinear in the pair (j, g), the desired relation will obtain for functions I and g
which are trigonometric polynomials, that is, finite linear combinations of
the en' It thus appears that (3.1.1) constitutes a hopeful starting point. We
proceed to the details forthwith.
Suppose that I and g belong to Ll. Then the Fubini Tonelli theorem is
applicable (see [W], Theorems 4.2b, 4.2c, .and 4.2d; [HS], pp. 384-386, 396;
[AB], pp. 154-155) and shows that the integrand appearing on the right of
(3.1.1) is, for almost all x, an integrable function of y, so that (3.1.1) effectively
defines 1* g(x) for almost all x; moreover the function so defined almost
everywhere is measurable and
(3.1.2)
In particular, I * g ELI.
50
[3.1] DEFINITION AND FIRST PROPERTIES OF CONVOLUTION 51
(J * g)"'(n) I*
= 2~ f g(x)e-1n;r dx
= (2~r II
= (2~rI{I
e-1n;r f(x - y)g(y) d(x, y)
I
= 21T I .
g(y)e- 1nll {f(n)} dy
Since (Ji * gil converges in Ll to 1* g, it follows that (3.1.1) holds for almost
all x. This shows in particular that y -+ f(x - y)g(y) is integrable for almost
all x whenever I and g are nonnegative functions in Ll. IfI and g are replaced
by If I and Igl, it is seen that the same is true whenever f and g belong to Ll.
Once (3.1.1) is established for general I, g E Ll, (3.1.3) results immediately.
[The reader should note that it is not the case that every nonnegative
integrable function h is equal a.e. to the limit of an t sequence (h n ) of
continuous nonnegative functions. As a counterexample, take h to be defined
on [0, 2rr] as the characteristic function of the complement, relative to [0, 2rr],
of a closed, nowhere dense set K ; [0, 2rr] having positive measure.]
Similar techniques are applicable in connection with 3.1.4 to 3.1.6.
At this point it is convenient to summarize what little we do know so far
about convolution and to mention a few questions that arise, which will
guide some of the subsequent developments.
g(ak)1 ~ e
k=l
y) - g(a k - y)1 ~ e
k=l
L II*g(bk) -I*g(ak)1
r
~ IIIlIloe
k=l
III*gllp ~ IIIlIlolIglip
Proofo For any h E U the Fubini-Tonelli theorem gives
The converse of Holder's inequality (see Exercise 3.6) now goes to show that
1 * g E LP and 111 * gllp ~ Ilflll Ilgllp, as alleged.
Remarkso (1) The argument can be made less sophisticated by assuming
first that 1 and g are continuous. One may then assume that h, too, is con-
tinuous. The required versions of the Fubini-Tonelli theorem and the Holder
inequality and its converse then become simpler. This leads to the stated
result for 1 and g continuous. In general, we may assume that p < co, since
otherwise the result is contained in 3.1.4, and then approximate 1 and g in Ll
and in LP, respectively, by continuous functions I .. and g" (n = 1,2, .. ).
By (3.1.2), I .. * g" -+ 1 * g in LI, and so a subsequence converges almost
everywhere. By applying the result already established for continuous 1 and
g to the terms of such a subsequence, and making use of Fatou's lemma ([W],
Theorem 4.Id) on the way, the desired result appears.
[3.1] DEFINITION AND FIRST PROPERTIES OF CONVOLUTION 57
(2) The assertion in 3.1.6 can be improved in several ways; see 12.7.3 and
Exercise 13.5. A result that combines and extends 3.1.4 and 3.1.6 will be
obtained in Section 13.6.
3.1.8. Convolution and Translation. Both 3.1.2 and 3.1.3 hint at close
connections between translation operators T a and convolution. This will be
borne out as we progress (see especially Sections 16.2 and 16.3). Meanwhile
here is a basic result in this direction.
3.1.9. Let JELl and let E denote anyone of the normed spaces C or LP
(1 ~ P < (0). If gEE, thenJ * g is the limit in E of finite linear combinations
of translates of g.
Proof. Let gEE be given. Denote by Vg the closed linear subspace of E
generated by the translates TaU of g, that is, the closure in E of the set of all
finite linear combinations of elements TaU. Denote also by 8 the set of JELl
such that J * g E VII' It has to be shown that 8 = Ll. Now it is evident that
8 is a linear subspace of Ll; and from 3.1.6 it follows that 8 is closed in Ll. It
will therefore suffice to show that 8 contains a subset 80 such that the finite
linear combinations of elements of 8 0 are dense in Ll.
If E = C, a convenient choice of 8 0 is C (see [W], Theorem 4.3b). We will
leave to the reader the task of showing that in fact J * g is the uniform
limit of finite linear combinations of translates of g, whenever J .and g are
continuous. (Hint: Approximate the integral definingJ * g by Riemann sums,
using uniform continuity of the functions involved.)
We pass on to the remaining cases.
Suppose, then, that E = L" (1 ~ P < (0). In this case a convenient choice
of 80 is the set of functions J which coincide on [0, 21T] with the characteristic
function of an interval I = [a, b], where 0 < a < b < 21T, and which are
defined elsewhere by periodicity (compare [W], Theorem 4.3a). In this case
58 CONVOLUTIONS OF FUNCTIONS
II!* y -
1
-2
'TT1e
L: Illel' T"kyllp ~ 2-1 L: IIhlell
'TT
p (3.1.9)
Since
[3.2] APPROXIMATE IDENTITIES FOR CONVOLUTION 59
(3.2.1)
and
lim
n_CXl
r
)6<1%1<"
\Kn(x)\ dx = 0 (3.2.3)
lim r
"... "" J6 "Ixl""
X,,(x) dx =0
for each fixed 8 satisfying 0 < 8 < 1T, constitutes an approximate identity.
Thus, we might take for Xn (n = 1,2, .. ) the function which is defined on
[ -1T, 1T) to be ?Tn times the characteristic function ofthe interval [ -lin, lin],
and defined elsewhere by periodicity.
The name "approximate identity" is justified by the following result.
gives
(3.2.4)
say. Being assigned any e > 0, choose and fix 8 satisfying 0 < 8 ~ ?T, so
that IIT],.j-fll", ~ dar Iyl ~ 8.
Then
(3.2.5)
[3.2] APPROXIMATE IDENTITIES FOR CONVOLUTION 61
where M is independent of n. Since also IITlIf - fll .. ~ 211fll .. for all y, the
second integral is majorized by
as a *-module over Ll, this latter set is just the closed submodule of E
generated by I. Further results of this type will appear in Section 11.1.
lim 21 fKII(x)/(x) dx
A_CD 'Ir
= 1(0)
for each continuous I. (Actually, scrutiny of the proof would show that this
holds for any IE L'" such that I is continuous at 0.) This means that the
sequence (Kn) is of the type which is often said to converge (in some un-
specified sense) to the so-called Dirac S-Iunction. Complete precision will be
attained in terms of the ideas to be studied in Chapter 12; see especially
12.2.3 and 12.3.2(3).
is an invariant integral on G (see 2.2.2 and Exercise 2.5), a little thought will
show that the group algebra may very well be pictured as the algebra of
K-valued functions on G, the linear space operations being point-wise and
the product being convolution; the sum
L: /(x -
Ilea
y)g(y)
2~ J f(x - y)g(y) dy
used in Section 3.1 to define the convolution of two functions on the group
T.
It is no part of our purpose to carry forward the study of the group algebra
of a finite group (see, for example, [Bo] and [vW]): the concept is mentioned
merely because it is the forerunner of one that holds an important place in
the modern .developments in harmonic analysis (see 3.3.2). For purposes of
subsequent comparison it is to be noted that the study of the group algebra
of a finite group leads ultimately to a good deal of information about the
structure of the underlying group, albeit only when combined with the study
of representations of the group.
EXERCISES
3.3. Take p = 1 in the preceding exercise. Show that I and ,J are non-
closed in Ll.
Hint: Assume the result of the computations in 5.1.1 and use the uniform
boundedness principle as stated in Appendix B.2.1 or in B.2.2.
3.4. (1) Let E be the set off ELl such thatf(x) = f( -x) a.e. Is E an ideal
in L1? Is it a subalgebra of L1? Is E closed in L1? Give reasons for your
answers.
(2) Prove that W1 (f * g) ~ II! III . w1g for f, g E Ll. Deduce that the set
1= {fE Ll: wd(a) = O(laI112) as a~ O}
is an ideal in Ll. Is I closed in Ll? Give reasons.
Hint for (2): Observe that I is everywhere dense in L1.
3.5. Suppose that f E La> is such that the function a ~ Taf is continuous
from R into L'" for the normed topology on the latter space (that is, that
IITaf-fII",~O as a~O). Prove thatfis equal almost everywhere to a.
continuous function. (The converse is true and almost trivial.)
Hints: Take an approximate identity (K n );:'=l in Ll and consider the
functions fn = Kn *f. Show that the fn are equicontinuous and uniformly
bounded. Let (xdt": 1 be a sequence that is everywhere dense in (0, 21T) and
pick strictly increasing sequences of natural numbers (nk(i))k'=l so that
(nk (1+ 1)k'= 1 is a subsequence of (nk(f))k'= 1 and limk~ fn<(O(x,) exists finitely
<Xl
for each i. Deduce that there exists g E C such that, if n k = nk(k) (the
"diagonal subsequence"), then fnk ~ g uniformly. Use 3.2.2 to compare /
and g.
Remarks. This is the special case, for the group R/21fZ, of a result due to
D. A. Edwards [1] for general groups. An analogous and older result for Radon
measures (see Chapter 12) is the work of Plessner and Raikov. Both types of
result are treated in R. E. Edwards [2]. See also Exercises 11.22 and 12.23. The
existence of a uniformly convergent subsequence of (in) is a special case of
Ascoli's theorem; see [E], Section 0.4.
3.6. (Converse of Holder's inequality) Suppose that 1 ~ P ~ 00 and that
/ E Ll is such that
(1)
3.8. Let (K,,);'ml be any approximate identity. Show that, if p > 1, then
lim KN
N-",
*1 = 1
uniformly for each continuous J. Deduce that (KN)N= 1 is an approximate
identity (see 3.2.1).
3.11. Assume that a is a real number such that a/TT is irrational and that
f is a measurable complex-valued function such that Tal = f a.e. Show
that I = const a.e. (Recall that all functions considered have period 211'.)
3.12. Suppose 1 ~ P ~ 00. Show that the convolution algebra L" has no
nonzero generalized (or topological) nilpotents, that is, elements 1 such that
inf" 111*"11/'" = 0,
where 1*1 = 1 and 1*(1<+ 1) = 1 *1*" (k = 1,2, .. ). See 11.4.18(1).
3.13. Assume that K E Ll and IIKlll < 1. Given gEL" (where 1 ~ P ~ (0),
show that the equation
f = g + L'" K*" * g,
,,=1
EXERCISES 67
neZ
2: Ic,.j(n)i < 00
LetJbe the periodic function that coincides on [-17', 17'} with the characteristic
function of S. Apply 3.1.4 to J * j. .
Notes: The result is due to Steinhaus (Sur les distances des points des
ensembles de mesure positive. Fund. Math. 1 (1920), 93-104). It could be for-
mulated entirely in terms of the group T; in fact, the result is valid for any
locally compact group whatsoever, the proof being a simple adaptation of tha.t
which is proposed above. A corollary is that a subgroup is either open or has
zero interior measure.
For some generalizations, see Ray and Lahiri [I], Mueller [I], and the
references cited there.
There is an analogous result due to Banach, Kuratowski, and Pettis applying
to topological groups that are not necessarily locally compact, measure-
theoretic concepts being repla.ced by category; see [K], p. 211 and (for special
cases) [Zd, p. 250, Example 2. For a converse, see MR 51 # 6286.
3.17. Let (an}:'=l be a sequence of real numbers and suppose that
lim n _ '" exp (ianx) exists for each x belonging to a set of real numbers having
positive interior measure (see the preceding exercise). Prove that lim n _ co an
exists finitely.
Hints: The set S of points of convergence of the sequence (exp (ianx)) is
evidently a subgroup of R. Use Exercise 3.16 to conclude that S = R, so that
g(x) = lim n _ '" exp (ianx) exists for all real x. By integration theory, therefore,
lim
n-aJ
r J(x} exp (ianx) dx = JRr J(x}g(x} dx
JB
for every function J which is Lebesgue-integrable over R. Deduce first that
(a n):'= 1 is bounded (an adaptation of 2.3.8 will be needed here), and then (by
choosing J suitably, or by a compactness argument) that this sequence is
convergent.
3.18. Let (Cn}:'=l be a sequence of complex numbers and (an):'=l a
sequence of real numbers. Suppose that lim n _ co Cn exp (ianx) exists for each
x belonging to a set of real numbers having positive interior measure (see
Exercise 3.16). Show that (i) (c n ) is convergent to some complex number and
(ii) if lim n _ '" Cn i= 0, then (ex n ) is convergent in R.
3.19. Let X be a measurable character of T. Prove that X is continuous.
Hints: Use Exercise 3.16 to show that X is bounded. Then establish
continuity of X by adapting the reasoning' used in 2.2.1 to show that a
continuous character is differentiable.
Alternatively, see [HR), p. 346, where the result is stated and proved in a
more general form.
Remark. Despite the stated result, there exist characters of T which are
both hounded and nonmeasurahle.
CHAPTER 4
Homomorphisms of Convolution
Algebras
In this brief chapter we introduce the reader to two problems typical of the
current outlook on harmonic analysis. The first problem, which will be solved
in detail in Section 4.1, arises on choosing anyone of the convolution algebras
E mentioned in Subsection 3.3.2 and seeking to determine all the homo-
morphisms I' of E into the complex field. The answer highlights the funda-
mental im portance of the Fourier transformation in relation to group structure.
The second problem is concerned with the (self-) homomorphisms of E
(that is, the homomorphisms of E into itself). Of the available choices of E,
only the cases L2 and L1 are fully solved. The former case is easy and of
relatively little interest (compare Exercise 8.1). The case E = L1 is, on the
contrary, comparatively very complex, and we shall be able only to indicate
how the solution of the complex homomorphism problem allows a useful
reduction to be made, and to indicate the solution for this case.
An incomplete account of the homomorphism problem is inserted at this
early stage because it has been learned in 3.1.9 and 3.1.10 that convolution
is related in a very basic way to the group structure, and because, granted
the ensuing fundamental role of convolution, the homomorphism problem
begs for recognition without delay. This problem has, in fact, proved to be
one focus of interest in contemporary work.
Yn(J) = j(n)
for fE Ll.
Proof. The uniqueness of It is clear.
We shall offer two proofs of the existence of n, the relative merits of which
will be weighed in 4.1.3.
First Proof. Define Cn = y(e n ) for n E Z. Since y is continuous and
nontrivial, the density theorem of 2.4.4 entails that Cn i: 0 for at least one
integer n. Since also en * en' = en or 0 according as n' is, or is not, equal to n,
an application of Y shows that Cn Cn' = Cn or 0 according as n' is, or is not,
equal to n. Thus Cn , is equal to 1 or to 0 according as n' is, or is not, equal
to n. Linearity of Y shows then that y(J) = j(n) for all trigonometric
polynomialsf. Continuity of y, together with the density theorem, accordingly
show that y(f) = j(n) for allfE Ll.
Second Proof. In view of the continuity of y and the fact that C is
everywhere dense in V, it will suffice to show that, for some nEZ, the
formula
y(J) = j(n)
holds for each continuous f.
Now, again since y is continuous, we can choose and fix a continuous fo
such that y(Jo) is nonzero. Consider the function X defined on Rj21TZ by
x(x + y) = X(x)X( y) .
(4.1.4)
Now take any continuousf. Then (compare the proof of 3.1.9), given any
B > 0, fo * f is uniformly approximated to within
B by any sum of the form
[4.1] COMPLEX HOl'tlOMORPHIS}IS AND FOURIER COEFFICIENTS 71
where 0 = Xo < Xl < ... < xm = 271 is any partition of[O, 271] whose "mesh"
max"(x,, - X"-l) is sufficiently small (depending upon '/0' and/). A fortiori,
for any such partition. Therefore, if we apply y and use (4.1.2), we conclude
that
Using (4.1.1), dividing through by y(/o). and using (4.1.3) and (4.1.4), we
find that
(4.1.5)
for all partitions of sufficiently small mesh. But, as the mesh of the partition
tends to zero, the sum appearing on the left of (4.1.5) converges (since / is
continuous) to the integral
f -
1 /(x)e-1n:r dx = /(n).
271
In the case of LOO, one change is necessary owing to the fact that C is not
dense in LOO. However, (4.1.1) and the fact thatf*gEC whenever fEL""
and gEL"', ensures that Y, being by hypothesis not identically zero on L"",
cannot vanish identically on C. Hence fo can be chosen as before. The
preceding argument shows that y{f) = j(n) = Yn{f) for all f E C. But then,
if gEL"', we have (since fo * g E C)
y{fo * g) = Yn{fo * g),
that is,
y{fo)y(g) = Yn{fo)Yn(g)
Since y{fo) = Yn{fo) # 0, so y(g) = Yn(g) for all gEL"'. The conclusion is thus
valid for LOO also.
The second proof is clearly more complicated and laborious than the first.
It has been included, because it can be adapted to cases in which the under-
lying group is noncompact (the groups Z and R, for example). In such cases
the first proof breaks down completely because V then contains no con-
tinuous characters at all. Moreover, the second proof bypasses the density
theorem and relies on fewer facts concerning harmonic analysis.
4.1.5. A full exploitation of the results obtained in 4.1.2 and 4.1.3 depends
on the Gelfand theory of complex commutative Banach algebras. Were we
seeking to develop harmonic analysis on a general group, it would at this
point be advantageous to embark on the Gelfand theory and reap the fruits
of its application. As it is, however, we shall defer this sowing and harvesting
until Section 11.4.
for f, gEE. (By using the closed graph theorem as indicated in 4.2.3, together
with the necessary continuity of complex homomorphisms of E, one could
show that any homomorphism of E into itself is necessarily continuous.)
In Exercises 4.2, 4.3, and 4.8, the reader's attention is directed to the
simplest and most obvious such homomorphisms, namely those of the type
for n E Y and fEE. This' relationship entails that n' is uniquely determined
by n E Y, so that a mapping
a: Y -* Z
"L/(a(ne" =
"eZ fteY
"L I(a(n))e" E E (4.2.4)
whenever fEE. The reader will notice that (4.2.4) is intended to mean simply
that there exists gEE (unique by 2.4.1) such that g = loa; no statement is
intended concerning the convergence of the series in (4.2.4). (The series do in
fact converge distributionally, as will appear in Chapter 12; no use is made of
this fact.)
lim
Ie ......
lie = 0 in E, lim Tile = g in E
Ie ......
(4.2.6)
But, in all cases here envisaged, limle .... 00 lie = 0 in E entails that limle .... 00 A = 0
pointwise on Z. Similarly the second clause of (4.2.6) entails thatlim/c ..... (TI/c)"
= gpointwise on Z. Since (Tltr' = l/c 0 a, a(Z) c Z v {<Xl}, andA(<Xl) = 0, it
follows that g = 0 and so (by the uniqueness theorem once again) the
conclusion (4.2.7) is obtained. This completes the proof that T is continuous.
EXERCISES
4.4. Suppose that T and IX are as in 4.2.1 and 4.2.2. Show that T is
one-to-one and T(E) contains all trigonometric polynomials if and only if
y = Z and IX maps Z one-toone onto itself.
EXERCISES 77
L:
"Ea- 1({m
e.. ll" ~ k.
In this chapter we introduce the so-called Dirichlet and Fejer kernels and
their elementary properties. These kernels are basic in the study of pointwise
convergence and summability, respectively, of Fourier series. From their
properties we shall derive the localization principle, together with alternative
proofs of the uniqueness and approximation theorems of 2.4.1 and 2.4.4.
Included in the text and in some of the attached exercises are a few
properties of Cesaro summability as applied to general series, most of which
will later be applied in the case of Fourier series.
8N!(X) = L
Inl';;N
j(n)e 1nz
When we speak of the convergence of the Fourier series of! we shall always
mean the convergence of these symmetric partial sums; see the end of 2.2.2.
Inserting the integral expression for j(n) we find that
DN(x) = 2:
Inl';N
e1nx = sin (l!
sm
-:- Yz)x;
Y2x
(5.1.2)
(5.1.3)
Furthermore,
(0 < 8 ~ Ixl ~ 1T). (5.1.4)
The Nth Cesaro sum (see the end of 2.2.2) of the Fourier series of f is
the arithmetic mean of the first N + 1 terms of the sequence of symmetric
partial sums thereof, namely,
(5.1.5)
These are also spoken of as the (C, I)-means of the Fourier series off' .. "c"
for Cesaro and the" 1" indicating first order arithmetic means. Using (5.1.1)
and some elementary calculations, we find that
= 2: (1 - N_In_I
Inl.;N +1
)e 1nX
(5.1.8)
80 THE DIRICHLET AND FEJER KERNELS. CEsARO SUMMABILITY
Moreover,
(5.1.10)
IIDNl11 =~ I"
2rr -11
ISin (~ + %)Xl dx
sm %x
2 i,,/2lsin
= -rr 0 . + l)YI d y
(2N
smy
(tt
pu mg y = 11 )
72 X
I DNl11 = -2 L
2N
1/( %krr) 5" /2
uk(s) ds + 0(1)
rr k= 1 0
=-
2
rr
L -2
2N
krr
1t=1
+ 0(1).
Since L.~~1 11k = log N + 0(1), (5.1.10) is established.
[5.2] THE LOCALIZATION PRINCIPLE 81
5.1.2. New Proofs of 2.4.1 and 2.4.4. It is worth noting that 3.2.2 and
the opening sentence of 5.1.1 combine to yield independent proofs of 2.4.1
and 2.4.4.
Consider, for example, 2.4.1. If j = 0, then (5.1.6) shows that FN *f = 0
for all N. Since, as has been remarked in 5.1.1, the FN form an approximate
identity in V, 3.2.2 shows thatf, the mean limit in V (and, iff is continuous,
the uniform limit) of F N * f as N -+ 00, is zero almost everywhere (or
everywhere).
The deduction of 2.4.4 follows similarly from (5.1.6) and 3.2.2, the former
showing that FN * f is a trigonometric polynomial for all N and all fELl.
The same arguments and sources yield results that may be interpreted in
terms of the Cesaro summability of Fourier series, but we defer this develop-
ment until Chapter 6.
5.2.1. If f and g are integrable functions and if, for a given point x, the
function
y-+f(Y) - g(y)
y-x
is integrable over some neighborhood of the point x, then
(5.2.1)
and
(5.2.2)
Proof. Since both h -+ SNh and h -+ uNh are linear operators, we may
assume without loss of generality that g = 0 throughout. Again, since
(5.1.1), (5.1.6), and 3.1.2 show that
we may take x = O.
Now, by (5.1.1),
SNf(O) = 217
1 I"
-ll f( y)D N( y) dy
= ~I"
217 -ll
{!(y) }'Sin(N
sm Y2y
+ Y2)ydy. (5.2.3)
82 THE DIRICHLET AND FEJER KERNELS jESARO SU~L\[ABILITY
Consider a two-way infinite series Lnez Cn and define the partial sums
We recall (see the end of 2.2.2) that the given series is said to be Cesaro (or
(C, 1i-) summable to sum s if and only if
[There is a Cesaro method (C, a) for every a i= -1, - 2, .. " but we shall
make n,) use of this concept; see, for example [Zl]' p. 76.]
Consider as an example the series Lneze1nx. Equation (5.1.2) shows that
this series converges (to a finite sum) for no real values of x. On the other
hand, (5.1.7) shows that this same series is Cesaro-sum mabie to 0 for all real
x tc. 0 (mod 211"). It will appear in Chapters 6 and 10 that this very special
example is surprisingly significant in relation to the behavior of Fourier
series in general.
Turning to generalities, we shall first verify that the Cesaro method of
summability is stronger than, and consistent with, ordinary convergence.
5.3.2. W~ have in 5.3.1 tacitly assumed that the limit s is finite. However,
if we look at real-valued sequences SN' the result extends to properly divergent
ones. For example, the reader will verify easily that the preceding proof
adapts easily to show that, if lim SN = 00, then also lim aN = 00. More
generally, indeed, one has in all cases
2: e
neZ
ln %
5.3.4. If the SN are real and increasing (for example, if en ~ 0), then Cesaro
summability implies convergence.
Proof. There is no loss of generality in assuming that SN ~ O. Then
EXERCISES
for some a > O. Let N be any integer such that N a > 1. Prove that f*N is
equal almost everywhere to a continuous function, where f*N = f * ... * f
(N factors). (Compare with Exercise 8.4.)
Note: In the following exercises the notation is as in Section 5.3.
5.3. Show that if limN ... ",aN exists finitely, then 8N = o(N) as N _00.
5.4. . Give detailed proofs of the statements in 5.3.2.
5.5. Show that tlN == 8N - aN is expressible as
tlN = (N + 1)-1 L
I"I",N
Inlc".
Show that if
M == L InIP-
neZ
1 Ic"IP < 00
where A is an absolute constant. Notice that lim supN ... ..,I~NI is unaffected if,
for any k, en is redefined to be 0 for Inl :s;; k.
5.6. (I) Suppose that en _ 0 as Inl-oo, and that e" = 0, except perha.ps
for n = 0, n" (k = 1,2, .), where I :s;; n 1 < n 2 < ... and inf" nk+ljn"
== q > 1. Show that if lim.v",,,,uN = 8, then lim N ... ..,8 N = 8.
(2) Show also that the same conclusion is valid whenever en = o(Ijn).
Hint: Define tlN as in Exercise 5.5, and write
L u,qr-".
Ie
I~NI :s;; 2
ral
1 (nm1) '.
2:-=0-
n/c
/c>m
and that
UN./C = SN + ""
L... (1 - Inl k- N) cn
N<)n) .. N+/C
Deduce from the former relation that if k = kN ~oo with N in such a way
that NlkN remains bounded, and if limN_",uN = s, then limN_",uN./C = s.
5.8. Suppose that Cn = O(l/lnl) as Inl ~oo and that limN_",uN = s.
Show that limN_",sN = s. (This is Hardy's theorem referred to in 5.3.5.)
Hint: Use the second relation established in the preceding exercise to
show that
2: J(n)e1nZo = 1(zo)'
nEZ
Hint: Use 5.2.3.
CHAPTER 6
It follows from Exercises 3.1 and 3.2 that the second part of 6.1.1 is not
true, if uNf is replaced by 8 Nf; the first part also fails. In both cases the
failure is a consequence of the fact that
itself a corollary of (5.1.10). (For the case k = 0 of 6.U, see also 10.3 below.)
For a homogeneous Banach space B over T satisfying IlenfllB = IIfllB for
all n E Z and all fEB, it is known (see [Kz], p. 49) that
lim
N-J
IIf - 8 NfJIB = 0 for all fEB
if and only ifB admits conjugation; that is to say, if and only iff E B implies
JE B, J
where the conjugate function is defined as in 12.8.1. [This last is
equivalent to demanding that, for every fEB, there is a function JE B such
that
However,
UN(n) = (1 - )~ l)c n
6.1.4. Behavior of IIf - uNf lip as N ~co. Although we know from 6.1.1
that, as N ~CO,
[6.1] UNIFORM AND MEAN SUMMABILITY 89
for all n E Z and N = 1,2, . '. Therefore, in view of (1), it will appear from
(6.1.2) that
I)(N(n) - 11 ~ %
for all n E Z and infinitely many N. For any such N, it thus appears that
lim inf IKN(n)i ~ %> 0,
Inl-'"
uNf(x) = L en e1n%,
Inl';'"
I(J - I
P)(x) ~ e
6.2.4. Bernstein Polynomials. In just the same way that 6.1.1 includes
and refines 2.4.4, a famous theorem of Bernstein includes and refines the
Weierstrass theorem in 6:2.2. Bernstein's theorem asserts that, if f is a con-
tinuous function on [0, 1], then the associated so-called Bermtein polynomial8
..,
B..,f(x) = 2: f(nfN) ..,Cn xn(l - x)N-n (N = 0, 1, 2, ... )
n=O
We now turn to some deductions from 6.l.1 more closely connected with
Fourier series.
92 CEsARO SUMMABILITY OF FOURIER SERIES
2 II
7T
f(x)g(x)dx= Lf(n)"g(-n),
neZ
"'. A
lim
N-oo
2:
Inl",N
j( -n) g(n),
passage to the limit under the integral sign being justified by Lebesgue's
theorem ([W], Theorem 4.lb). From this the stated results follow.
2~ I U(X)SN,p(X) dx = Lf u(x)g(x) dx
2~ f u(x)h(x) dx = 2~ f u(x)g(x) dx
6.2.8. If J E L\ then
f b
a
J(x) dx = 2: j(n)
neZ
elnb
:-
~n
e 1na
,
2:
Inl .. N
j(n)' 21Tg( -n).
(6.3.2)
It is our aim to give simple conditions sufficient to ensure that the expression
(6.3.2) tends to zero as N -,)000. In studying this we shall not assume from the
outset that 8 is the" right" value, namely f (x). Indeed, it is the behavior of
n in the neighborhood of y = 0 (that is, of f in the neighborhood of x) which
is significant, and the prime feature is not the value ofn at o (that is, of f
at x) but rather the limiting behavior of f~ near 0 (that is, of f near x). In
this section we consider the simplest case, in which we assume outright the
existence of the limit off~(y) as y-,)o +0.
The limit in (6.3.4) is attained uniformly on any set on which the limit in
(6.3.3) is attained uniformly.
Proof. The formula (6.3.2) will be applied, taking therein
8 = Y2[f(x + 0) + f(x - 0)).
say. Then
[6.4] POINTWISE SUM)IABILITY ALMOST EVERYWHERE 95
by (5.1.8). (Notice that we are here using positivity of F 1'1: this is not available
for DN and the substitution of IDNI for DN would vitiate the argument
irreparably.) On the other hand, the use of (5.1.9) leads to
If t: and S are held fixed while N is allowed to tend to infinity, it is seen that
Since t: is a freely chosen positive number, the desired conclusion now follows.
(6.4.1)
which is easily established by examining separately the ranges 0 :s;; Y :s;; TT/ N
and TT/N :s;; Y :s;; TT; in the first interval, F N( y) is majorized by a multiple of
N, and in the second by a multiple of N-ly-2. This inequality shows in
turn that
(6.4.3)
96 CEsARO SUMMABILITY OF FOURIER SERIES
6.4.2. Regarding the theorem of Lebesgue it suffices to recall that for any
integrable f, periodic or not, and almost all x it is true that
when s is taken to be f (x). Such points x are usually termed Lebesgue points
of f and the set of such points the Lebesgue set of f.
Now we can state and prove the main theorems of this section.
Assuming that (6.4.4) is satisfied, we suppose that e > 0 is given and choose
8 > 0 so that
By (6.4.1),
Now, for any S > 0, the inequality (5.1.9) shows that the integral appearing
in (6.4.7) tends to zero as N ---+00. So, keeping e and S fixed, we infer from
(6.4.7) that
lim sup laNf(x) - 81 ~ 7T- l (%A + B)e.
N->
.
Ilu*J lip ~ ApllJ lip if J E LP and p > 1, (6.4.11)
of the ordinary partial sums of the Fourier series of J, see 10.3.5 and 10.4.5.
6.4.S. The Estimate (6.4.9). The relation (6.4.9), valid for g E V", is close
to being the best possible. Indeed, if we are given any set E of measure zero,
there exists a nonnegative function f belonging to LP for every p < 00 and
such that
(x E E). (6.4.15)
To verify this we may assume that E lies in (0,21T) and choose numbers
a" > 0 and c" ~ 0 (k = I, 2, ... ) so that
(6.4.16)
for every p < 00; for example, c" = log k, a" = k - 2. Since E has measure
zero, we may choose open sets E" such that E c E" c (0, 21T) and having
measure m(E,,) E; ale' Let fIe denote the characteristic function of E,,, extended
by periodicity, and put
Then f ELP for every p < 00, as follows from the second clause of (6.4.16) and
Beppo Levi's theorem ([W], Theorem 4.1e). Observe also thatf is nonnegative
and lower semicontinuous. According to 6.3.1,
lim uNf,,(x) = 1
N~"
for all x E E" and, a fortiori, for x E E. Since FN is nonnegative, one has for all
x and all k
[6.5] APPROXIMATION BY TRIGONOMETRIC POLYNOMIALS 99
see Exercises 6.12 and 6.13, and also 10.1.1, 10.1.4, 10.2.1, and 10.2.3. As will
appear in 10.3.4, the relation (6.4.17) is hopelessly over-optimistic for general
feLl.
6.5.1. The Functionals PN and EN' For definiteness we shall work within
the Banach space C of continuous functions, but the reader will scarcely
need to be told that each question posed in this setting has some sort of
analogue for the case in which C is replaced by V' (or, indeed, by anyone
of a number of other quite natural function spaces).
In order to examine the questions raised in vague terms in 6.2.1(2), we
introduce two sequences of functionals PN and EN (N = 0,1,2, .. ) defined
on C in the following way:
(6.5.1)
where TN denotes the set of trigonometric polynomials of degree at most N.
Plainly, the relative magnitude of PHf and EHf provides a sensible measure
of just how good UHf is as an approximant to f, when compared with other
elements of T H'
It is evident that
EN+d:!O; EHf " PH!.
and that PN! = 0 if and only if! is a constant function. Moreover (see
Exercise 6.5), the infimum EH! is actually an a.ssumed minimum; as a
consequence it follows that E Hf = 0 if and only if! e T H'
According to 6.1.1, PH! = 0(1) as N -00 for each f e C; yet, by 6.1.4, the
rela.tion PNf = O(eN) is false for some f e C (in fact. for a nonmeager set of
100 CEsARO SUMMABILITY OF FOURIER SERIES
IE C) whenever E,y = 0(1) is given. It is also easily shown (see Exercise 6.10)
that the relation p,yl = o(ljN) holds if and only if I is a constant. This
simple result is noteworthy when it is compared with Exercises 6.8 and 6.9;
it then shows conclusively that, for sufficiently smooth nonconstant con-
tinuous functions I, aNf is far from being the optimal approximant to f
among all elements of TN. If f is very smooth, sNf is a decidedly better
approximant to I than is aNf. Crudely speaking, the great advantage of aNf
is to be seen for general.continuous functions f.
After these preliminary remarks we now proceed to establish some
improved estimates for PNI in case I satisfies certain Lipschitz (or Holder)
conditions. This latter type of condition is for our immediate purposes best
expressed in terms of a modified modulus of continuity, namely,
compare with (8.5.1) and the definition of wd in 2.3.7. Applying (6.3.1) and
choosing s = I(x), we see that
(6.5.2)
PN f ~
AN IllN
7T 0
r\
Ua:
I( Y)dY + ~
N
7T
I"
lIN
O",/(y) dy
Y
2 (6.5.3)
on some interval (0, e), where e > 0, while y-aw(y) is decreasing on the same
interval for some choice of a satisfying < a < 1. Suppose further that f is
continuous and satisfies
D.",f(y) = O[w(y)] (respectively,o[w(y)]) as y~ +0. (6.5.4)
Then
as N ~ ct) (6.5.5)
Proof. If e < 1T, and if we extend w by setting w( y) = w(e) for e < y ~ 1T,
this extended function satisfies the required conditions on the interval
(0, 1T]. Thus, we may as well assume from the outset that c = 1T.
Now, without appeal to 6.5.2. it follows from (6.5.4) that
{liN (liN
N Jo D.",f( y) dy = 0 (respectively, 0) (N . Jo w( y) dy)
= 0(respectively, 0) (N'N-l.w(~))
= 0 (respectively, 0) (w(~)), (6.5.6)
liN
It now remains but to combine (6.5.3), (6.5.6), and (6.5.7) in order to derive
(6.5.5).
6.5.4. Remarks (1). The majorization given in 6.5.3 is. at least for certain
natural choices of w, the best poBBible (see [Bad. p. 206). Somewhat similar
102 CEsARO SUMMABILITY OF FOURIER SERIES
results are known for functions! which satisfy a mean Lipschitz condition of
the type IITJ - flit = O(iala) as lal->- 0; see [Zd, p. 117. See also MR 53
# 6203.
(2) Inasmuch as EN! "" PN!, Subsection 6.5.3 and its analogues yield majorants
for EN! for restricted functions!. Superior results are obtainable by estimating,
not
PN! = Ilf - UN! II "',
but rather II! - 'TN!II"" where
'TN! = 2U2N-d - uN-deT2N - l ;
some calculations of this nature are proposed in Exercises 6.6 to 6.9. More
elaborate results appear in Timan [1]. One might also use to the same end
the socalled Jackson polynomials J N * f, where
J N(X) = sin
eN ( .
%N'X)'
1; ,
sm ZX
where N' = [%N] + 1 and where the number eN is chosen to make IIJNlll = 1;
see [L 2 ], pp. 55-56. See also W. R. Bloom [3], [4].
The Abel means of the Fourier series of! are the functions
Arf(x) = Lr
neZ
lnl j(n)e tnx
= P T *!(x), (6.6.1)
EXERCISES
then
-.!..ffg
21T
dX = lim L
N_oo l"I<iN
(1 - N 1n1+ l)i{n)y(-n).
6.3. Write, for f E L1,
Prove that
6.4. Is
( _1)"e 1nx
~ (1 + Ini)log (2 + n 2 )
EXERCISES 105
TNI(x) =:;;:2f<Xl
0
[ f ( x + Ny) + f ( x Y)] k(y)
- N ~. dy
where
h(y) = sin 2 y - sin 2 Yzy = Y2(cOS y - cos 2y).
6.7. The notation being as in Exercise 6.6, define
Verify that
(I) fa<Xl IHt(y)i dy < 00 (i = 0, 1,2,),
-4f<Xl H o( y) dy = 1,
1T 0
~ fa<Xl cos (1)' Ho(Y) dy = 1.
Integrate the second relation by parts and use the first to obtain
its Nth Cesaro mean. Show that the given series is a Fourier-Lebesgue series
if and only if
lim IluN - UN' III = O.
N.N'-.:o
N-l 2:
Inl<iN
Inj(n)I = 0(1) as N --+00,
then limN_.,sNf(x) = f(x) for almost all x; and that if, furthermore, f is
continuous, then limN_",sNf = f uniformly.
H?:nt: Use Exercise 5.5 and the results appea~ing in 6.1.1 and 6.4.4.
Note: One can show, by using the same method in combination with
8.3.1, that if Cn = o( I/Ini) as Inl --+ co, then the trigonometric series Lnezcnelnr
is convergent for almost all x.
6.13. Suppose that fE Ll and that j(n) = 0 save perhaps for n = 0,
nt, n 2 ,' " where 0 < nl < n 2 < ... and inf nk+1/nk > I. Prove that
limN_",sNf(x) = f(x) for almost all x; and that if, in addition,J is continuous,
then limN_",sNf = f uniformly.
Hint: Use Exercise 5.6, and Subsections 6.1.1 and 6.4.4.
Note: These so-called lacunary series will receive further attention in
Chapter 15; see especially the remarks in Section 15.6. The result, like that in
Exercise 5.6, admits some generalization.
6.14. Let the function rN (N = 1, 2, ... ) be defined to be equal to l7N in
[ -liN, liN), to 0 in [-17, -liN) and in (liN, 17), and be defined elsewhere so
as to be periodic. Put RN = rN * rN. Verify that (RN);=l is an approximate
identity. Compute flN and deduce that
where Qr( y) = - sin y P~( y). Verify that Qr ~ 0, IIQr !Il = r, and that
lim
r-1 - 0 6"
sup Qr( y) =
Iyl""
for any 8 E (0, 17]. Mimic the proof of 6.4.3.
6.17. Sometimes (see 6.4.7, 12.9.9, 13.8.3, and 13.10, Exercises 13.21 to
13.23; [Zd, pp. 170-175; [Z21, pp. 116-119, 158) one wishes to consider the
set L ~ of measurable functions 1 such that <1>( I1 I) E 11, <1> being a fixed non-
negative function of a suitable type, and one wants to know that the set T
of trigonometric polynomials is contained and everywhere dense in L~ in
the sense that, given any 1 E L~ and e > 0, there exists t E T such that
In this chapter we assemble a few results about two special types of series,
namely,
Yz ao + 2:"" an cos nx = 2:
n=1 neZ
cne'n:r , (0)
2:"" an sin nx = 2:
n= 1 neZ
cne'n:r, (8)
where Cn = (1/2i) sgn n alnl' We shall assume throughout that the an are
real-valued, and write SN and aN for the Nth partial sum and the Nth Cesaro
mean, respectively, of whichever series happens to be under discussion.
The series (C) and (8) are examples of so called conjugate series, a topic to
which we return in Section 12.8.
A primary concern will be the determination of conditions under which
these series are Fourier-Lebesgue series. While the results are rather special,
inasmuch as they assume heavy restrictions on the sequence (an), they
frequently play an important role in handling questions of general significance
(as, for example, in Section 7.5).
For an extended study of more special series, see [Zd, Chapter V.
The results we shall obtain can be easily recast into statements about the
definition and nature of ~ (see Section 2.5) for rather special functions cp on Z.
sin (N + Yz)x
(7.1.1)
= 2sinYzx '
109
110 SOME SPECIAL SERIES AND THEIR APPLICATIONS
the symbol - marks the passage to the so-called conjugate series, a topic
dealt with in Section 12.8 in some generality.
The modified kernels defined by
(7.1.4)
and
. 1 - cos Nx
Y:! D N#(X) = Y:! DN(X) - Y:! sm N x = 2t Yl
an 2X
(7.1.5)
play useful, if transient, roles. Notice that DN# is even, and that DN# is odd.
Moreover,
(0 < x < 1T), (7.1.6)
and
(0 < x < 1T). (7.1.7)
(7.1.8)
(7.1.9)
(7.1.10)
(7.1.11)
as N ---+ 00; see Exercise 12.20 and [Zd, pp. 49, 67.
[as is the case with (e)]; only minor changes are needed to take care of
sequences indexed from n = 1 [as is the case with (8)].
For any real-valued sequence (an):'=o we define the sequence of differences
~an = an - a n+ 1 In particular, then, (an) is decreasing (in the wide sense),
if and only if ~a" ~ O. The sequence of second differences is defined by
2: (n + 1)1 ~2a,,1
<Xl
< 00;
,,=0
of bounded variation (BV) if and only if
2: l~a,,1
<Xl
< 00.
,,=0
It is worth explaining at once that the intervention of the difference
sequences is explained by frequent use of the technique of partial summation,
which has already been used without special comment in earlier chapters;
see also [H], pp. 97 ff. Given two sequences (a,,) and (b,,), the formula for
partial summation reads
2:
p~,,~q
a"b" = aqBq - apBp_l + L
p~,,~q-l
~a,,' B,,; (7.1.12)
where r is any fixed integer satisfying r ~ p and such that a" and btl are
defined for n ~ r; an empty sum (that is, a sum extending over a range that
is empty) is always understood to have the value zero. Repetition of the
technique introduces the second differences ~2an'
The following simple result about convex sequences helps to illuminate
7.3.1 to follow.
L'" (n + 1) .~2a" = ao -
,,=0
lim a".
n-co
(7.1.14)
(3) If (an) is quasiconvex and convergent (to a finite limit), then (an) is
BVand (7.1.13) is true.
Proof. (1) Convexity signifies that (&I n ) is decreasing. If &J", = C < 0,
then ~an ~ c for n ~ m, in which case an __ 00 as n -- 00, contrary to the
assumed boundedness of (an). So ~an ~ 0, (an) is decreasing, and therefore
boundedness implies that a = lim an exists finitely. Now
(7.1.15)
which we now know to be a series of nonnegative terms converging to
ao - a. Since also ~(~an) ~ 0, it follows easily that n.~an __ 0, which is
(7.1.13). Finally, (7.1.14) follows on applying partial summation to (7.1.15)
and using (7.1.13).
(2) We have
(7.1.16)
[This may be proved by induction on n, thus: It is trivially true for n = o.
If it is true for n, then
+ (n + 2)~am+n+l;
2:
n=O
Ibnl ~ (q + l)/bq l + 2: nl6.b l
n~O
n
[7.1] SOME PRELIMINARIES 113
q 2
,,=0 n=O
2: (k + 1)la a/cl
2"
I(n + 2)aa2"u1 ~ la"+l - a211+21 + 2
/czn+l
In Section 7.4 we shall need to know that there exist sequences (an) which
are positive and convex, and which tend to zero arbitrarily slowly. How
such sequences may be constructed will appear from the next two results.
7.1.4. Let a > 1 and suppose that (Nk)r=l is a strictly increasing sequence
of positive integers such that
N2 ~ [1 + Yz(a - 1)-l]Nl'
(7.1.18)
(k + I)Nl<+l ~ 2kN/c - (k - I)N/C-l (k = 2,3, ... ).
If (an) is the sequence defined so that ao == a, an = 11k for n = N /c (k
= 1,2, .. ), and so as to be linear for values of n satisfying 0 < n < Nl or
Nk < n < N/c+l(k = 1,2, ... ),then(an)ispositive,decreasestozero,andis
convex.
Proof. It is evident that (an) is positive and decreases to zero. Conditions
(7.1.18) express the convexity of (an), :which amounts to saying that the
negatives of the slopes of the line segments, obtained by joining neighbors
in the sequence of points of the plane having coordinates
lim c"
n-oo
= o.
114 SOME SPECIAL SERIES AND THEIR APPLICATIONS
The sequence (a,,) constructed in 7.1.4 will then sa.tisfy all requirements.
L:
P~"~Q
a"sin na: = aQ Y2D Q(x) - ap ' Y2Dp _ I (x) + L:
P~"~Q
~a". Y2D,,(x)
The statements concerning (S) follow from this and the hypotheses. Those
about (0) are proved simila.rly, using (7.1.1) in pla.ce of (7.1.3).
Conditions for the uniform convergence of (8) are not so superficial.
7.2.~. Suppose that a" t o. Then
(1) the series (S) is uniformly convergent, if and only if M" = 0(1);
(2) the series (8) is bounded1y convergent if and only if nan = 0(1);
(3) the series (8) is the Fourier series' of a continuous function if and only
if na ll _ 0;
(4) the series (8) is the Fourier series of a function in La> if and only if
M" = 0(1).
Proof. (1) Suppose first that (8) is uniformly convergent. Putting
x = 'tT/2N we have
i
[YsN)+ I
all sin nx ~ sin (i) .aN' iI,
(YoN] + I
L a" sin nx = L
,,>m m.. ,,<m+N
+ L
">,,,+N
= R + R'.
Then
IRI = I L
m",,<m+N
a"sinnxl ~ X'
"'''"<,,,+N
L naIl
~ 2a",+N'IT
x
~ 2(N + l)a"'+N ~ 2b",. (7.2.2)
aN ( ; ) ~ L (I __
n .. YoN"
a
N + 1
n ). ~ 7Tn.
'IT N
So (7.2.3) entails
N-l 2: na,,_O,
" .. YoN
hence
N-1al'/.NJ L n_ 0,
" .. YoN
and so finally NaN _ O.
116 SOME SPECIAL SERIES AND THEIR APPLICATIONS
(4) The proof is entirely similar to that of (3), using (2) and the fact that
IluNII",
= 0(1), if 2::'.. 1 an sin nx is the Fourier series of the bounded
(measurable) function/; see (6.4.9).
Remark. In cases (3) and (4) the proofs show that any function, of
which (8) is the Fourier series, is equal almost everywhere to the sum
function of the series (8). This also follows from 6.4.5.
7.2.3. Define
~n = n L
m~"
I~aml
Then
(1) if an =0(1) and ~" = 0(1), (8) is boundedly convergent;
(2) if an =0(1) and ~n = 0(1), (8) is uniformly convergent.
Proof. We may and will suppose that 0 < x ~ 'TT. Denote by N = N:z;
the positive integer such that
'TT 'TT
-- <x~-'
N+ 1 N
Then, if mE {I, 2, ... },
L an sin nx =
n~m
L
m~n<m+N
+ L
m~m+N
= R+ R'.
Here,
IRI ~ L
m~"<m+"
la"lonx
~ x L
m~n<m+N
nlanl = NxN-l L
m~"<m+N
nlanl
~ Nxoa m ~ 'TTa m (7.2.4)
and [using partial summation and noting that necessarily an = 0(1))
IR'I = IL"~m+"
~n Y2l>,,(x) - am+NoY2l>m+N-1(X) I
~ L
n~m+N
l~anl'TTX-1 + lam+Nl o'TTX- 1
~ (lam+NI + L
n~m+N
l~a"I)(N + 1).
Since
a" ~ nlanl for all n E {I, 2, .},
[7.3] THE SERIES (C) AND (8) AS FOURIER SERIES 117
therefore
Hence
IR'I ~ IXm+N + (m + N) 2:
n;om+N
I~anl
2: (n + 1)1~2anl <
ao
00. (7.3.2)
n=O
2: (n + 1)~2an' Y2 F n(x)
co
8N(X) - !(x) = (7.3.3)
n=O
pointwise for x t=. 0 (mod 27T),! E L1 and {OJ is the FS ofJ. The hypotheses are
satisfied if an t 0 and (an) is convex (see 7.1.3) in which case I ~ o.
Proof. Two applications of partial summation yield, via the formulas in
7.1.1, the equality
N-2
8N(X) = 2:
n=O
(n + 1)~2an'Y2Fn(x) + Y2N~aN_lFN_l(X) + Y~NDN(X).
(7.3.4)
By 7.1.3(3), n~an - 0, and so pointwise convergence for x t=. 0 (mod 27T) is
clear from (7.1.1), (7.1.2) and (7.3.4). Since also IIFnill = 1, the series
2: (n + 1)~2an' Y2 F n
co
(7.3.5)
n=O
converges in L1 to J. [We are not here asserting that 8NI in L1; see 7.3.2(1).]
It remains to show that (0) is the FS off. We may and will assume without
loss of generality that a o == o. Consider
2: n 2: a~ sin na.
ao co
g{x) = -1 an sin nx = (7.3.6)
n=1 n-1
118 SOME SPECIAL SERIES AND THEIR APPLICATIONS
a: = sup L1 ~
mla:1 = sup L1 ~
laml
k ~"<m<"+k k ~"<m<"+k
~ sup
m;lo"
laml-O,
aa*" = aa"
-n + n.
-..+
(I
1 -
n - --
n
I)
+1 ,
and so
= f.
o
Z (n + l)a~a"o%F,,(y)ay + %NaaN_l of.z F
ZN-~
,,-0 0
N _ 1 (y)ay
+ Y~N LZ
DN(y) ay.
f:
So
IlsNlb = 0(1).
From this one may conclude (see 12.5.2 and 12.7.5) that (0) is the Fourier
(-Stieltjes) series of a measure.
See also Teljakowski [2]; MR 48 # 794; 52 # 1480Sa,b; 54 # 13436.
(3) It is easily shown (see Exercise 7.5) that, if p > 1,
(7.3.9)
where (lIp) + (lIp') = 1. From this and (7.3.4) it appears that the con-
ditions
n1-<l/p)a ll = 0(1), (7.3.10)
n 2 -(1/p) Aa ll = 0(1) (7.3.11)
.
Ln2-(1/p)IA2alll < 00 (7.3.12)
11'"0
case it may be shown (see 12.7.6) that (O) is the Fourier series of a function
inll'.
For the case in which an t 0, a sharper result appears in 7.3.5(2}.
(4) If (an) is quasiconvex and an = 0(1}, (O) is the Fourier series of the
function I ELl; it converges in Ll to I, if and only if (7.3;7) holds; and
L: n -la
N
gN(X) = n sin nx,
n=l
g(x) - g( 8) = i% I(y) dy
for 0 < 8 ~ x ~ 'IT. By (7.3.3) and (7.3.2),feLl. Since also g is continuous
and g(O) = 0, it follows that
g(x) = 1% I(y) dy
[7.3] THE SERIES (C) AND (S) AS FOURIER SERIES 121
for O:s;; x :s;; 17. A similar argument yields the same equation for
-17 :s;; X :s;; O. From this point on, the argument proceeds as before.
Moreover, if an t 0, one can even dispense to some extent with (7.3.2),
though one cannot now conclude that f ELI. However, it will still be true
that
for -17 :s;; Xo :s;; x < 0 and for 0 < Xo :s;; x :s;; 17. Because of this one may
integrate by parts in the formula
to conclude that
recall that g is continuous and g(O) = 0, and that we are assuming (as before)
that ao = o. In other words, (0) is the series
L /(n)e
neZ
ln ""
where now
is a Cauchy principal value. Thus (0) is, in this wider sense, still the Fourier
series of f.
We next turn to a few analogous results for the sine series (8).
L a"sinnx,
co
f(x) =
n=1
L 2an
co
n=1
<00, (7.3.13)
Proof. The reader will verify easily that, under the stated conditions on
(a,,), (7.3.13) is equivalent to
n ' (7.3.15)
,,=1
If we replace herein D" by D,.# we obtain a function
L: Aa" %D,,#(x}
co
f#(x} = (7.3.16)
n=l
which differs from f by
L:
'"
,,=1
Aa,,' %sin nx,
L: IAa,,1 <
CD
00.
,,=1
Thus fELl if and only if f# ELl.
On the other hand, since D,,# is odd and is nonnegative on (0,11) [see
(7.1.5)], f# E Ll if and only if
Then 2:'=1 a" sin nx, although everywhere convergent, is not a Fourier-
Lebesgue series.
Proof. This follows immediately on combining 5.3.1,6.4.5, and 7.3.3.
[7.3] THE SERIES (C) AND (8) AS FOURIER SERIES 123
7.3.5 Supplements.
(i) 7.3.3 can be generalized ..For example, Teljakowskii [1] has proved that,
if (an) is quasiconvex and an ~ 0, then (8) is a Fourier series, if and only if
Also, Kano and Uchiyama [1] have proved that, if (an) is quasiconvex and
bounded, then (8) converges in Ll, if and only if
L: n-Ilanl <
II)
00 and anlogn~O,
n=1
while
if and only if
L: n-Ilanl <
II)
(7.3.18)
n=1
124 SOME SPECIAL SERIES A..~D THEIR APPLICATIONS
where
(Ta)" = n-1(a1 + ... + an) for n E {I, 2, ... }.
On the other hand, G. and S. Goes [1] proved that if (an) is BV and an ~ 0,
then (7.3.18) is a Fourier series, if and only if
L: n-1lanl
co
< 00;
,,=1
(7.5.3)
L'" aNII! -
Nal
FN *!II < 00, (7.5.4)
and
(7.5.5)
'"
g =/+ L
N=l
aN(j- FN*f) (7.5.6)
g = J. [1 + L
'"
aN(1 - PN)); (7.5.7)
N=l
the reader will observe that, for a fixed nEZ, 1 - PN(n) = O(ljN) for
large N, so that (7.5.5) ensures the pointwise convergence on Z of the series
appearing in (7.5.7).
In view of 2.4.1, (3.1.5), and (7.5.7), to establish (7.5.1) for some hE Lt, it
suffices to show that the sequence (a")"EZ defined by
(7.5.8)
Noting that a" = a\II\' 7.3.1 shows that we have only to verify the following
two points:
(1) all t 0 as n too;
(2) (a"),, .. o is convex.
As to (1), we have
1 __ n
ifO~n..;;N,
PN(n) = {0 N +1
ifn > N,
asn~oo.
(7.5.9)
a.nd so, since aN ~ 0 and bll ..;; bn+ 1 ~ bn+ 2' it is clear that (7.5.9) holds.
Thus (2) is satisfied a.nd the proof is complete.
In the second place, having arranged that the sequence (a,,),,>o decreases to
zero and is convex, equation (7.3.4) shows that the function h is nonnegative,
so that Ilhlll = ao and therefore
(7.5.11)
(2) With but little further effort 7.5.1 may be sharpened to the following
extent.
Suppose that El is a acompact subset of E (by which it is meant that El is
contained in the union of a countable sequence AI (i = 1, 2, ... ) of compact
subsets of E). Then there exists an hE V, which is nonnegative and satisfies
(7.5.11), with the property that to each. 1 EEl CJorresponds agE E such that
(7.5.1) holds. (The choice of h may be made the same for all 1 E Ed
The basis of the proof of this extension is the remark that, since AI is a
compact subset of E, the numbers
tend to zero as N ---+ 00 for each fixed i. (This in turn depends on the fact that.
in any metric space, any compact set can be covered by a finite number of
balls of arbitrarily small radius; this remark is used in conjunction with the
observation that 111 - F N .111 ~ 2111 I for all N and all 1 E E.) This being so,
the numbers aN ;;. 0 are chosen to satisfy (7.5.3) and (7.5.5), while in place of
(7.5.4) we impose (as we may) the demand that
( 7.5.4')
Cl = C. AC, (7.5.12)
EXERCISES
7.1. Show that if '2:'=0 e" is convergent, and if '2:'=0 IDaA,,1 < 00, then
'2:'=0 A"e" is convergent.
7.2. Prove the converse of the result in Exercise 7.1, that is, that if
'2:'=0 A"e" converges whenever '2:'=0 e" converges, then '2:'=0 IDaA,,1 < 00.
Hint: Show that the hypothesis means that '2:'=0 DaA" 8" is convergent
for any sequence 8" _ O.
7,3. Prove that, if '2:'= 0 e,,/log (2 + n) is convergent, then '2:'=0 e,,/(l + n)'"
is convergent for any a > O.
7.4. Write 8" = '2~=0 ele and ale = (n + 1)-1(80 + 81 + ... + 8,,). If we
are given that (a,,) is a bounded sequence and that 8" = o(log n) as n-oo,
prove that '2:'=0 e,,/log (2 + n) and '2:'=0 e,,/(l + n)'" are convergent for
a > O.
Hint: Write A" = 1/log (2 + n) and apply partial summation twice in
succession to the series '2:'= 0 A"e"
7.5. Verify that, if p > 1,
as N _00, where l/p + l/p' = 1 and where A,. and B,. are positive and
independent of N. [Compare (7.3.9).]
7.6. Suppose that a(x) is defined and continuous for x ~ 0, and that the
second derivative a"(x) exists and is nonnegative for x > O. Show that the
sequence (a(n:'=o is convex.
7.7. Show that '2:'=2 cos nx/log n is the Fourier series of an integrable
function! ~ 0, but that '2:'=2 sin nx/log n is not a Fourier-Lebesgue series at
all. (This example will gain in significance in 12.8.3.)
7.S. Using 7.3.2(3), show that if 0 < IX ~ 1, then '2:'=1 cos nx/n'" is the
Fourier series of a nonnegative function! E V' provided 1 ~ P < (1 - a)-l,
and that then
lim
N..... II! - i:
"=1
cos ,"nxll
n ,.
= O.
Cm = cm -1 * AC = C * AC * ... * AC
~
m (acton
for m = 1,2,.
EXERCISES 129
7.11. Prove that [with the notation used in Remark (4) following 2.3.5 and
7.5.4(3)]
(1) AC = Ll * AC;
(2) AC = L1 * BV.
Note: From (2) we see that Ll * BV is a proper subset ofBV. In proving
that L1 * BV c AC, it is useful to note that V (J) = I Df III for each trigono-
metrio polynomial f. (Actually, V(J) = IIDflll is true for each absolutely
continuous f; see [Na], p. 259.)
CHAPTER 8
Fourier Series in L2
(j, g) = 211T f
jg dx, (S.l)
(n E Z) (S.2)
form an orthonormal base in V. This last means that the family (en) is
orthonormal, in the sense that
(m,nEZ), (S.3)
and that
a.e. (S.4)
see, for example, [E], Corollary 1.12.5, or [HS], pp. 245-246, or [AB], pp.
239-240.
Despite this ready-made solution to the problem before us, we shall not
assume a knowledge of Hilbert space and will give all the necessary details
pertaining to the present situation. For general orthogonal expansions, see
[KSt], Kapitel III.
With the exception of Section S.6, what little we have to say about point-
wise convergence is included in Chapter 10.
130
[8.2] MEAN CONVERGENCE OF FOURIER SERIES IN L' 131
(8.Ll)
lim II! -
N-QI)
8N ! 112 = lim II! - UN!112
N_co
=0 (8.2.1)
and
(8.2.2)
II! - L: i(n)e,,112 ~ e.
neF
(8.2.3)
132 FOURIER SERIES IN L2
lim
N->
111 - aNl112 = 0,
111 - sJ112 2 = L
neZIF
Ij(n)i2.
(j, g) == 21 jl(x)g(x) dx =
7r
L j(n)g(n) ,
neZ
(8.2.4)
21 jl{x)g(x) dx
7r
= L j(n)g(-n),
neZ
(8.2.5)
the orthogonality relations (8.3) yield for 111 < N the equality
1 8M - 8.'11112 2 = L
M<lnl<;!V
Ic l2
n
8.3.2. Remarks. The result 8.3.1 is known to be the best possible in the
following sense: given any sequence (cn)nez for which
it is possible to choose the signs in such a way that the series L cne tnx is
not the Fourier series of- any integrable function (nor even the Fourier-
Stieltjes series of any measure, as defined in 12.5.2). Such questions will
receive more attention in Chapter 14; see especially 14.3.5 and 14.3.6.
It can also be shown that the relation Lnez Ij(n)j2 < 00, valid for fE L2,
cannot be much improved even for continuous fun.ctions f. For example (see
Exercise 8.9), given any positive function w on Z such that lim 1nl _ oo w(n)
= 00, there exist continuous functions f such that
neZ
L w(n)jj(nW = 00. (8.3.4)
L IJ(n)12-e. = 00
neZ
where,B > 1 and c > 0; see [Zl]' p. 200, Theorem (4.11); this example is quite
different in nature from those indicated in Exercise 15.13. See also [Ba l ],
p.337.
For some valid extensions of 8.3.1, see Sections 13.5 and 13.11. Regarding
extensions of 8.2.1 and 8.3.1 to general groups, see 13.5.2.
8.3.3. Dual Version of 8.3.1. As has been hinted at in Section 2.5, 8.3.1
can be recast into the form of a dual result. It says in effect that, if t/J E 12 ,
then the trigonometric polynomials
(8.5.1)
WpJ(a + b) I
= Tauf - flip ~ IITa+d - Tbfllp + IITbf - flip
= wpf(a) + wl,j(b) , (8.5.3)
It has been seen in Exercises 5.1 and 5.2 that restrictions on the rate of
decrease of wd(a) as a -+ 0 bear upon smoothness properties ofJ. It will now
be seen, by using the results of 8.3, that further results of this nature may be
expressed in terms of wd and w2f. Similar and much more elaborate results
will be mentioned at the end of 10.4.6.; see also 10.6.2.
then f E L2 and
(8.5.5)
\ 2: Inj(n)12 ~ [w 2f (a)]2
'TT Inl""la a
~ 2:
nez
Inj(n)12. (8.5.6)
8.5.4. Suppose thatf E L2. Then the following four conditions are equivalent:
(1) after correction on a null set, f is absolutely continuous and Df E L2;
(2) Lnez Inj(n)J2 < 00;
(3) lima.:.oa-1(T -af - f) exists in mean in 12;
(4) w2f(a)/a = 0(1) as a-+O.
If anyone of these conditions is fulfilled, the limit mentioned in (3) is DJ.
Proof. That (1) implies (2) follows directly from 2.3.4 and (8.2.2).
Assuming (2), 8.3.1 ensures that there exist g E L2 such that g(n) = inj(n)
for all n E Z. Then 6.2.8 shows that
L j(n)(elnb -
nez
elna ) = [b g(x) dx
Ja
(8.5.8)
for all a and b. Moreover (2) entails that Lnez Ij(n)1 < 00, and 2.4.2 shows
that, after correcting f on a null set,
f(x) = 2: j(n)eln:r:
neZ
for all x. So (8.5.8) and Lebesgue's theorem on the derivation of integrals
([W), Theorem 5.2g) combine to show that f is absolutely continuous and
[8.6] CONCERNING SUBSEQUENCES OF 8NI 137
= L: la-1(e
neZ
'na - 1) - inI211(n)i2. (8.5.9)
Now
lim a-1(e 1na - 1) - in = 0 (n E Z),
a-O
i
r=1
fgr < 00,
then ([W), Theorem 4.1e) ~;o..1 gr is integrable, hence finite. valued almost
everywhere, and so gr -+ 0 almost everywhere. In view of these two remarks,
it will be sufficient to show that
...
S = L IlaNJ - BN"JII22 < 00. (8.6.1)
"'-1
Now, by (8.2.2),
with the understanding that No be read &8 O. On the other hand, since (N/c) is a
Hadamard sequence, it is easily seen that
L N m -2 ~ O'Nj-lI,
m;.j
[8.7] A(Z) ONCE AGAIN 139
(8.6.3)
Moreover,
(8.6.1) follows and the proof is complete. More general results of similar nature
are proved in [Z2], Chapter XV; see especially p. 231.
L:
aD
in V 3, where
BU = L {,,(n - k) - "(n)}e,,.
1"lsN
So (see the proof of Theorem 4.58 of [WJ) one can choose sequences (N,(Ie');"..1
(k = 2,4, ... ) of integers such that (N~Ic+1::'1 is 8 subsequence of (N,(Ie'):"_1
and
140 FOURIER SERIES IN L2
for almost all x and all k = 1,2,.. Since; is known to belong to co(Z), this
last relation may be written
for almost all x and all k = 2, 4, .... This relation makes it plain that there
exists a measurable periodic function I such that
a.e., (8.7.3)
which, by (8.7.3), is equal toJ(n - k) - J(n). If we let m_oo and use 2.3.8
and the hypothesis that; e co(Z), it appears that; = J. The proof is complete.
[8.7] A(Z) ONCE AGAIN 141
where f3 is a positive real number and the supremum is taken with respect to
all strictly increasing finite sequences (nle)~ ~ 0 of integers. The supremum
may, of course, be 00.
It is evident that V 1(4)) = Ila4> I 1, where the difference operator a is
defined as in 7.1.2. Also, if 4> is bounded,
for some IX > 2, and that V8 (4)) < 00 for some f3 satisfying 1 ~ f3 < 2. Then,
for k = 1, 2, 3, .. "
\I T 1e4> - 4>1\2 ~. A a 8 V 8(4))BCa-2l12Ca-8lKaC2-8)12(a-/l l kCa - 2 )(2Ca-Bl. (8.7.7)
Beside this,
Ie-I
L:
"eZ
1<I>(n) - 4>(n + kW = 2 L:
1-0 mEZ
I~(mk + j) -4> (mk + j - k)I'
(8.7.9)
and, by (8.7.6),
The estimate (8.7.7) follows at once on combining (8.7.8), (8.7.9), and (8.7.10).
The final statement is a consequence of (8.7.7) and 8.7.1.
142 FOURIER SERIES IN L2
EXERCISES
Tf = 2: /(a(n))e".
"EZ
Verify that T is an isometric isomorphism of the convolution algebra L2 onto
itself.
Construct permutations a of Z such that T is not the restriction to V~ of
any homomorphism of the convolution algebra Ll into itself.
Hint: For the second part refer to conditions (3) and (4) of 4.2.6, and
show how to construct permutations a of Z with the property that, for any
integer q > 0, the relation,
lim
k~'"
lIakllll/k = lIa I ""
8.8. The spectral radiU8 formula for L2 (see Exercise 3.12 and Subsection
11.4.14). Show that, iff E VI', then
au
u(X, 0) = f(x), at (x, 0) = g(x) ,
2: Pn =
"eZ
00 (1)
and
2: p,,2 <
"eZ
00. (2)
A(f) = LJf(-x)g(X)dX.
Use (3) and Appendix B.5.1 to show that there exists ex E tl(Z) such that
A(f) = L ex(n)Pn-1!(n)
neN
(fEE); (4)
in doing this you will need to verify that each continuous linear functional on
co(Z) (see 2.2.5) is expressible as
q, --+ L ex(n)q,(n)
neZ
for some ex E tl(Z). Deduce a contradiction of (1) from (4), using Exercise
3.14 on the way.
Remarks. The condition (2) is not essential: it is included merely to
shorten the proof somewhat. One can in any case show that f may be chosen
to satisfy (i) and (ii) and to belong to V for every p < <Xl; for details, see
Edwards [3].
8.13. Suppose that f E L2. By applying. the Parseval formula to the
function
show that
lim r
r- co
L 1!(n)j2 sin (n217)r
neZ
2 = O. (2)
Hint: Observe that if If(~ + 0) - f(~ - 0)1 = d > 0, then, for all
large r and almost all x, at least one term in the sum appearing on the right-
hand side of (1) contributes an amount not less than (d/3)2,
EXERCISES 145
N-'"
lim N-l L
Inl .. N
Inj(n)1 = 0; (3)
see, for example, [Bad, pp. 214-215. In this final form the criterion of
continuity for functions of bounded variation is due to Wiener. It follows in
particular tha t any function f of bounded variation, for which j (n) = o( I / Inil,
is continuous. Compare with the remarks following 2.3.6. See also MR 38
# 487; 44 # 7220.
S.14. (I) Suppose that f, g, k e L2. Prove the following extension of
Appolonius' identity:
For this, as well as many other related results, see [Hel] and [Ho]. Compare
also Exercise 15.17.
8.16. The so-called isoperimetric problem is that of determining, among all
closed plane curves with given perimeter (here conveniently chosen to have
the value 27T), that (or those) enclosing the greatest possible area.
Assuming sufficient smoothness on the part of the admissible curves, and
expressing these curves in terms of arc length s as parameter, the problem
can be formulated analytically as follows: among all pairs (j, g) of real-valued,
absolutely continuous periodic functions I and g such that DI, Dg E L2 and
(Df)2 + (Dg)2 = 1 a.e.,
is a maximum.
Use the Parseval formula to solve this version of the problem.
Notes: For a discussion of the isoperimetric problem for plane polygons,
see [Ka], pp. 27-34. Chapter VII of [HLP] contains a brief account of the
classical approach to this and some similar problems by the methods of the
calculus of variations. Two other problems of this sort that have had a pro-
found influence on mathematics are the Dirichlet problem and the Plateau
problem; see [CH] and the references cited there, [E], Section 5.13, and [Am].
Variational methods are applicable to the study of many linear functional
equations, in which connection they often suffice to yield useful information
about the eigenvalue distributions in cases where the equation is not con-
veniently soluble explicitly; see, for example, [So], Chapter II. The existence
theorems appropriate to many variational problems are crucial and difficult:
they represent the concrete origins of, and the initial motivation for, the modern
study of compactness in function spaces and of numerous other functional
analytic techniques. Exercise 8.14 above includes a simple variational principle,
and Exercise 8.15 illustrates a concrete application of this.
8.17. The reader is reminded that f F dx is defined (possibly 00) for any
nonnegative measurable function F on T to be lim k ... ", f inf (F, k) dx; and
that 2"&#,, is defined (possibly 00) for any nonnegative function (c,,),,&Z
on Z to be lim k ... ", 21"I<kC",
Verify that, with these definitions, the Parseval formula
21
7T
JIII2 dx = L li(n)!2
"eZ
Define
for a E R, where (compare Exercise 8.17) 00 112 is taken to mean 00. Prove
that N is lower semi continuous (see Appendix A.4).
By using Exercise 3.16 and Appendix A.5 show that, if S either has positive
interior measure or is nonmeager, then there exist positive numbers 8 and A
such that
2: 1/(nW sin
neZ
2 Y2na ~ A
9.1 Mise-en-Scene
the importance ofthe Fourier transformation, this fact adds to the significance
of the positive definite concept.
In this brief chapter we shall deal rapidly with a part of the theory of
positive definite functions on the group T, and observe how it might be
used as an alternative approach to the 12 theory and the Parse val formula.
(see Chapter 8). The present situation, dealing as it does with a compact
Abelian group, is technically much simpler than that involving a general
locally compact group.
Another special case of the general setting is the dual one, in which the
underlying group is Z rather than T. There is a perfect Bochner theorem
applying to his case, but its discussion must be deferred until Section 12.13.
For remarks concerning the theory in a more general setting, see Section
9.4.
f * u * u*(O) = 4~2 ff
f(x - y)u(x)u( y) dx dy ;:.: 0 (9.2.1)
for each continuous function u. (The reader is reminded that u* denotes the
function x-+ u( -x); see the start of Section 2.3.)
9.2.2. The set of positive definite functions does not, of course, f.arm a
linear space. It is true, however, that any linear combination with real
nonnegative coefficients of positive definite functions is again positive
definite.
It is easily verified that any continuous character is positive definite; the
same is therefore true of any trigonometric polynomial with nonnegative
coefficients.
Further examples appear in 9.2.5.
9.2.3. A function fELl is positive definite if and only if (9.2.1) holds for
each trigonometric polynomial u.
Proof. The necessity is plain. Suppose, conversely, that (9.2.1) holds for
each trigonometric polynomial u. If u is any continuous function, choose (as
is possible by 6.1.1) a sequence (Un);'=l of trigonometric polynomials con-
verging uniformly tou. Then, by 3.1.4 and 3.1.6, f * Un * u! -+ f * u * u*
uniformly. Since, by hypothesis, f * un * u~(O) ;:.: 0, it follows that
f * u * u*(O) ;:.: 0, showing that the hypothesis of 9.2.1 is fulfilled.
150 POSITIVE DEFINITE FUNCTIONS AND BOCHNER'S THEOREM
= 2:lcn I 2j(n) ~ 0,
n
9.2.5. By using the criterion 9.2.4 it is very simple to verify that for any
g E L2 the continuous function
holds for any finite sequence (x!J)~ = 1 of real numbers and any finite sequence
(zn)~ = 1 of complex numbers.
Proof. We leave it to the reader to show that (9.2.2) implies (9.2.1),
remarking merely that it suffices to approximate the integral appearing in
(9.2.1) by Riemann sums
k
2: I(x m - xn)u(xm)u(xn)(xm - x m- 1 )(xn - x n- 1 )
m.n=l
UN/tO) = 2:
Inl .. N
j(n) (I - )nl+ I)' (9.2.5)
This, together with (9.2.4) and the fact thatj ~ 0, shows that LnEzj(n) < 00.
The equality (9.2.3) is now a consequence of 2.4.2. Further, if/ is known to be
continuous, we may appeal to 2.4.3 to infer that (9.2.3) holds everywhere.
2: j(n) =
neZ
lim
N-<Xl Inl.;;N
2: j(n) (I - Nlnll)
+
~ m;
then (9.2.3) and 9.2.4 show that 1/(x)1 ~ m for almost all x.
152 POSITIVE DEFINITE FUNCTIONS AND BOCHNER'S THEORE;\f
9.2.10. The converse of 9.2.5 can now be established with ease: any
continuous positive definite function I is expressible in the form I = g * g*
for some g E V.
In fact, by 9.2.4 and 9.2.8, j ~ 0 and
2: j(n) < 00.
neZ
From 8.3.1 it therefore follows that a function g E L2 exists such that g = j1l2,
in which case 2.3.1, (3.1.5), and 2.4.1 combine to show that I = g * g*.
The reader should compare the conjunction of 9.2.5 and its converse (just
established) with the criterion of M. Riesz mentioned in 10.6.2(4), and
applying to functions with absolutely convergent Fourier series.
deal of work has been done in this direction. Here we insert a few biblio-
graphical indications for the benefit of the interested reader with an ample
supply of ambition and energy.
What would appear to be perhaps the "natural" extension of the Bochner
theorem is the perfect version applying to any locally compact Abelian group.
For the details, see [B], pp. 220 ff: [R], p. 19; [N], pp. 404 ff.; [We], Chapitre VI;
[E], Sections 10.3 and 10.4; Bucy a.nd Maltese [1]; [Ph]. The third reference
displays the use of Banach-algebra techniques; see the remarks in 11.4.18(1).
Infinitesimal fragments of such extensions of the Bochner theorem appear in
Section 12.13 and in Exercises 12.34 and 12.35.
Slightly less complete versions of the theorem have been worked out for
non-Abelian locally compact groups; see [N], pp. 394 ff. and Godement [1],
especially pp. 50-53.
In pursuit of still more generality, Ito and M. G. Krein have devised formula-
tions of the theorem applying to functions on sets and spaces that are not
groups. For an example, see [N], pp. 427-428; for a brief survey, see [Hew],
pp. 145-149 and the references cited there.
Owing in part to impulses transmitted by mathematical physicists with
interests in the quantum theory of fields, special attention has been paid to
forms of the Bochner. theorem and of harmonic analysis in general applying
to Abelian nonlocally compact topological groups. A special, but particularly
relevant, such group is the underlying additive group of an infinite-dimensional
Hilbert space, for which case the reader should consult [G] and the references
there, and [GV] , Chapter IV. Possibly the Hilbert space example, although
especially significant, is too special to be genuinely typical-a state of affairs
due perhaps to the existing variety of possible methods of approach tending to
obscure the underlying essentials. Be that as it may, more general cases have
been examined with some success. For a nonlocally compact Abelian group
there may exist no genuine invariant (=Haa.r) integral having all the customary
properties (as described in, for example, Section 15 of[HR]). However Shah [1]
has shown that considerable progress can be made if there exists a suitable
stand-in for the missing invariant integral. Such a stand-in does exist for
Hilbert space and this is probably the underlying reason for success in this case.
For further developments, see MR 37 ## 1893, 5610, 5611; 40 # 6224;
51 # 13582; 52 # 14870; 55 ## 3678, 3679.
EXERCISES
Prove that
L (1 + n2)j(n) ~ /(0) + A.
neZ
9.3. Given that / is positive definite and that / E COO(U) (or is analytic
on U), for some open neighborhood U of 0, show that / is Coo (or analytic)
everywhere.
Hint: Use Exercise 9.2 to majorize the sums Lnez InI2kJ(n) for k
= 0,1,2,.
9.4. Prove that the (pointwise) product of two positive definite functions
in Leo is again positive definite.
9.5. A positive definite function/ E Leo is said to be minimal if, whenever
gEL'" is positive definite and such that / - g is positive definite, g is equal
almost everywhere to a scalar multiple of /. Verify that the only minimal
functions are the nonnegative scalar multiples of the continuous characters
e1nx
Note: This characterization of the continuous characters is important in
the Cp..:tanGodement account mentioned in Section 9.!.
9.6. Suppose that/ E Lao is positive definite. Show that If/is equal almost
everywhere to a positive definite function in LOO, if and only if / is equal
almost everywhere to a positive multiple of a continuous character e1nx
9.7. Show that the linear subspace of C generated by the set of continuous
positive definite functions is precisely the set A of continuous functions /
such that
L IJ(n)1
neZ
< 00.
Po}ntwise Convergence
of Fourier Series
In this chapter we shall deal rather summarily with some positive and
negative results about the pointwise convergence of Fourier series. The
reasons lor not according this topic a fuller treatment are discussed in
Chapter 1. The reader who is particularly attracted by these aspects may
consult [Z], especially Chapters II, VIII, and XIII; [Ba], especially Chapters
I, III-V, VII, IX; [HaR], especially Chapter IV; [I], pp. 23 If., pp. 103 If.;
[A]; and the work of Carleson mentioned in 10.4.5.
Our account is particularly terse in relation to the many known sufficient
conditions for convergence at a particular point. Out of a veritable multitude
of such results, increasing almost daily, we shall in fact prove only the very
familiar criteria associated with the names of Jordan and Dini, respectively.
These are perhaps the most useful aids to handling the functions that occur
naturally and with appreciable frequency.
On the other hand, partly in order to reinforce the remarks in Chapter 1
that bear upon the difficulties of characterizing Fourier series directly in
terms of pointwise convergence, and partly to exhibit some of the chara-
teristic functional analytic techniques of modern analysis, we shall devote
quite a large fraction of the chapter to telling part of the story of the mis-
behavior of Fourier series in relation to pointwise convergence.
Although our main concern throughout this book is with Fourier series of
functions on T, it is worth noting at this point that, rather late in the his-
torical development of harmonic analysis, it came to be recognized that the
classically natural groups T and its finite powers are" difficult" in respect of
questions about pointwise convergence of Fourier series. More precisely,
there are compact Abelian groups which are, from a classical viewpoint,
rather bizarre, and yet relative to which Fourier series behave in a fashion
simpler and more civilized than they do when the underlying group is T. The
simplest (infinite) instance of such a group is the so-called Cantor group rc,
which will be discussed in Chapter 14 as an aid in the study of Fourier series
of functions on T. It turns out that, if f is any continuous complex-valued
function on rc, then the Fourier series of f converges uniformly to f.
Concerning the dual aspect of the problems handled in this chapter, see
the remarks in Section 6.7.
155
156 POINTWISE CONVERGENCE OF FOURIER SERIES
The reader is again reminded that, in the absence of any statement to the
contrary, the convergence of a numerical series L:nezcn is defined to signify
the existence of a finite limit
lim -2
1
N-> 1T
i"0
g( y)D N( y) dy = %g( + 0) (10.1.1)
for any real-valued function 9 which is integrable over (0, 1T) and of bounded
variation on some right-hand neighborhood [0, S] of 0, where 0 < [j < 1T. In
doing this we may, without loss of generality, assume that 9 is increasing on
[10.1] FUNCTIONS OF BOUNDED VARIATION AND JORDAN'S TEST 157
[0, S] and that g( +0) = O. [The latter reduction is possible in view of the
obvious fact that (10.1.1) holds when g is a constant; see (5.1.3).] This being
so, the second mean value theorem of the integral calculus gives
~ (" g(y)DN(Y) dy
Jo
=~ r + ~217 J6("
Jo
5: r
21T 21T
I ~ J(6~ D N( y) dyl
21T
= I! J
1T
(6 sin (N .+ %)y dYI ~
~ 2 sm%y 1T ~
I! Jr 6
Y
sin (N + %)y dyl
Since y-l - %cosec %y is integrable over (0, 1T), 10.1.2 to follow shows that
(10.1.3)
(10.1.4)
Since g(y) cosec %y is integrable over (S, 1T), we may allow N to tend to
infinity and apply 2.3.8 to the second term. on the right in (10.1.4) and 80
conClude that
lim sup r 21
N-aJ ..". Jor" g( y)DN( y) dyl ~ e.
Since e is freely chosen, (lO.l.1) is thus established. See also Izumi [1].
10.1.2. The integral f: t- 1 sin t de is bounded for all real values of a and b,
that is,
sup
4. bell
lIb t-1sintdt\ <
4
00.
158 POINTWISE CONVERGENCE OF FOURIER SERIES
10.1.3. As has been indicated in 5.3.5, by using 2.3.6, 5.2.1 and a Tauberian
theorem of Hardy (see Exercise 5.8), 10.1.1 could be inferred directly from
6.3.1.
An examination of the proof of 10.1.1 leads to the following global version
thereof.
for any increasing function g on [0,17] such that g( +0) = O. Now the second
mean value theorem of the integral calculus gives for such g the relation
r
Proof. The function g defined by
L (in)-lJ(n)e inX ;
" .. 0
see 2.3.4. The result follows on applying Remark (2) following 10.1.4.
[10.2] REMARKS ON OTHER CRITERIA FOR CONVERGENCE 159
10.1.6. Remarks. (1) The proof of 10.1.5 can be extended to show that
L.n .. o{l.(n)/n is convergent for any measure fL (see Sections 12.2 and 12.5).
(2) From 10.1.5 it follows that L.:'=2 sin nx/log n is not a Fourier-Lebesgue
series. (Nor, by Remark (1) immediately above, is it even a so-called
Fourier-Stieltjes series; see 12.5.2.) See also 7.3.4 and Exercise 7.7.
(10.2.2)
Sincef!( y) cosec %y is integrable over (8, 17") for any 8 satisfying 0 < 8 ~ 17",
2.3.8 .shows that
it is necessary and sufficient that for some 8 satisfying 0 < 8 ~ 17" it is true
that
lim -1 f6 n(y)DN(y) dy = O. (10.2.5)
N- 00 '1T 0
10.2.2. In order that (10.2.4) be true, it is sufficient that to any e > 0 shall
correspond a number 8(e) satisfying 0 < 8(e) ~ 17" and a positive integer
N(e) such that
10.2.3. (Dini's test) In order that (10.2.4) be true, it is sufficient that for
,I:
same 8 satisfying 0 < 8'~:1rdne has
10.2.4. Remarks. (1) If (1.0.2.7) holds for any value of sat all,and iff
has at x at \vorst a 'jump;discontinuity, tl~e'value ?f s must be Y:!U(x + 0)
+ f(x - 0. .
(2) It is evident that (10.2.7) is fulfilled, with 8 = f(x), if, for example,
l(x + y) - f(x) = 0(1 ylti) for some ex > o.
(10.3.3)
f(x) = L'"
/(=1
Ct.k tpk,qk(X). (10.3.5)
= lak\ L: r-
r=1
1
it follows from (10.3.6) that the partial sums of the Fourier series of fare
unbounded at the origin.
Numerous variations may be played on the preceding construction; see
[Z1]' pp. 299-300; [Bad, 45; [Kz], p. 51, Proof B; Edwards and Price [1].
ISNf(xk)1
Pk(f) = !~~ PN log (N + 1)
[10.3] THE DIVERGENCE OF FOURIER SERIES 163
(10.3.9)
TtJ.is contradiction establishes the desired result. See also [Kz], p. 51, Proof A.
(xEE), (10.3.11)
interval. For S we choose the set specified in the proof if 10.3.2, where it is
shown that S is meager. Take f E C\S and define
everywhere. That such functions exist has been established by Stein ([1],
Theorem 6) on the basis of a general theorem which is cited in 16.2.8 and which
provides a powerful general approach to many existence theorems of this type.
Granted the existence of one such function, it is relatively simple to deduce
that they exist in abundance; see Exercise 10.21. See also Chen [1] and M. and
S.-1. Izumi [2].
Regarding similar results when the operators 8N are replaced by something
similar but more general, see Exercises 10.23 and 10.24.
For Kolmogorov's functionj, as for those in 10.3.2 and 10.3.3, it is the case
that 8Nj(X) is unbounded. In 1936 Marcinkiewicz showed that integrable
functions j exist for which 8Nj(X) is boundedly divergent for almost all x (see
[Zd, p. 308, and [Bad, pp. 430-443), but it is apparently still unknown whether
"almost all" can here be replaced by "all."
There is an elegant discussion of divergence for homogeneous Banach spaces
(including the Kolmogorov example cited above) in [Kz], pp. 55-61; see also
[Moz], Appendix C.
See also MR 33 # 6266; 35 # 3349; 39 # 7342; 41 # 8906; 51 # 13578.
10.3.5. Majorants for the sNf. The example given by Kolmogorov and
mentioned in 10.3.4 shows that, if we write
s*f(x) == sup ISNf(x)1 (~oo),
N
s*f(x) = 00
for all x. The readel' should compare this with the results quoted in 6.4.7
and relating to the majorant a*f of the Cesaro means aNf. See also 10.4.5
and Exercise 10.2, and 13.10 below.
In spite of this disconcerting situation, there are some remarkable assertions
of a type which say that, iff is real-valued and if the sNf are bounded below
in a suitable sense, then they are also bounded above in a corresponding
sense. A typical statement of this type affirms that if fEll is real-valued
and such that infN8.vf Ell, then SUPNSNf E LP for any p satisfying 0 < p < 1.
For other similar results and the relevant details, see [Z2]' pp. 173-175.
(10.3.12)
the limit being taken with respect to the norm on E. A similar definition applies
to sequences (an):=l' (The reader will recall that a not necessarily countable
family (an) of elemcnts of E forms an algebraic, or Hamel, base for E if each
166 POINTWISE CONVERGENCE OF FOURIER SERIES
Within the circle of ideas thus suggested appear certain problems concerning
the pointwise convergence of Fourier series which, while remaining intrac-
table for a very long while, provided the incentive for a great deal of
fruitful research; see 10.4.2 and 10.4.4 to 10.4.6.
[lOA] THE ORDER OF MAGNITUDE OF 8Ni 167
18Nf(x) - 81 ~ Ae log N +e
for all sufficiently large N, which implies (10.4.1).
(2) In the general case, we proceed as before save that the range of
integration (0, 1T) is divided into (0, 1T/N) and (1T/N, 1T). Using the estimates
IDN(y)\ ~ AN, \ D",(y) 1 ~ A/y, where A is again a positive absolute
constant, we find that
A -118",f(x) - 81 ~ ~ ("IN
1T Jo
In(y)IN dy +~
1T
f"
nlN
In(y)\ dy
Y
(10.4.2)
The choice of the number 8 is so far immaterial. Defining, as in (6.4.6),
Jl~~NJ(;).
Moreover, by partial integration,
168 POINTWISE CONVERGENCE OF FOURIER SERIES
Thus, by (10.4.2),
A -II SN f()
x - SI ::;; -1 Nl(7T)
- + -1(7T) If" l(y)dy
2 + - --2-' (10.4.3)
7T N 7T 7T "IN Y
is false for almost all x; and that there exists at least one f E V for which the
relation
sNf(x) = O(eN log log N) (N -+(0)
if f E LP, 1 ~ P ~ 2 (10.4.4)
for almost all x. Yet, for an equally long time, no success attended attempts to
establish (10.4.4) for the case where p > 2, or even to show that (10.4.4) holds
for fEe and p > 2; see [Z2]' pp. 161-162, 166-167. The present position will be
outlined in Subsection 10.4.5.
It is relatively simple to show on the basis of 10.4.3 that the trigonometric
series
(10.4.5)
[10.4] THE ORDER OF MAGNITUDE OF BN! 169
2:
neZ
Icn l2 log2 (1 + In/) < 00; (10.4.6)
2: !(n)e1nX
neZ [log (2 + Inl )]l/P
converges almost everywhere whenever f E LP and 1 ~ P ~ 2; and that
(10.4.5) converges almost everywhere provided
2:
neZ
Icn l2 log (1 + In/) < 00. (10.4.7)
This last result is due to Kolmogorov, and Seliverstov, and Plessner; see [Z2],
p. 163 or [Bad, p. 363.
On the other hand, it is known (see [Bad, p. 483, Problem 1) that there exist
functions fEe such that
2:
neZ
1/(n)1 2 Iog (1 + In/) < 00
and yet the Fourier series off diverges at infinitely many points.
what can be said regarding the size of the set E of points of divergence of the
aeries
[10.5] MORE ABOUT THE PARSEVAL FORMULA 171
It has turned out that, at least when W(N) increases more rapidly than log N
(compare (10.4.7)), new concepts of the thinness of sets, more refined than, that
provided by vanishing of the Lebesgue measure, come into play. Owing to a
strong formal similarity with definitions in potential theory, the appropriate
refined set functions are called capacities; see Section 12.12. The general type of
conclusion to be drawn from (10.4.11) is that E is a set of zero capacity, where
the particular type of capacity referred to depends on the weight function W.
Several results of this sort are known and due to Beurling, Salem, and Zygmund
and Temko (1940 onward). For details, see [Bad, pp. 398-414; [KS], Chapitre IV.
As has been seen in Section 8.6, if (10.4.5) is the Fourier series of some
function f E L2, then BNJ --+ f almost everywhere for any Hadamard sequence
(N Ie);:= 1 of positive integers (that is, any sequence of positive integers for
which inf N1<+l/N" > 1). Also, by the Kolmogorov-Seliverstov-Plessner
theorem, if (10.4.7) holds, then BNf--+f almost everywhere. To Salem we owe
a general investigation of such assertions, based upon hypotheses of the type
(10.4.11). This study produced further conditions on the sequence (N");:=l
sufficient to ensure that (10.4.11) entails that BNkf --+ f almost everywhere. One
such condition is that
for some A > 0; the case W(N) = log N includes the Kolmogorov.Seliverstov-
Plessner result. Perhaps even more remarkable is the fact that Salem obtained
conditions on the N" sufficient to ensure that BN.! --+ f almost everywhere for
any preassigned fELl: the condition reads
where wdis defined as in 2.3.7. The details are presented in [Bad, pp. 389-397.
21
~
f
j(x)g(x) dx = ". .
L.., j(n)g( -n).
neZ
(10.5.1)
10.5.1. Ifj ELI and g E BV, then (10.5.1) holds, the series being convergent.
Proof. We have seen in 10.1.4 that
SNg(X) - Y2{g(x + 0) + g(x - O)}
172 POINTWISE CONVERGENCE OF FOURIER SERIES
boundedly. Moreover, the right-hand side here is equal to y(x) save perhaps
for a countable (and therefore null) set of x. So ([W], Theorem 4.lb) we have
lim 21
211r flY dx = N~CIO T1'
JI' 8 NY dx
= lim
N
2: i( -n)g(n) ,
Inl .. N
IIIIIA == 2: V(n)1
neZ
< 00; (10.6.1)
'" sm nx
n~2 (n log n)'
the sum function of this series is even absolutely continuous (see 12.8.3(2))
and does not belong to A. See also Exercise 10.18.
Incidentally, since the sum function I is absolutely continuous (hence of
bounded variation), it ensues from the Remark following 10.6.2(1) below that
I satisfies no Lipschitz condition of order a > O.
(3) It will appear in 12.11.3 that the problem of determining all the
continuous linear functionals on A leads to a significant class of distributions.
[w2f(a)J2 = .2 4 sin
neZ
2 %na'IJ(n)12
and so
[w2f(2a)]2 = 4 .2 sin2 naIJ(n)l2.
neZ
= (8 + 1)-lN-1 S 7T /2
~ (4N)-171'.
Hence
y~ == .2
Inl;oN
IJ(n)[2:!":; [02f(27TN-1)]2. (10.6.3)
Moreover
IIil11 ==
neZ 1
ao n
= li(O)1 + .2 2: n- 1<1J(n)1 + IJ( -n)l)
n= 1 N= 1
= li(O)\ + 2:
'"
N=ln;oN
.2 n -l(!i(n)1 + li( -n)j)
.2 .2 n- 12V'<IJ(n)1 2 + li(-n)1 2)V2
00
:!":; li(O)1 +
r
n=ln;oN
= li(O)1 + i
N=l
2N- y.( L
\n\;oN
!1(X)1 2 )
L N-%r,v.
<Xl
= li(O)1 +2
N=l
[10.6] ABSOLUTELY CONVERGENT FOURIER SERIES 175
By (10.6.3), therefore
2: IJ(x + lea) -
4N
J(x + (k - l)a)12 ~ D.",J(2TTN-l). V(f).
k=l
Integrating with respect to x and using translation invariance of the integral,
and so
D.d(2TTN-l)2 ~ (8TTN)-1D.",J(2TTN-l). V(f).
Hence, (10.6.4) yields
2:
"" (D.",J(2- TT))Y. <
k=l
k 00, (10.6.6)
2:'"
N=2
N-1D.""f(2TTN-1)y' = 2:
""
k=22k-l"N<2k
2: N-1D.",f(2TTN-1)y'
~ L"" 2k-1.(2k-l)-1D.""J(TT2-k+2)Y.
k=2
= L
'"
k=2
D.",J(2- k + TT)Yz 2
cf. Section 6.5, where E~oolJ is written as ENf, and Exercise 6.10. Among such
results we note that of Steckin, which asserts that, if f E L2, then i E tl
whenever
2: N- YoE<J1 <
00
<Xl;
N=l
2: li(n)1
00
< <Xl
neS
for preassigned subsets S of Z. See also Yadov [1] and Zuk [1J.
(4) It follows easily from 8.2.1 and 8.3.1 that f E A if and only if it can be
expressed in the form f = u * v with u, v E L2. This criterion was first noted
by M. Riesz; unfortunately, it is difficult to apply in specific cases which are
not already decidable in more evident ways.
Somewhat similar results have been given by M. and S.-I. Izumi [4] and
others.
[10.6] ABSOLUTELY CONVERGE~T FOURIER SERIES 177
2: N-Y.,.E ! <
<Xl
N 00,
N=l
where
the infimum being taken as S ranges over all (not necessarily harmonic)
trigonometric polynomials of the form
2: cn,exp (iAnx) ,
N
S(x) =
n=l
where the Cn are complex numbers and the An are distinct real numbers; see
[Ba2]' p. 186; [Kah 2], p. 10. Compare this with the sufficient condition given
in (3), noting that evidently
L
neZ
Jj(n)1 < 00.
Fora proof, see (Zl], pp. 242 and 287. Compare this with F. and M. Riesz'
theorem mentioned in 12.8.5(4). See also Exercise 12.19.
(8) Let E denote a closed subset of T. Almost every question so far posed
in relation to A can also be posed in relation to A(E), the set of restrictions
to E of elements of A. For such variants we must refer the reader to [KS],
Chapitre X, especially pp. 130 ff and [Kah 2], p. 19.
It is of course evident that in all cases A(E) is a subset of C(E), the space
of all continuous complex. valued functions on E. Much less evident is the
fact that there are infinite closed sets E such that A(E) exhausts C(E): such
sets E are termed" Helson sets," concerning which a little more will be said
in Section 15.7; see especially Subsection 15.7.3. See also [Kah 2 ], Chapitres
III, IV, IX; MR 40 ## 630,7731; 43 # 6660; 55 ## 966,970.
178 POINTWISE CONVERGENCE OF FOURIER SERIES
10.6.3. The Modern Approach. (1) In more recent times emphasis has
been placed on structural properties of A as a whole. Thus, one of the problems
on which attention has been centered is the following: under what conditions
upon the function F, defined on some subset D of the complex plane, is it
true that Fo/EAwhenever/EAand/(R) =/(T) c D?
As a matter of fact, some aspects of this problem were first considered by
Levy as long ago as 1934; and even prior t.o this Wiener had obtained a
special case of Levy's result. Wiener showed that 1- 1 E A whenever I E A
and I is nonvanishing, while Levy showed that a function F has the property
mentioned in the preceding paragraph whenever D is open and F is analytic
at each point of D. The modern approach to these problems differs from the
original (for which see [Zl], pp. 245-247; [Ba 2 ], pp. 186-194; [Kah 2], pp. 57,
58) in the methods utilized, namely, the theory of Banach algebras. We shall
deal with these theorems by use of this technique in 11.4.17; the dual results
appear in 11.4.13 and 11.4.16.
In 1958 Katznelson discussed the necessity of Levy'S sufficient condition,
and variants and analogues of t.he Levy-Katznelson results have been
examined for cases in which the underlying group T is replaced by a more
general group. Katznelson, Helson, Rudin, and Kahane have all contributed
to these problems; see Herz [1], Rudin [2], [5], [R], Chapter 6; [Kah], Capitulos
IV to VI; [Kah 2], Chapitre VI; [Kz], Chapter VIII. Here we mention merely
the versions ofthe Levy-Katznelson results appropriate to the groups T and
Z, namely:
(a) If F is defined on [-1, 1] and F 0 I E A whenever I E A and I(T)
c [-1,1], then F is analytic on [-1,1].
(b) If F is defined on [ -1, 1] and F 0 4> E A(Z) whenever 4> E A(Z) and
4>(Z) c [ -1, 1], then F is analytic at 0 and F(O) = O.
Of these statements, (a) is due to Kat.znelson and (b) to Helson and Kahane
jointly. The assertion that F is analytic on [-1, 1] (respectively at 0)
signifies that F is extendible into a function analytic on some open subset of
the complex plane containing [-1, 1] (respectively 0).
From (a) it follows as a corollary that there exist functions I E A such that
III does not belong to A; the analogous assertion with A{Z) in place of A
follows likewise from (b). (Both of these corollaries were established a little
earlier by Kahane [2], [3].) On the other hand, Beurling has shown that if
I E A is such that 1/( n)1 ~ Cn (n = 0,1,2" .. ), where Cn .j. 0 anrl
"2.:'=0 Cn < 00, then Ifl E A; see [I], p. 78.
(2) There are analogous problems in which A(Z) is replaced by AP(Z)
== {/ : IE LP}. The case in which 2 ~ p ~ 00 is solved by Rider 12], who
shows (among other things) that a complex-valued function F on the complex
plane has the property that F 0 4> E AP(Z) whenever 4> E AP(Z) if and only if
it has the form
F(z) = az + bE + IzI2!P'c(z) ,
[10.6] ABSOLUTELY CONVERGENT FOURIER SERIES 179
defined in Subsection 10.6.2(8), see de Leeuw and Katznelson [1], McGehee [1],
and [Kah 2 ], Chapitre IX. (In the terminology introduced in Chapter 11,
A(E) is isomorphic to the quotient algebra of f1(Z) modulo the ideal IB
composed of those .p E (l(Z) such that $ vanishes on E.)
For further reading, consult MR 34 ## 3226, 3365, 3366, 4793, 4796; 35
## 670, 3373; 36 ## 1924,6880; 37 ## 6672, 6690; 41 # 744; 49 ## 5723, 7700;
50 ## 895, 5338; 51 ## 1289, 6627; 52 ## 8777, 8798, 8805, 14816; 53 ## 8791,
8792; 54 ## 856, 858, 8160, 8163.
EXERCISES
L j(n)e tnx
neZ log (2 + InJ)
is uniformly convergent.
Hints: Use the preceding exercise and examine the proof of 10.4.3.
Alternatively, put S(f) = Lnezj(n)/log (2 + In!> and S,v(f) = L'n, . . ,vj(n)/
log (2 + In!). Verify that SN (f) ---;. S(f) for each fEe and use the uniform
EXERCISES 181
0
+ Y.'x sin t dt
--} =
t
ouniformly forO ~ x ~ 17';
(3) I1m . (a)
SNJ N = la sintdt
--
t
for a > 0;
1" sintdt
N-oo ~ 0
sup
(4) I1m SNJ
. ()
x ~ --
N-oo,x-+O 0 t
>Y217'=j(+O).
Notes: Conclusion (4) shows that the sequence (SNj) offunctions exhibits
the so-called Gibbs phenomenon on right-hand neighborhoods of zero. A
182 POINTWISE CONVERGENCE OF FOURIER SERIES
Using the uniform boundedness principle (Appendix B.2.1) and the result
(included in 10.1.4) that SUPN ISNI(O)I < 00 for IE CBV, deduce inequality
(10.1.5), that is, the existence of a number m ~ 0, independent of I, such
that
for IE CBV.
Remark. I am grateful to Professor G. Goes for the remark that, by
7.2.2(2), the series h(x) = 2 2;:'= 1 n -1 sin nx is boundedly convergent. Iff is of
bounded variation, integration by parts yields
and so
for some integer n and all x E R, and having the property that f 0 .p E A
whenever f E A. Prove that
sup Ilell<<bII A < 00.
keZ
whenever f E E\E o
Hint8: See [KSt], p. 25 or [GP], p. 153.
Remark. The results are true if E is a Frechet space; and Lp can be
replaced by more general spaces of functions.
10.21. Assuming Kolmogorov's theorem (asserting the existence of at
least one integrable function whose Fourier series diverges almost every-
where), deduce that the set of functions fELl, whose Fourier srrirs diverge
almost everywhrre, is a comeagrr subset of V (that is, is a subset of Ll whose
complement is meager and which is therefore itself nonmeager).
Hint: Use the preceding exercise ..
10.22. Employ the uniform boundedness principle (see Appendix B.2.1)
to prove the following statement: if 4> is a nonnegative function on Z such
that the set N = {n E Z : 4>(n) > O} is infinite and
217T J
UTal(x + a) da = sNI(x)
for IE T.
Defining
It is assumed that the reader is familiar with the definition and simple
properties of a metric space and its metric topology. The aim of this appendix
is to introduce the concepts of meager and nonmeager sets, to prove Baire's
"category theorem," and to give some corollaries thereof (some of which
form the basis of results given in Appendix B). See also [K], pp. 200-203, or
[HS], p. 68.
where d denotes the metric on E. (To allay possible confusion resulting from
the notation, we point out that E{x, e), although closed in the topology
defined by d, is in general not the closure of B(x, e).)
We may assume that E is nonvoid. The same is therefore true of U 1 (since
U l is everywhere dense). Choose freely Xl E U l and then el > 0 so that
(A.2.1)
If n > m we have Xn E B(xn' en) c B{xm' em), by (A.2.I), so that d(xn' xm)
< em < 11m. The sequence (xn) is thus Cauchy and so, E being complete by
hypothesis, X = lim Xn exists in E. Since Xn E B(xm' em) for n > m ~ 1, so
X E ..8(xm' em) for all m, and (A.2.1) shows then that x E U. Thus U is nonvoid.
Take now any closed ball E = B(xo, S) in E. It is easy to verify that
Un n E is everywhere dense in the complete metric space B (a subspace of E).
So, by what we have proved, n:=1 (B nUn) isnonvoid, that is, n:=lU n = U
meets E. This being so for any B, U is everywhere dense in E. This proves (1).
(2) Let M be any meager subset of E. Then we can write M = U.7'= lAn,
where An is nowhere dense. So Un = E\An is everywhere dense and open.
By (1), n:=lU n is everywhere dense. That is, E\U:=lA n is everywhere
dense, so that U:= IAn has no interior points. The same is therefore true of
Me U::lAn.
A.3 Corollary
If E is a complete metric space, it is nonmeager in itself.
A.5 A Lemma
Let E be a metric space (or a topological space) and (f1)leI an arbitrary
family of lower semi continuous functions on E. If
sup fl(x) < 00
leI
holds for each x in some nonmeager subset 8 of E, then there exists a number
m < 00 and a nonvoid open subset U of E such that
SUpft(X) ~ m (x E U).
leI
B.l.4. Remarks. (1) For a given Frechet space there are many different
defining families of seminorms.
(2) Given a linear space E and a countable family (Pk) of seminorms on E
such that (b) and (c) are fulfilled, we can topologize E in just one way as a
Frechet space for which (Pk) is a defining family of seminorms. Namely, we
agree (as a matter of definition) that the sets (B.l.I) shall constitute a base of
neighborhoods of 0; and that, for any Xo E E, the images of these sets under
the translation x -+ x + Xo shall constitute a base of neighborhoods of Xo.
The properties of seminorms ensure that in this way one does indeed obtain a
topological linear space which satisfies conditions (a) to (c).
(3) If E is a Frechet space, it can be made into a complete metric space
whose topology is identical with the initial topology on E, and this in several
ways. One way is to define the metric
(4) In a Frechet space one can always choose the defining family (Pk) so
that the index set is the set of positive integers and so that Pl ~ P2 ~ ....
(By repeating seminorms we may suppose that the original set of indices is
the set of natural numbers, define new seminorms q" = sup {Pk : 1 ~ k ~ h},
and take the (q,.) as the desired defining family.] If this be done, a neighbor-
hood base at 0 is comprised of the sets {x E E : Pk(X) < E} when k and E > 0
vary; a neighborhood base at 0 is also obtained if E is restricted to range over
any sequence of positive numbers tending to 0; or again, if the strict
inequalities Pk(X) < E are replaced throughout by Pk(X) ~ E.
(5) The product of two Frechet spaces, or a closed linear subspace of a
Frechet space, is a Frechet space. If (Pk) and (qh) are defining families for
Frechet spaces E and F, respectively, the semi norms rkh(x, y) = Pk(X) + q,,( y)
constitute a defining family for E x F.
the proof of completeness of E' is exactly like that in the special case dealt
with in 12.7.1.
for each x in a nonmeager subset S of E (in particular, if (B.2.I) holds for
each x E E), then the II are equicontinuous on E, that is, to each e >
corresponds a neighborhood U ofO in E such that x E U entails supd/l{x)l < e.
IfE is a Banach space, the conclusion reads simply suplerll/,1I < 00.
Proof. (I) In interpreting the hypotheses on Pk we agree that a + 00
= 00 + 00 = 00 for any real a ~ 0, that 0'00 = 0, and that a'OO = 00 for
any real a > 0. For k, r = 1,2" . " let Sk.r denote the set of x E E such that
Pk{X) ~ r. Since Pk is lower semicontinuous, each Sk.r is closed in E. Plainly,
S c U:'r=ISk,r so that, since S is nonmeager by hypothesis (see A.3 for the
particular case in which S = E), some Sk.r has interior points. Thus there exists
Xo E E and a neighborhood U of in E such that Pk{X) ~ r for x E Xo + U.
Then the identity x = %(xo + x) - %(xo - x) combines with the prop-
erties. of Pk to show that Pk{X) ~ r for x E U. Consequently, given e > 0,
we have Pk{X) ~ e provided that x E U1 == {r- 1 e)U. Now U 1 is again a
neighborhood of in E, because the function x -* (re -1)X is continuous from
[B.3] OPEN MAPPING AND CLOSED GRAPH THEORE;\IS 195
E into E (see the axioms of a topological linear space in B.l.l). This shows
that PIc is finite and continuous on E and so completes the proof of (1).
(2) This follows immediately from (1) on defining PIc = p for all k, where
PIc = sup
tel
iqi 0 Tt (k=I,2,.).
Remark. In the proof of B.2.2 no essential use is made of the fact that
F is a Frechet space: all that is necessary is that F be a topological linear space
whose topology can be defined in terms of continuous seminorms. In other
words, B.2.2 extends to the case in which F belongs to the category of so-
called "locally convex" topological linear spaces. For further developments,
see [Jj:], Chapter 7.
L P1(Xk)
n
P1(X) = lim P1(X 1 + ... + xn) ~ lim inf
n-CJJ n-CXl k=l
~ lim inf
n_co
L Pk(Xk) ~ L k-
n
k=l
'"
k=l
2 = r,
exists finitely for 1 ~ j ~ i andi = 1,2, '. We now take the "diagonal
subsequence" {3 defined by
This has the crucial property that ({3(n:'= 1 is a subsequence of (CLl CLj(n:'= 1
so that
lim f8(n)(XI) (B.4.1)
,,_<Xl
exists finitely for each i.
198 CONCERNING TOPOLOGICAL LINEAR SPACES
uniformly in nand n'. The existence of the limit (B.4.1) then shows that
B.4.2. We remark that there is an analogue of B.4.1 which is valid for any
(not necessarily separable) Frechet space E, and indeed in a still wider
context.
Suppose that E is any topological linear space and that (/n):.1 is any
equicontinuous sequence of linear functionals on E. Although it may not be
possible to extract a subsequence of (/n) which converges weakly in E', the
following statement is true.
There exists a continuous linear functional Ion E with the property that,
corresponding to any given e > 0, any finite subset {Xl" .. , x r } ofE, and any
integer no, there is an integer n > no for which
(i = 1,2, .. ,r).
(x E L). (B.5.I)
Proof. This may be taken verbatim from pp. 53-55 of [E], but note the
misprint on the last line of p. 54, where "~ " should read "~."
B.5.2. Let E be a Frechet space (or, indeed, any locally convex topological
linear space; see the Remark following Subsection B.2.2), A any nonvoid
subset of E, and Xo an element of E. Then Xo is the limit in E of finite linear
combinations of elements of A if and only if the following condition is
fulfilled: iff is any continuous linear functional E (that is, iff EE') such that
f(A) = {O}, thenf(xo) = O.
Proof. The" only if" assertion is trivial in view of the linearity and
continuity of J.
Suppose, conversely, that the condition is fulfilled. Let Lo denote the
closed linear subspace of E generated by A and suppose, if possible, that
Xo ~ Lo Since Lo is closed, and since the topology of E is defined by a family
of continuous semi norms , there is a continuous seminorm p on E such that
p(.y - x o) > I for all Y E L o, and hence also
B.5.3. Let E be a normed linear space and E' its dual (see B.1.7). If Xo E E,
there exists fEE' such that
Ilfll ~ 1
and
f(x o} = Ilxoll
Proof. Let L be the linear subspace of E generated by Xo. Define fa on
L by foU'.x o} = "llxoll (" E <1>). It now suffices to apply B.5.l, taking p(x} =
Ilxll: any extension f of fo of the type specified in B.5.l satisfies all
requirements.
APPENDIX C
The aim of this appendix is to give a proof of two results used in the
main text.
(C.1)
Proof. T~at any g E LP' satisfying (C.l) also satisfies (C.2), follows at
once from Holder's inequality and its converse (see Exercise 3.6). So we
confine our attention to the proof of the existence of g [its essential uniqueness
being a corollary of (C.2)]. This proof will be based upon use of the Radon-
Nikodym theorem (see [W], Chapter 6; [HS], Section 19; [AB], p. 406).
For this purpose we consider Borel subsets E of the interval X = (0,21T)
(compare [W], p. 93). The characteristic function of E may be extended by
periodicity, the result being denoted by XE and being a member of LP. The
number
(C.3)
201
202 THE DUAL OF L'(l ~ p <CXl)
the series being convergent in V' since p < 00. (Here is the major reason for
breakdown of the theorem for p = 00.) The continuity of F, together with its
linearity, shows that therefore
F(f) = L ffg dX
holds for allf which are finite linear combinations of functions XE' Knowing
this, it is easy to conclude that g E V" and that
compare Exercise 3.6. Holder's inequality shows then that the difference
Fo(f) = F(f) - L fh dx
li~ L fkl dX
exists finitely for each I] E L"'. Then there exists a function i ELI to which
(f;) is weakly convergent, that is, which is such that
L: F(u,,)t-"-I,
ao
4>(t) = (D.6)
n-O
again for sufficiently large Itl. Now (D.6) shows first that 4> can be extended
to 3:\K in such a way as to be holomorphic on a neighborhood of 00 (simply by
setting 4>(00) = 0); and second, in conjunction with (D.2), that 4> vanishes
on a neighborhood of 00. Since, by (1), 4\K is connected, the same is true of
a\K = (4\K) U {oo}, and it follows that 4> must vanish throughout the whole
of 4\K. In particular, (D.4) is true. This completes the proof.
Remarks. For a different approach to Runge'S theorem and related
questions, see [He], pp. 149-153, and the references cited there. The theorem
has ramifications and analogues extending to Riemann surfaces (see, for
example, the relevant portions of Behnke and Stein's "Theorie der analy.
tischen Funktionen einer komplexen Veranderlichen") and to functions of
several complex variables or on complex analytic manifolds (see Gunning
and Rossi's "Analytic Functions of Several Complex Variables," Chapter I,
Section F, and Chapter VII).
There are also real variable analogues of Runge's theorem, applicable to
solutions of linear partial differential equations; see [E], p. 396, and the
references cited there.
Bibliography
[Hi] HIRSCHMAN, I. I., JR. (editor). StudieB in Real and Complex AnalyBiB.
StudieB in MathematicB, Vol. 3. The Math. Association of America;
PrenticeHall, Inc., Englewood Cliffs, New Jersey (1965).
[HLP] HARDY, G. H., LITTLEWOOD, J. E. AND P6LYA, G. InequalitieB.
Cambridge University Press, New York (1934).
ALJANCIC, S.
[1] On the integral moduli of continuity in Lp (1 < P < C() of Fourier series
with monotone coefficients. Proc. Amer. Math. Soc. 17 (1966), 287-294.
BLOOM, W. R.
[1] Bernstein's inequality for locally compact Abelian groups. J. Austr. Math.
Soc. XVII(I) (1974), 88-102.
[2] A converse of Bernstein's inequality for locally compact Abelian groups.
Bull. Austr. Math. Soc. 9(2) (1973), 291-298.
[3] Jackson's theorem for locally compact Abelian groups. ibid. 10(1) (1974),
59-66.
[4] Jackson's theorem for finite products and homomorphic images of locally
compact Abelian groups. ibid. 12(2} (1975), 301-309.
BOAS, R. P. JR.
[1] Beurling's test for absolute convergence of Fourier series. Bull. Amer.
Math. Soc. 66 (1960), 24-27.
[2] Fourier Series with positive coefficients. J. Math. Anal. Appl. 17 (1967),
463-483.
BRED ON, G. E.
[1] A new treatment of the Haar integral. Mich. Math. J. 10 (1963), 365-373.
CARLESON, L.
[1] On convergence and growth of partial sums of Fourier series. Acta Math.
116 (1966), 135-157.
213
214 RESEARCH PUBLICATIONS
Cm:N, Y. M.
[1] An ahnoet everywhere divergent Fourier series of the class L(log + log + L )1-
J. London Math. Soo. 44 (1969), 643-654.
COllEN, P. J.
[1] Factorization in group algebras. Duke Math. J. 18 (1959), 199-205.
[2] On homomorphisms of group algebras. Amer. J. Math. 81 (1960), 213-226.
[3] On a conjecture of Littlewood and idempotent measures. Amer. J. Math. 81
(1960), 191-212.
[4] A note on constructive methods in Banach algebras. Proo. Amer. Math.
Soo. 11 (1961), 159-163.
EssiN, M.
[1] Studies on a convolution inequality. Ark. Mat. 5 (9) (1963), 113-152.
GoDEMENT, R.
[1] Les fonctions de type positif et 1& theorie des groupes. TraM. Amer. Math.
Soc. 63 (1948), 1-84.
GOES, G. AND GoES, S.
[1] Sequences of bounded variation and sequences of Fourier coefficients. I.
Math. Z. 118 (1970), 93-102.
HARDY, G.H.
[1] Notes on some points in the integral calculus, LXVI, The arithmetic mean
of a Fourier constant. MU8enger oj Math. 58 (1928), 50-52.
HEBz, C. S.
[1] Fonctions operant sur certains semi-groupes. O. R. Acad. Sci. Paris 155
(1962), 2046-2048.
HEWl'rr, E.
[1] The ranges of certain convolution operators. Math. Scani/,. 15 (1964),
147-155.
HEWIT.r, E. AND HEWITT, R. E.
[1) The Gibbs-Wilbraham phenomenon: an episode in Fourier analysis. To
appear The Mathematical Intelligencer.
HmscHMAN, I. I., JR.
[1] On multiplier transformations. Duke Math. J. 16 (1959), 221-242.
IZUMI, M. AND IZUMI, S.-I.
[1] Fourier series of functions of bounded variation. Proc. Japan Acad. 44 (6)
(1968), 415-417.
[2] Divergence of Fourier series. Bull_ A'UBtr. Math. Soc. 8 (1973), 289-304.
[3] On absolute convergence of Fourier series. Ark. JOr Mat. 7(12) (1967),
177-184.
[4] Absolute Convergence of Fourier series of convolution functions. J. Approz.
Theory 1 (1968), 103-109.
KAHANE, J.-P.
[1] Idempotents and closed subalgebras of A(Z). Function Algebras, pp. 198-
207. Scott, Foresman and Company, Chicago (1966).
[2] Sur certaines classes de mea de Fourier absolument convergentes. J.
Math. PUreB et Appl. 35 (1956), 249-258.
[3] Sur un probleme de Littlewood. Nederl. Akad. Wetensch. Proc. Ser. AGO =
Indag. Math. It (1957), 268-271.
[4] Sur les rtSarrangements des suites de coefficients de Fourier-Lebesgue.
O. R. Acad. Sci. Paris. Ser. A-B 265 (1967), A310-A312.
KAHANE, J.-P. AND KATZNELBON, Y.
[1] Sur les series de Fourier uniformement convergentes. O. R. Acad. Sci. Paris
181 (1965), 3025-3028.
KANo, T. AND UOHIYAMA, S.
[1] On the convergence in L of some special Fourier series. Proc. Japan Acad.
53 A(2) (1977), 72--77.
216 RESEARC.H PUBLICATIONS
MCGEHEE, O. C.
(1] Certain isomorphisms between quotients of group algebras. Notices Amer.
Math. Soc. 13 (1966), 475.
MEN'SHOV, D. E.
[1] On convergence in measure of trigonometric series. (In Russian) Trudy Mat.
Inst. im V. A. Steklova 32 (1950), 3-97.
MITJAGIN, B. S.
[1] On the absolute convergence of the series of Fourier coefficients. (In
Russian) Dokl. 4kad. Nauk SSSR 157 (1964), 1047-1050.
MUELLER, M. J.
[1] Three results for locally compact groups connected with the Haar measure
density theorem. Proe. Amer. Math. Soc. 16 (1965), 1414-1416.
NEWMAN, D. J.
[1] The non existence of projections from LI to HI. Proc. Amer. Math. Soc. 12
(1961), 98-99.
[2] An LI extremal problem for polynomials. Proc. Amer. Math. Soc. 16 (1965),
1287-1290.
RAY, K. C. AND LAHIRI, B. K.
[1] An extension of a theorem of Steinhaus. Bull. Calcutta Math. Soc. 56 (1964),
29-31.
RIDER, D.
[1] Central idempotents in group algebras. To appear Bull. Amer. Math. Soc.
[2] Transformations of Fourier coefficients. To appear Padlu; J. Math.
ROONEY, P. G.
[1] On the representation of sequences as Fourier coefficients. Proc. Amer.
Math. Soc. 11 (1960), 762-768.
RUBEL, L. A.
[1] Uniform distributions and densities on locally compact Abelian groups.
Symposia on Theor. Physics and Math., Vol. 9,183-193. Plenum Press, New
York (1969).
RUDIN, W.
[1] Representation of functions by convolutions. J. Math. Meeh. 7 (1958),
103-116.
[2] A strong converse of the WienerLevy theorem. Canad. J. Math. 14 (1962),
694-701.
[3] Idempotents in group algebras. Bull. Amer. Math. Soc. 69 (1963), 224-227.
[4] Some theorems on Fourier coefficients. Proc. Amer. Math. Soc. 10 (1959),
855-859.
[5] Positive definite sequences and absolutely monotonic functions. Duke
Math. J. 26 (1959), 617-622.
RUDIN, W. AND SCHNEIDER, H.
(1] Idempotents in group rings. Duke Math. J. 31 (1964), 585-602.
RESEARCH PUBLICATIONS 217
RYAN, R.
[I] Fourier transforms of certain classes of integrable functions. TraM. Amer.
Math. Soc. 105 (1962), 102-111.
SHAH, T S.
[I] Positive definite functions on commutative topological groups. Scientia
Sinica (5)XIV (1965), 653-665.
SRINIVASAN, T. P. AND WANG, J.-K.
[1] Weak Dirichlet Algebras. Function Algebras, pp. 216-249. Scott, Foresman
and Company, Chicago (1966).
STEIN, E. M.
[1] On limits of sequences of operators. Ann. Math. 7' (1961), 14:0-170.
TAIBLESON, M.
[1] Fourier coefficients of functions of bounded variation. Proc. Amer. Math.
Soc. 18 (1967), 766.
TELJAKowsKir, S. A.
[1] Some estimates for trigonometric series with quasiconvex coefficients
(Russian). Mat. Sbornik 63 (105) (1964), 426-444 = A.M.S. Translations (2)
86 (1970), I09-13l.
[2] On the question of convergence of Fourier series in the L-metric (Russian).
Mat. Zametki 1 (1967), 91-98. MR 34 # 8067.
TIMAN, M. F.
[I] Best approximation of a function and linear methods of summing Fourier
series. (In Russian) lzv. Akad. Nauk SSSR Ser. Mat. 29 (1965), 587-604.
WEISS, G.
[I] Harmonic Analysis. Studi68 in Mathematics, Vol. 3, pp. 124-178. The
Math. Assoc. of America; PrenticeHall, Inc., Englewood Cliffs, New
Jersey (1965).
YADAV, B. S.
[1] On a theorem of Titchmarsh. Publ. Inse. Math. (Beograd) (N.S.) 2(18)
(1962), 87-92 (1963).
YADAV, B. S. AND GOYAL, O. P.
[1] On the absolute convergence of Fourier series. J. MaharajaSayajirao Unit}.
Baroda 10 (1961), 29-34.
ZAlIlANSKY, M.
[1] Approximation et Analyse Harmonique, I. Bull. Sci. math., 2" serie, 101
(1977), 3-70.
[2] Approximation et Analyse Harmonique, II. Bull. Sci. math., 2" serie, 101
(1977), 149-180.
ZUK, V. V.
[1] On the absolute convergence of Fourier series. (In Russian) Dokl. Akad.
Nauk SSSR 180 (1965), 519-522.
Index
Nonmeager, 187
Jackson polynomial, 102 Norm, 191
Jordan's test, 149 V'" 27
11', 29
quotient, 195
Kernel, Nowhere dense, 187
Abel.Poisson, 103
conjugate, Dirichlet, 110
Dirichlet, 5.1, 79 Open linear operator, 195
Fejer,79 Open mapping theorem, 196
modified Dirichlet, 110 Order of magnitude,
Kolmogorov Seliverstov Plessner theo of 8Ni, 166 ff., 10.2
rem, 169, 171 of aNi, 97 ff., 6.18
Kolmogorov's theorem, 138, 171,10.21 Orthogonality relations, 3, 25, 26
Orthonormal base, 130