An Introduction To Banach Space Theory. R. E. Megginson PDF
An Introduction To Banach Space Theory. R. E. Megginson PDF
An Introduction To Banach Space Theory. R. E. Megginson PDF
Editorial 80ard
S. Axler FW. Gehring K.A. Ribet
An Introduction to
Banach Space Theory
Springer
Robert E. Megginson
Mathematics Department
3856 East Hall
University of Michigan
Ann Arbor, MI 48109-1109
Editorial Board
S. Axler F.W. Gehring K.A. Ribet
Mathematics Department Mathematics Department Mathematics Department
San Francisco State East Hali University of California
University University of Michigan at Berkeley
San Francisco, CA 94132 Ann Arbor, MI 48109 Berkeley, CA 94720-3840
USA USA USA
987654321
ISBN 978-1-4612-6835-2
To my mother and father
and, of course, to Kathy
Contents
Preface ix
1 Basic Concepts 1
1.1 Preliminaries 1
1.2 Norms . . . . 8
1.3 First Properties of Normed Spaces 17
1.4 Linear Operators Between Normed Spaces 24
1.5 Baire Category . . . . . . . . . 35
1.6 Three Fundamental Theorems . 41
1. 7 Quotient Spaces . . . . . . . . 49
1.8 Direct Sums . . . . . . . . . . . 59
1.9 The Hahn-Banach Extension Theorems 70
1.10 Dual Spaces . . . . . . . . . . . . 84
1.11 The Second Dual and Reflexivity 97
1.12 Separability . . . . . . . . . . . 109
*1.13 Characterizations of Reflexivity . 115
D Ultranets 541
References 547
Index 569
Preface
field. Albert Bennett came close to giving the definition of a normed space
in a 1916 paper [23] on an extension of Newton's method for finding roots,
and in 1918 Frederic Riesz [195] based a generalization of the Fredholm
theory of integral equations on the defining axioms of a complete normed
space, though he did not use these axioms to study the general theory
of such spaces. According to Jean Dieudonne [64], Riesz had at this time
considered developing a general theory of complete normed spaces, but
never published anything in this direction. In a paper that appeared in
1921, Eduard Helly [102] proved what is now called Helly's theorem for
bounded linear functionals. Along the way, he developed some of the general
theory of normed spaces, but only in the context of norms on subs paces of
the vector space of all sequences of complex scalars.
The first undisputed efforts to develop the general theory of normed
spaces appeared independently in a paper by Hans Hahn [98] and in Stefan
Banach's thesis [10], both published in 1922. Both treatments considered
only complete normed spaces. Though the growth of the general theory pro-
ceeded through the 1920s, the real impetus for the development of modern
Banach space theory was the appearance in 1932 of Banach's book Theorie
des Operations Lineaires [13], which stood for years as the standard refer-
ence work in the field and is still profitable reading for the Banach space
specialist today.
Many important reference works in the field have appeared since Ba-
nach's book, including, among others, those by Mahlon Day [56] and by
Joram Lindenstrauss and Lior Tzafriri [156, 157J. While those works are
classical starting points for the graduate student wishing to do research
in Banach space theory, they can be formidable reading for the student
who has just completed a course in measure theory, found the theory of Lp
spaces fascinating, and would like to know more about Banach spaces in
general.
The purpose of this book is to bridge that gap. Specifically, this book
is for the student who has had enough analysis and measure theory to
know the basic properties of the Lp spaces, and is designed to prepare such
a student to read the type of work mentioned above as well as some of
the current research in Banach space theory. In one sense, that makes this
book a functional analysis text, and in fact many of the classical results
of functional analysis are in here. However, those results will be applied
almost exclusively to normed spaces in general and Banach spaces in par-
ticular, allowing a much more extensive development of that theory while
placing correspondingly less emphasis on other topics that would appear
in a traditional functional analysis text.
It should be made clear that this book is an introduction to the general
theory of Banach spaces, not a detailed survey of the structure of the
classical Banach spaces. Along the way, the reader will learn quite a bit
about the classical Banach spaces from their extensive use in the theory,
examples, and exercises. Those who find their appetite for those spaces
Preface xi
Prerequisites
Appendix A contains a detailed list of the prerequisites for reading this
book. Actually, these prerequisites can be summarized very briefly: Any-
one who has studied the first third of Walter Rudin's Real and Complex
Analysis [202]' which is to say the first six chapters of that book, will be
able to read this book through, cover-to-cover, omitting nothing. Of course,
this implies that the reader has had the basic grounding in undergradu-
ate mathematics necessary to tackle Rudin's book, which should include a
first course in linear algebra. Though some knowledge of elementary topol-
ogy beyond the theory of metric spaces is assumed, the topology presented
near the beginning of Rudin's book is enough. In short, all of this book is
accessible to someone who has had a course in real and complex analysis
that includes the duality between the Lebesgue spaces Lp and Lq when
1 < p < 00 and p-l + q-l = 1, as well as the Riesz representation theorem
for bounded linear functionals on C(K) where K is a compact Hausdorff
space, and who has not slighted the usual prerequisites for such a course.
In fact, a large amount of this book is accessible at a much earlier stage
in a student's mathematical career. The real reason for the measure theory
prerequisite is to allow the reader to see applications of Banach space theory
to the Lp spaces and spaces of measures, not because the measure theory is
itself crucial to the development of the general Banach space theory in this
book. It is quite possible to use this book as the basis for an undergraduate
topics course in Banach space theory that concentrates on the metric theory
of finite-dimensional Banach spaces and the spaces fp and Co, for which
the only prerequisites are a first course in linear algebra, a first course in
real analysis without measure theory, and an introduction to metric spaces
without the more general theory of topological spaces. Appendix A explains
in detail how to do so. A list of the properties of metric spaces with which
a student in such a course should be familiar can be found in Appendix B.
Since the fp spaces are often treated in the main part of this book as special
Lp spaces, Appendix C contains a development of the fp spaces from more
basic principles for the reader not versed in the general theory of Lp spaces.
Though most of the results in this text are divided in the usual way
into lemmas, propositions, and theorems, a result that is really a theorem
occasionally masquerades as an example. For instance, the theorems of
Nikodym and Day that (Lp[O, 1])* = {O} when 0::; p < 1 appear as Exam-
ple 2.2.24, since such an Lp space is an example of a Hausdorff topological
vector space whose dual does not separate the points of the space.
The theory in this book is developed for normed spaces over both the
real and complex scalar fields. When a result holds for incomplete normed
spaces as well as Banach spaces, the result is usually stated and proved in
the more general form so that the reader will know where completeness is
truly essential. However, results that can be extended from Banach spaces
or arbitrary normed spaces to larger classes of topological vector spaces
usually do not get the more general treatment unless the extension has a
specific application to Banach space theory.
Any extensive treatment of the theory of normed spaces does require
the study of two vector topologies that are not in general even metrizable,
namely, the weak and weak* topologies. Much of the sequential topological
intuition developed in the study of normed spaces can be extended to non-
metrizable vector topologies through the use of nets, so nets playa major
role in many of the topological arguments given in this book. Since the
reader might not be familiar with these objects, an extensive development
of the theory of nets is given in the first section of Chapter 2.
This book is sprinkled liberally with examples, both to show the theory
at work and to illustrate why certain hypotheses in theorems are necessary.
This book is also sprinkled liberally with historical notes and citations
of original sources, with special attention given to mentioning dates within
the body of the text so that the reader can get a feeling for the time
frame within which the different parts of Banach space theory evolved. In
ascribing credit for various results, I relied both on my own reading of
the literature and on a number of other standard reference::> to point me to
the original sources. Among those standard references, I would particularly
like to mention the excellent "Notes and Remarks" sections of Dunford and
Schwartz's book Linear Operators, Part I [67], most of which are credited
to Bob Bartle in thc introduction to that work. Anyone interested in the
rich history of this subject should read those sections in their entirety. I
hasten to add that any error in attribution in the book you are holding is
entirely mine.
In many cases, no citation of a source is given for a definition or result,
especially when a result is very basic or a definition evolved in such a way
that it is difficult to decide who should receive credit for it, as is the case
for the definition of the norm function. It must be emphasized that in no
case does the lack of a citation imply any claim to priority on my part.
The exercises, of which there are over 450, have several purposes. One
obvious one is to provide the student with some practice in the use of the
results developed in the text, and a few quite frankly have no reason for
Preface xiii
their existence beyond that. However, most do serve higher purposes. One
is to extend the theory presented in the text. For example, Banach limits
are defined and developed in Exercise 1.102. Another purpose of some of
the exercises is to provide supplementary examples and counterexamples.
Occasionally, an exercise presents an alternative development of a main
result. For example, in Exercise 1.76 the reader is guided through Hahn's
proof of the uniform boundedness principle, which is based on a gliding
hump argument and does not use the Baire category theorem in any form.
\\T!th the exception of a few extremely elementary facts presented in the
first section of Chapter 1, none of the results stated and used in the body
of the text have their proofs left as exercises. Very rarely, a portion of an
example begun in t,he body of the text is finished in the exercises.
One final comment on the general approach involves the transliteration
of Cyrillic names. I originally intended to use the modern scheme adopted
by Mathematical Revie?'lS in 1983. However, in the end I decided to write
these names as the authors themselves did in papers published in West-
ern languages, or as the names have commonly appeared in other sources.
For example, the modern MR transliteration scheme would require that
V. L. Smulian's last name be written as Shmul'yan. However, Smulian
wrote many papers in Western languages, several of which are cited in this
book, in which he gave his name the Czech diacritical transliteration that
appears in this sentence. No doubt he was just following the custom of
his time, but because of his own extensive use of the form Smulian I have
presented his name as he wrote it and would have recognized it.
Synopsis
Chapter 1 focuses on the metric theory of normed spaces. The first three
sections present fundamental definitions and examples, as well as the most
elementary properties of normed spaces such as the continuity of their
vector space operations. The fourth section contains a short development
of the most basic properties of bounded linear operators between normed
spaces, including properties of normed space isomorphisms, which are then
used to show that every finite-dimensional normed space is a Banach space.
The Baire category theorem for nonempty complete metric spaces is the
subject of Section 1.5. This section is, in a sense, optional, since none of
the results outside of optional sections of this book depend directly on it,
though some such results do depend on a weak form of the Baire category
theorem that will be mentioned in the next paragraph. However) this sec-
tion has not been marked optional, since a student far enough along in his
or her mathematical career to be reading this book should become familiar
with Baire category. This section is placed just before the section on the
open mapping theorem, closed graph theorem, and uniform boundedness
xiv Preface
are two reasons for my not doing so. The first is that I wish to emphasize
that the proof is really based only on the elementary metric theory of
Banach spaces, not on arguments involving weak compactness, and the best
way to do that is to give the proof before the weak topology has even been
defined (though I do cheat a bit by defining weak sequential convergence
without direct reference to the weak topology). The second is due to the
reputation that James's theorem has acquired as being formidably deep.
The proof is admittedly a bit intricate, but it is entirely elementary, not all
that long, and contains some very nice ideas. By placing the proof as early
as possible in this book, I hope to stress its elementary nature and dispel
a bit of the notion that it is inaccessible.
Chapter 2 deals with the weak topology of a normed space and the
weak* topology of its dual. The first section includes some topological pre-
liminaries, but is devoted primarily to a fairly extensive development of
the theory of nets, including characterizations of topological properties in
terms of the accumulation and convergence of certain nets. Even a student
with a solid first course in general topology may never have dealt with
nets, so several examples are given to illustrate both their similarities to
and differences from sequences. A motivation of the somewhat nonintuitive
definition of a subnet is given, along with examples. The section includes
a short discussion of topological groups, primarily to be able to obtain a
characterization of relative compactness in topological groups in terms of
the accumulation of nets that does not always hold in arbitrary topological
spaces. Ultranets are not discussed in this section, since they are not really
needed in the rest of this book, but a brief discussion of ultranets is given
in Appendix D for use by the instructor who wishes to show how ultranets
can be used to simplify certain compactness arguments.
Section 2.2 presents the basic properties of topological vector spaces and
locally convex spaces needed for a study of the weak and weak* topologies.
The section includes a brief introduction to the dual space of a topologi-
cal vector space, and presents the versions of the Hahn-Banach separation
theorem due to Mazur and Eidelheit as well as the consequences for locally
convex spaces of Mazur's separation theorem that parallel the consequences
for normed spaces of the normed space version of the Hahn-Banach exten-
sion theorem.
This is followed by a section on metrizable vector topologies. This section
is marked optional since the topologies of main interest in this book are
either induced by a norm or not compatible with any metric whatever. An
F -space is defined in this section to be a topological vector space whose
topology is compatible with a complete metric, without the requirement
that the metric be invariant. Included is Victor Klee's result that every
invariant metric inducing a topologically complete topology on a group is
in fact a complete metric, which has the straightforward consequence that
every F-space, as defined in this section, actually has its topology induced
by a complete invariant metric, and thereby answers a question of Banach.
xvi Preface
The versions of the open mapping theorem, closed graph theorem, and
uniform boundedness principle valid for F-spaces are given in this section.
Section 2.4 develops the properties of topologies induced by families
of functions, with special emphasis on the topology induced on a vector
space X by a subspace of the vector space of all linear functionals on X.
The study of the weak topology of a normed space begins in earnest
in Section 2.5. This section is devoted primarily to summarizing and ex-
tending the fundamental properties of this topology already developed in
more general settings earlier in this chapter, and exploring the connections
between the weak and norm topologies. Included is Mazur's theorem that
the closure and weak closure of a convex subset of a normed space are
the same. Weak sequential completeness, Schur's property, and the Radon-
Riesz property are studied briefly.
Section 2.6 introduces the weak* topology of the dual space of a normed
space. The main results of this section are the Banach-Alaoglu theorem and
Goldstine's theorem. This is followed by a section on the bounded weak*
topology of the dual space of a normed space, with the major result of this
section being the Krein-Smulian theorem on weakly* closed convex sets:
A convex subset C of the dual space X* of a Banach space X is weakly*
closed if and only if the intersection of C with every positive scalar multiple
of the closed unit ball of X* is weakly* closed.
Weak compactness is studied in Section 2.8. It was necessary to delay
this section until after Sections 2.6 and 2.7 so that several results about
the weak* topology would be available. The Eberlein-Smulian theorem is
obtained in this section, as is the result due to Krein and Smulian that the
closed convex hull of a weakly compact subset of a Banach space is itself
weakly compact. The corresponding theorem by Mazur on norm compact-
ness is also obtained, since it is an easy consequence of the same lemma
that contains the heart of the proof of the Krein-Smulian result. A brief
look is taken at weakly compactly generated normed spaces.
The goal of optional Section 2.9 is to obtain James's characterization
of weakly compact subsets of a Banach space in terms of the behavior of
bounded linear functionals. The section is relatively short since most of the
work needed to obtain this result was done in the lemmas used to prove
James's reflexivity theorem in Section 1.13.
The topic of Section 2.10 is extreme points of nonempty closed convex
subsets of Hausdorff topological vector spaces. The Krein-Milman theorem
is obtained, as is Milman's partial converse of that result.
Chapter 2 ends with an optional section on support points and subreflex-
ivity. Included are the Bishop-Phelps theorems on the density of support
points in the boundaries of closed convex subsets of Banach spaces and on
the sub reflexivity of every Banach space.
Chapter 3 contains a discussion of linear operators between normed
spaces far more extensive than the brief introduction presented in Sec-
tion 1.4. The first section of the chapter is devoted to adjoints of bounded
Preface XVll
bases are studied. The final section of Chapter 4 is optional and is devoted
to an investigation of James's space J, which was the first example of a
nonreflexive Banach space isometrically isomorphic to its second dual.
Chapter 5 focuses on various forms of rotundity, also called strict con-
vexity, and smoothness. The first section of the chapter is devoted to char-
acterizations of rotundity, its fundamental properties, and examples, in-
cluding one due to Klee that shows that rotundity is not always inherited
by quotient spaces. The next section treats uniform rotundity, and includes
the Milman-Pettis theorem as well as Clarkson's theorem that the Lp spaces
such that 1 < p < 00 are uniformly rotund. Section 5.3 is devoted to gen-
eralizations of uniform rotundity, and discusses local uniform rotundity,
weak uniform rotundity, weak* uniform rotundity, weak local uniform ro-
tundity, strong rotundity, and midpoint local uniform rotundity, as well as
the relationships between these properties.
The second half of Chapter 5 deals with smoothness. Simple smoothness
is taken up in Section 5.4, in which the property is defined in terms of
the uniqueness of support hyperplanes for the closed unit ball at points
of the unit sphere and then characterized by the Gateaux differentiabil-
ity of the norm and in several other ways. The partial duality between
rotundity and smoothness is examined, and other important properties of
smoothness are developed. Uniform smoothness is the subject of the next
section, in which the property is defined using the modulus of smoothness
and characterized in terms of the uniform Frechet differentiability of the
norm. The complete duality between uniform smoothness and uniform ro-
tundity is proved. Frechet smoothness and uniform Gateaux smoothness
are examined in the final section of the chapter, and Smulian's results on
the duality between these properties and various generalizations of uniform
rotundity are obtained.
Appendix A includes an extended description of the prerequisites for
reading this book, along with a very detailed list of the changes that must
be made to the presentation in Chapter 1 if this book is to be used for
an undergraduate topics course in Banach space theory. Appendices B
and C are included to support such a topics course. They are, respectively,
a list of the properties of metric spaces that should be familiar to a student
in such a course and a development of fp spaces from basic principles of
analysis that does not depend on the theory of Lp spaces. Appendix D is a
discussion of ultranets that supplements the material on nets in Section 2.1.
Dependences
No material in any nonoptional section of this book depends on material
in any optional section, with the exception of a few exercises in which
the dependence is clearly indicated. Where an optional section depends on
Preface XIX
Acknowledgments
There are many who contributed valuable suggestions and various forms
of aid and encouragement to this project. Among those, I would particu-
larly like to mention Sheldon Axler, Mahlon Day, Alphonso DiPietro, Fred
Gehring, John LeDuc, Don Lewis, Tenney Peck, M. S. Ramanujan, Ira
Rosenholtz, Mark Smith, B. A. Taylor, and Jerry Uhl. I would also like
to express my appreciation for the support and editorial assistance I re-
ceived from Springer-Verlag, particularly from Ina Lindemann and Steve
Pisano. This book would hardly have been possible without the day-to-
day support of my wife Kathy, who did a wonderful job of insulating me
from the outside world during its writing and who did not hesitate to pass
stern judgment on some of my more convoluted prose. Finally, my special
thanks go to my friend and mentor Horacio Porta, who read an early ver-
sion of the manuscript and made some extremely valuable suggestions that
substantially improved the content and format of this book.
This chapter contains the basic definitions and initial results needed for a
study of Banach spaces. In particular, the material presented in the first
twelve sections of this chapter, with the exception of that of Section 1.5,
is used extensively throughout the rest of this book. Section 1.13, though
containing material that is very important in modern Banach space theory,
is optional in the sense that the few results and exercises in the rest of the
book that depend on this material are clearly marked as such.
1.1 Preliminaries
Here are some of the definitions, conventions, and notation used throughout
this book. Whenever a definition contains two or more different names for
the same object, the first is the one usually used here. The alternative names
are included because they are sometimes encountered in other sources.
The set of positive integers is denoted by N. The fields of real and complex
numbers are denoted by IR and C respectively. The symbol IF denotes a field
that can be either IR or C. The elements of IF are called scalars.
In a topological space, the closure of a set A, denoted by ::4, is the smallest
closed set that includes A, that is, the intersection of all closed sets that
include A. The interior of A, denoted by A 0, is the largest open subset
of A, that is, the union of all open subsets of A.
A vector space or linear space over IF is a set X of objects called vectors
along with an operation + from X x X into X called addition of vectors
2 1. Basic Concepts
1 In this bO'O'k, finite sets as well as cO'untably infinite O'nes are said to' be cO'untable.
1.1 Preliminaries 3
x+A = {x + y: YEA},
x- A= {x - y : YEA},
A+B = {y + z: YEA, z E B},
A- B = { Y - z : YEA, z E B },
aA = { ay : YEA}, and
-A = {-y: YEA}.
The set x + A is called the translate of A by x. Notice that A - B represents
the algebraic difference of the sets A and B. The set-theoretic difference
{ x : x E A, x f/- B} is denoted by A \ B.
Suppose that A is a subset of a vector space X. Then A is convex if
ty + (1 - t)z E A whenever y, z E A and 0 < t < 1. If aA ~ A whenever
lal ~ 1, then A is balanced. The set A is absorbing if, for each x in X,
there is a positive number Sx such that x E tA whenever t > Sx' The
following properties of these special types of sets are not difficult to prove;
see Exercises 1.1 and 1.3.
(1) Absorbing sets always contain O. So do nonempty balanced sets.
(2) If A is a balanced set, then aA = A whenever lal = 1, which in
particular implies that -A = A.
(3) Arbitrary unions and intersections of balanced sets are balanced.
(4) Arbitrary intersections of convex sets are convex.
(5) Translates and scalar multiples of convex sets are convex.
(6) The set A is convex if and only if sA+tA = (s+t)A whenever s, t > O.
The convex hull or convex span of A, denoted by co(A), is the smallest
convex set that includes A, that is, the intersection of all convex sets that
include A. It is not difficult to show that co(A) is the collection of all con-
vex combinations of clements of A, that is, all sums of the form 2::7=1 tJx J
such that n E N, Xl, ... , Xn E A, t 1 , ... , tn 2: 0, and 2::;'=1 tj = 1. See Ex-
ercise lA. If X has a topology, then the closed convex hull or closed convex
span of A, denoted by co(A), is the smallest closed convex set that in-
cludes A, that is, the intersection of all closed convex sets that include A.
The linear hull or linear span of A, denoted by (A), is the smallest subspace
of X that includes A, that is, the intersection of all subspaces of X that in-
clude A. Notice that (0) = {O}. If A is nonempty, then it is not difficult to
4 1. Basic Concepts
20f course, the composite go f is defined for any functions f and g, linear or not,
such that the range of f lies in the domain of g, but in the nonlinear case it is customary
to insert the 0 symbol and not to use the term product, especially in situations in which
composites could be confused with pointwise products of scalar-valued functions.
1.1 Preliminaries 5
Zorn's Lemma. (M. Zorn, 1935 [249]). A preordered set in which each
chain has an upper bound contains at least one maximal element.
Exercises
Unless stated otherwise, in each of these exercises X is a vector space, the ele-
ments x and yare vectors in X, the sets A, B, and C are subsets of X, and 0:
and (3 are scalars.
3The symbol Qt is the Fraktur letter A. Uppercase Fraktur letters Qt, '23, ([, ... will
often be used to denote certain types of sets, particularly those whose members are sets
or functions. See the List of Symbols for the list of Fraktur letters corresponding to the
uppercase Roman letters.
1.1 Preliminaries 7
1.1 (a) Show that if A is an absorbing set or a nonempty balanced set, then
o E A.
(b) Show that if A is balanced, then aA =A whenever lal = 1.
(c) Suppose that Q3 is a collection of balanced subsets of X. Show that
u{ s :
S E Q3 } and n{
S : S E Q3 } are both balanced.
(d) Suppose that <!: is a collection of convex subsets of X. Show that
n{S : S E <!:} is convex.
(e) Show that if A is convex, then x +A and aA are convex.
1.2 (a) Show that the "addition" and "multiplication by scalars" defined for
sets obey the commutative and associative laws for vector spaces.
That is, show that A+ B = B +A, that A+ (B +C) = (A+B) +C,
and that a(j3A) = (aj3)A. Show also that (x + A) + (y + B)
(x+y)+(A+B).
(b) Show that alA + B) = aA + aBo
(c) Show that (a + j3)A ~ aA + j3A but that equality need not hold.
1.3 (a) Prove that A is convex if and only if sA + tA = (s + t)A for all
positive sand t. (Consider the special case in which s + t = 1.)
(b) Use (a) and Exercise 1.2 to prove that if A and B are convex, then
so is A + B.
1.4 Show that the convex hull of A is the collection of all convex combinations
of elements of A. Show that if A is nonempty, then the linear hull of A is
the collection of all linear combinations of elements of A.
1.5 Prove that if A and B are balanced, then so is A + B.
1.6 Suppose that A is balanced. Prove that A is absorbing if and only if the
following holds: For each x in X there is a positive number tx such that
x E txA.
1. 7 Identify all of the balanced subsets of IC. Do the same for ]R2.
1.8 The balanced hull bal(A) of A is the smallest balanced subset of X that
includes A. There is a simple expression for bal(A) as the union of a
certain collection of sets. Find it.
1.9 Prove that each convex absorbing subset of C includes a neighborhood
of o.
1.10 Is the conclusion of the preceding exercise true for nonconvex absorbing
subsets of C?
1.11 Let X and Y be vector spaces. Suppose that A ~ X, that B ~ Y, and
that T: X -> Y is a linear operator.
(a) Prove that if B is convex, or balanced, or absorbing, or a subspace,
then T- 1 (B) has that same property.
(b) Prove that if A is convex, or balanced, or a subspace, then T(A)
has that same property. Prove that if T maps X onto Y and A is
absorbing, then T(A) is absorbing.
8 1. Basic Concepts
1.12 Let X be the vector space JR2 with the topology whose only member
besides the entire space and the empty set is {(a,f3) : (3 > a 2 }. Find a
subspace A of X such that A is not convex (and therefore is not a subspace
of X), so that neither of the equations [A] = (A) and co(A) = co(A) holds.
(It will be shown in Sections 1.3 and 2.2 that both equations do hold for
every subset A of a topologized vector space under certain restrictions on
the topology, in particular when it comes from a norm.)
1.13 Let M be a metric space with metric d. Suppose that A is a nonempty
subset of M and that x and yare elements of M. Show that d(x, A) :::;
d(x, y) + d(y, A). Conclude that Id(x, A) - d(y, A)I :::; d(x, y).
1.2 Norms
Suppose that X is a vector space. For each ordered pair (Xl, X2) of elements
of X, define (Xl, X2] to be {y : Y = (l-t)XI +tX2, 0 < t :::; 1 }. In particular,
this implies that (x, x] = {x} for each x in X. It is not difficult to check
that if (Xl, X2] = (Yl, Y2], then Xl = Yl and X2 = Y2; see Exercise 1.18.
Define an equivalence relation on the collection of all of these "half-open line
segments" in X by declaring that (Xl, X2] is equivalent to (Yl, Y2] whenever
there is a z in X such that Yl = Xl + z and Y2 = .T2 + z. That is, two of
1.2 Norms 9
these segments are equivalent when one is a translate of the other. Let the
arrow ~ with head X2 and tail Xl be the equivalence class containing
(Xl, X2]' It is easy to see that the collection Xa of all such arrows is a vector
space over IF when given these operations:
----> )
a Xl,X2 = aXl,aX2
It is also easy to see that the map X 1--* 0,1 is a vector space isomorphism
from X onto Xa. Thus, the vectors of X can be regarded either as its
elements or as directed line segments in X that remain fundamentally un-
changed when translated. In fact, the familiar "arrow vectors" of calculus
are obtained by treating the vectors of ]R2 or ]R3 the second way.
The fundamental metric notion for vectors can be either distance or
length, depending on the way vectors are treated. If they are thought of as
points, then the distance between two vectors is a reasonable notion, while
the length of a vector would seem to be a concept best left undefined. On the
other hand, the length of a vector is an entirely natural concept if vectors
are considered to be arrows. The distance between two vectors would then
seem to be a more nebulous notion, though it might be reasonable to obtain
it, as is done with the arrow vectors of calculus, by joining the tails of the
two vectors and then measuring the length of one of the two "difference
vectors" between their heads.
The concept of a norm comes from thinking of vectors as arrows. A norm
on a vector space is a function that assigns to each vector a length. There
are some obvious properties that such a function should be required to
have. A nonzero vector should have positive length; the additive inverse of
a vector should have the same length as the original vector; half a vector
should have half the length of the full vector; a vector forming one side of
a triangle should not be longer than the sums of the lengths of the vectors
forming the other two sides; and so forth. These requirements are embodied
in the following definition.
The term norm is commonly used for both the function of the preceding
definition and its values, so that Ilxll is read as "the norm of x." Another
10 1. Basic Concepts
As will be seen in the next chapter, there are other natural topologies that
normed spaces possess besides their norm topologies. Henceforth, when-
ever reference is made to some topological property such as compactness or
convergence in a normed space without specifying the topology, the norm
topology is implied.
Suppose that X is a normed space and that B(x, r) is the closed ball
centered at a point x of X and having radius r. It follows easily from the
defini tion of the metric of X that B (x, r) = x + r B( 0, 1). A corresponding
relationship exists hetween open balls in X and the open ball with radius 1
centered at the origin. Because of this, and for other reasons that will
become apparent later in this book, the balls of radius 1 centered at the
origin playa special role in the theory of normed spaces and are therefore
given special names.
1.2.4 Example: IF. The prototype for all norms is the absolute value
function on IF. It is the norm implied whenever IF is treated as a normed
space without the norm being specified. Notice that this norm induces the
standard metric on IF.
1.2.6 Example: Lp(0., E, /1-), 1 :::; p :::; 00. Let /1- be a positive (that is,
nonnegative-extended-real-valued) measure on a cr-algebra E of subsets of
a set 0.. For each p such that 1 :::; P :::; 00, the Lebesgue space Lp(0., E, /1-)
is a normed space with the norm II lip given by letting
Ilfllp = {(llfI P
d/1-) lip if 1:::; p < 00;
inf{t:t>O,/1-({X:XE0., If(x)1 >t})=O} if p = 00.
For each p, the elements of Lp(0., E, /1-) are equivalence classes of either real-
or complex-valued functions, giving, respectively, real or complex normed
spaces. In the rare instances in which it is necessary to be specific about
the scalar field, the terms "real Lp(0., E, J.L)" and "complex Lp(fl., E, J.L)" are
used. A similar convention applies to the spaces defined in the following ex-
amples. For many ofthe Lebesgue spaces appearing in this book, the set fl. is
the interval [0,1], the cr-algebra E is the collection of Lebesgue-measurable
subsets of [0, 1], and the measure /1- is Lebesgue measure on [0,1]' in which
case Lp(0.,~, /1-) is abbreviated to Lp[O,l]. Other abbreviations such as
Lp[O,oo) have analogous meanings.
1.2.8 Example: CP ' 1 :S p < 00. Let p be a real number such that p ?: 1.
The collection of all sequences (aj) of scalars for which L~llaj IP is finite
is a vector space with the vector space operations of the preceding example.
Let the norm II lip be defined on this vector space by the formula
1.2.9 Example: C;,1 :::; p :::; 00. Let p be such that 1 :::; p ::: 00 and let n
be a positive integer. Define a norm on the vector space lF n by letting
The resulting normed space is called R.; (pronounced "little ell p n"). Notice
that C2 is just Euclidean n-space. By convention, the space cg is Euclidean
O-space. It is also possible to represent cg as a Lebesgue space by using
the fact that there is exactly one scalar-valued function on the empty set,
namely, the function represented by 0 x IF; see [65, p. 11, Ex. 3]. It follows
that cg can be defined to be Lp( 0, {0}, J1), where of course J1(0) = O.
1.2 Norms 13
It is easy to check that this defines a norm on the vector space. The re-
sulting normed space is denoted by C(K). By analogy with Lp[O, l], the
abbreviation C[O, 1] represents C([O, 1]).
It turns out that each of the norms in the preceding examples induces
a complete metric. For IF, this is a basic fact from analysis; for C(K),
this is just a special case of the fact that every uniformly Cauchy se-
quence of scalar-valued continuous functions on a topological space con-
verges uniformly to a continuous function; and for all of the other spaces
except rca(K), this follows from the completeness of the metrics of Lebesgue
spaces. The completeness of rca(K) is not difficult to prove from basic
principles, as is done in Exercise 1.28, and also follows from a result in
Section 1.10 about the completeness of dual spaces, as is mentioned in
the comments following Theorem 1.10.7. It is also not difficult to prove the
completeness of all of the spaces f.p and f.;, and therefore of all of the finite-
dimensional Euclidean spaces, without invoking the general fact that every
Lebesgue space is a complete metric space. See the proof of Theorem C.10
in Appendix C.
It is finally time for the definition of the objects that are the main focus
of this book.
1.2.14 Example: Lp(ll') and Hp, 1 ~ p ~ 00. Fix p such that 1 ~ p ~ 00.
Let 11' be the unit circle {z : z E C, Izl = I} in the complex plane, and let X
be the set of all complex-valued functions f on 11' with the property that
if g: [-Jr,Jr) -+ C is defined by letting g(t) = f(e it ), then 9 E Lp[-Jr,Jr).
As would be expected, two functions hand 12 in X are considered to be
the same if h (e it ) = 12 (e it ) for almost all t in [-Jr, Jr). Since Lp [-Jr, Jr) is
a Banach space, it is clear that X is also a Banach space with the obvious
vector space operations and the norm 1Illp given by letting
Ilfllp = {( ~ 1"
2Jr_"
If(eit)IP dt) lip
if p
~
=
P<
00.
00;
This Banach space is called Lp(ll'). Notice that Lp(ll') is essentially just
Lp[-Jr,Jr), except that [-Jr,Jr) has been identified with 1I' and Lebesgue
measure A has been replaced by normalized Lebesgue measure (2Jr)-1 A so
that the measure of 1I' is 1.
Suppose that f E Lp(1I'). For each integer n, the nth Fourier coeffi-
cient j( n) of f is defined by the formula
j(n) = ~
2Jr
lIT
-1\"
f(eit)e-int dt.
The following vector space, with various norms, is useful for generating
counterexamples.
Exercises
1.18 With all notation as in the first paragraph of this section, show that
if (Xl, X2] = (YI, Y2J, then Xl = YI and X2 = Y2. (This is easy when
Xl = X2 Suppose instead that Xl -=1= X2. It may help to show first that
[XI,X2] = [YI,Y2J, where [XI,X2] = {y: Y = (1- t)XI +tX2, 0 ~ t ~ I}.
To this end, notice that (XI,X2] and [Xl, X2] are both convex and that
[Xl, X2] has exactly one more point than does (Xl, X2]. In how many ways
can (Xl, X2] be augmented by one point so that the resulting set is convex?)
1.19 The definition of a norm contains some redundancies. Prove that an equiv-
alent definition is obtained by replacing (1) in that definition with
(1') IIxll -=1= 0 whenever X -=1= o.
1.20 A sphere in a metric space X is a set of the form {y : Y E X, d(x, y) = r},
where X E X and r > o.
(a) Prove that if X is a normed space, then every closed ball is the
closure of the corresponding open ball, every open ball is the interior
of the corresponding closed ball, and every sphere is the boundary
of the corresponding open and closed balls.
16 1. Basic Concepts
(b) Find a metric space X such that the three etmclusions of part (a)
all fail to hold for X.
1.21 Let X be a normed space. Prove that if (Bn) is a sequence of balls in X
such that Bn 2 Bn+1 for each n, then the centers of the balls form a
Cauchy sequence. Give an example to show that this result can fail if X
is only assumed to be a metric space.
1.22 The purpose of this exercise is to show that the Banach spaces of Examples
1.2.7, 1.2.8, and 1.2.9 can be obtained by subspace arguments beginning
with the Banach spaces Lp[O, 00).
(a) Show that 00 can be identified with a closed subspace of Loo[O, 00)
by identifying (OJ) with the function whose value on each interval
[j - 1, j) is OJ. Conclude that 00 is a Banach space.
(b) Use a similar argument to prove that p is a Banach space when
1:::; p < 00.
(c) For each p such that 1 :::; p :::; 00 and each nonnegative integer n,
identify ; with a closed subspace of p and conclude that e; is a
Banach space.
(d) Give analogous arguments for (a) and (b) based on the spaces Lp[O, 1]
instead of Lp [0, 00 ).
1.23 Let X be the vector space of all continuous functions from [O,IJ into IF.
For each p such that 1 :::; p :::; 00, let II lip be the norm that X inherits
from Lp[O, 1]. For which values of p is this a Banach norm?
1.24 Let X be the vector space of finitely nonzero sequences. Show that X is
a dense subspace of Co and of p when 1 :::; p < 00, but not of 00.
1.25 Let c be the collection of all convergent sequences of scalars with the vector
space operations and norm as given for 00. Show that c is a Banach space.
1.26 (a) In each of the spaces Co and p such that 1 :::; p :::; 00, identify the
linear hull ({ en : n EN}) of the collection of unit vectors.
(b) Do the same for [{ en : n E N }J.
(c) Do the same for co({ en: n EN}).
1.27 (a) Suppose that the unit sphere of 1 contains the sequence (x(n)
and the element x. Show that (x(n) converges to x if and only
if limn x;nl
= Xj for each positive integer j.
(b) Show that (a) fails if the requirement that IIxll = 1 is removed.
(c) Show that (a) fails if 1 is replaced by Co.
1.28 Suppose that E is a a-algebra of subsets of a set n. Let ca(n, E) be
the normed space formed from the collection of all finite scalar-valued
measures on E by using the obvious vector space operations and by letting
the norm of a measure be its total variation. It is not difficult to show that
ca(n, E) really is a normed space; see [202, pp. 116-119J for the details.
(a) Prove that ca(!!, E) is a Banach space.
1.3 First Properties of Normed Spaces 17
(b) Now suppose that i.8 is the a-algebra of Borel subsets of a com-
pact Hausdorff space K. Prove that rca(K) is a closed subspace of
ca(K, i.8) and is therefore a Banach space.
1.29 Fix a positive real number p. A scalar-valued function f on [0,1] is said
to satisfy a Lipschitz condition of order p if there is a nonnegative real
number M such that If(s) - f(t)1 ~ Mis - W whenever s, t E [0,1]. The
collection of all such functions is denoted by Lip p.
(a) Show that Lipp is a subspace of the vector space of continuous
scalar-valued functions on [0,1]. Show that if p > 1, then Lipp
consists only of the constant scalar-valued functions on [0, 1].
(b) Define a function II II: Lip p -+ IR by the formula
I/(s) - f(t)1 }
Ilfll = If(O)1 + sup { Is _ W : s, t E [0,1]' s =I- t .
.
hm sup
{ If(s)
I - f(t)1
IP : s, t E [0,1], s =I- t, Is - tl ~ 8
} = O.
6~O+ S - t
Prove that lip p is a Banach space under the norm inherited from
Lipp.
and
1.3.6 Example. Suppose that X is Cp , where 1 :::; p < 00, or co. Let (en)
be the sequence of standard unit vectors in X. It is easy to check that
1.3 First Properties of Normed Spaces 19
(an) = En ane n whenever (an) E X. This result does not extend to too,
for if (an) is a member of too whose terms do not tend to 0, then the
sequence (E:=l anen):=l is not Cauchy and therefore cannot converge.
as m -> 00, giving (b) and (c). A similar argument, with a replaced by T
and references to Yn omitted, proves (d). For (e), repeated applications of
the triangle inequality show that 112::=1 xnll ~ 2::=lllxnll for each finite
sum 2:::'=1 xn in X. If 2:n Xn is an infinite sum in X, then the continuity
of the map x ~ Ilxll assures that 112:::'=lxnll-> l12:nxnll as m -> 00, so
the desired inequality for infinite sums can be obtained by letting m tend
to infinity in the inequality 112::: 1 Xn I ~ 2:::'=lllxnII.
Of course, the converse of (a) is not in general true since the harmonic
series diverges. It should also be noted that the sum 2:nllxnll in (e) does
not have to be finite when 2:n Xn is a convergent infinite sum; otherwise,
20 1. Basic Concepts
from which it follows that the partial sums of Ln Xn form a Cauchy se-
quence and therefore that Ln Xn converges.
See Exercise 1.31 for some further observations about absolute and un-
conditional convergence of series in Banach spaces. The reader interested
in unconditional convergence should also see Propositions 4.2.1 and 4.2.3.
While the proofs of those two propositions rely only on material that has
already been covered, the results have been postponed until Section 4.2
since they are not needed before then.
1.3 First Properties of Normed Spaces 21
It is not in general true that balls in a vector space with a metric must
be convex; see Exercise 1.35. However, balls in a normed space always have
this property.
Balls centered at the origin of a normed space have some additional useful
features.
In the preceding proof, it is not really necessary to force the radii of the
balls to decrease to O. The fact that Bn :2 Bm whenever n ::; m is enough
to assure that their centers form a Cauchy sequence. See Exercise 1.21.
Exercises
1.30 Let K be a compact Hausdorff space and let X be a normed space. By
Corollary 1.3.4, the collection of all continuous functions from K into X
is a vector space when functions are added and multiplied by scalars in
the usual way. Define a norm on this vector space by the formula
1.33 This exercise shows that the order in which closures and hulls are taken
in Corollary 1.3.11 is important.
(a) Prove that if A is a subset of a normed space, then co(A) :;! co(A)
and [A] :;! (A).
(b) Show that if A is the collection of unit vectors in co, then the inclu-
sions in (a) are proper.
1.34 Suppose that 5 is a subspace of a normed space X. It is clear that S
need not be a subspace of X because S might be empty; consider {ot
in IF. It is true, however, that S is a subspace of X when S i 0. In
fact, a much stronger statement can then be made about S. What is that
statement?
1.35 Euclidean-space intuition might make it tempting to think that balls in
a vector space with a metric must be convex. To see that this is not so,
let X be the vector space ]R2 with the metric given by the formula
Show that d really is a metric on ]R2. Let B x be the closed unit ball
of X, that is, the set of all members of X no more than 1 unit from (0,0).
Sketch Bx. Notice that (1,0) and (0,1) are in Bx, but d,~) is not. Thus,
the ball B x is not convex.
24 1. Basic Concepts
1.36 Suppose that X is the vector space underlying 1, but equipped with
the 00 norm. Show that {(an) : (an) E X, Lnlanl :::; I} is a closed,
convex, absorbing subset of X whose interior is empty. (Notice that, by
Theorem 1.3.14, the normed space X cannot be a Banach space.)
1.37 Suppose that a vector space X has a nonempty subset B that is convex,
balanced, and has this strong absorbing property: For every nonzero x
in X, there is a positive Sx such that x E tB if t ~ Sx and x (j. tB if
o :::; t < sx. Show that there is a norm II liB on X for which B is the
closed unit ball.
1.38 Suppose that T is a linear operator from a Banach space X into a normed
space Y. Show that if T- 1 (By) is closed, then T is continuous at O.
(It can then be concluded from Theorem 1.4.2 that T is actually contin-
uous everywhere.)
1.39 The core of a subset A of a vector space X is the collection of all points y
in A with this property: For each x in X, there is a positive OX,y such that
y + tx E A whenever 0 :::; t < Ox,y. Prove that if X is a Banach space,
then the core of each closed convex subset of X is the interior of that set.
A function into a metric space is often called bounded when its range
is a bounded set. It is quite possible for a linear operator between normed
1.4 Linear Operators Between Normed Spaces 25
It follows from the preceding theorem that the collection B(X, Y) of all
bounded linear operators from a normed space X into a normed space Y is
the same as the collection of all continuous linear operators from X into Y.
Convention dictates that a member of B(X, Y) be called bounded rather
than continuous, which also explains why the notation B(X, Y) is used for
this collection rather than something like CL(X, Y). While L(X, Y) is used
in this book to denote the vector space of all linear operators from X into Y,
be warned that many authors use L(X, Y) to denote the corresponding
space of bounded linear operators.
Let V be the vector space of all continuous functions from a normed
space X into a normed space Y with addition of functions and multiplica-
tion of functions by scalars defined in the usual way; see Corollary 1.3.4.
Then B(X, Y) is clearly a subspace of V, and so is itself a vector space. It
would be nice if B(X, Y) could somehow be made into a normed space. To
do that, some notion for the "length" or "size" of a member T of B(X, Y)
is needed. The quantity sup{ IITxl1 : x E Bx} mentioned in part (g) of
Theorem 1.4.2 is in a sense a measure of the size of T(Bx), and therefore
would seem to be a plausible candidate.
1.4.3 Definition. Let X and Y be normed spaces. For each T in B(X, Y),
the norm or opemtor norm IITII of T is the nonnegative real number
sup{ IITxl1 : x E Bx }. The opemtor norm on B(X, Y) is the map T I---> IITII.
1.4.4 Example. Let X and Y be normed spaces. The zero opemtor from X
into Y is the zero element of the vector space B(X, Y), that is, the operator
that maps each x in X to the zero element of Y. This operator clearly has
norm O. The identity opemtor on X is the member I of B(X) defined by the
formula Ix = x. Notice that 11111 = 1 as long as X # {O}. More generally, for
each scalar a the linear operator Ta on X given by the formula TaX = ax
is bounded, and has norm lal if X # {O}. Such operators are called scalar
opemtors.
Here are some useful characterizations of the operator norm that are
sometimes used for its definition.
1.4 Linear Operators Between Normed Spaces 27
1.4.7 Proposition. Suppose that X and Y are normed spaces and that
T is a bounded linear operator from X into Y.
(a) IITII = sup{ IITxl1 : x E X, Ilxll < 1 }.
(b) If X i- {O}, then IITII = sup{ IITxl1 : x E Sx }.
(c) If x E X, then IITxl1 ~ IITIIllxll. Furthermore,
the number IITII is
the smallest nonnegative real number M such that IITxl1 ~ Mllxll
for each x in X.
PROOF. Define I: X -lR. by the formula I(x) = IITxll. Since I is continu-
ous, its supremum on the open unit ball of X is the same as its supremum
on the closure Bx of that ball, which gives (a). If x is a nonzero element
of B x , then IIxll-Ix E Sx and 1(llxll-Ix) = IIxll-I/(x) ;::: I(x). It follows
that the supremum of I on S x is the same as its supremum on B x when-
ever Xi- {O}, which gives (b). For (c), first notice that if 0 ::; M < IITII,
then there is an x in Bx such that IITxl1 > M ;::: Mllxll, so there is no
nonnegative real number M smaller than IITII such that IITxll ::; Mllxll for
each x in X. If x E X and x i- 0, then IIxll-IIlTxll = IIT(llxll-Ix) II ::;
IITII,
and so IITxll ~ IITllllxll. This last inequality is trivially true if x = 0, which
finishes the proof of (c).
So far, it has not been shown that the operator norm really is a norm.
It is time to remedy this oversight.
as n ---> 00.
The converse of Proposition 1.4.9 is in general false, as should be ex-
pected. Convergence in B(X, Y) represents uniform convergence on the
closed unit ball of X, so it is not surprising that such convergence is not
always implied by pointwise convergence on X. See Exercise 1.44.
Suppose that X, Y, and Z are normed spaces and that S E B(X, Y) and
T E B(Y, Z). Then the product TS is in B(X, Z) since the composite of
continuous functions is continuous. It is too much to hope that liT SII would
have to equal IITIIIISII; for example, it is easy to find nonzero members S
and T of B(JR2 ) whose product is the zero operator. The following, at least,
can be said.
1.4 Linear Operators Between Normed Spaces 29
This inequality also shows that IITSII ~ IITIlIiSII, since IITSII is the smallest
nonnegative real number M such that IITS(x)1I ~ Mllxll for each x in X .
As was shown in Example 1.4.5, there are normed spaces X and Y such
that some members of L(X, Y) are unbounded. Of course, this cannot hap-
pen when Y = {O}, for then L(X, Y) contains only the zero operator. This
observation, together with the following two theorems, completely settles
the question of which ordered pairs (X, Y) of normed spaces possess the
property that L(X, Y) has an unbounded member: this happens if and only
if X is infinite-dimensional and Y =I- {O}.
and let W be the normed space (V, I I). It follows immediately from the
definition of I I that a sequence (a~jlxl + ... + a}!lx n ) in W converges to
some member alxl + ... + anXn of W if and only if limj ag l = am when
m= l, ... ,n.
30 1. Basic Concepts
and that O:IXl + ... + O:nXn E Sw. Therefore every sequence in Sw has
a subsequence converging to a member of Sw, so Sw is compact. Since
I-I E L(W, X}, the function W t-+ III-lwll from W into IR is continuous
and so attains a positive minimum on the compact set Sw. If I were not
bounded, then there would be a sequence (Zj) in Bx such that /Iz j / ~ j
for each j. Letting Wj = /Izj/-lIzj for each j yields a sequence (Wj) in Sw
such that III-lwjll = /Izj/-lllzjll --> 0 as j --> 00, a contradiction. Thus,
the operator I must be bounded.
The significance of part (c) of the preceding proposition comes from the
fact that, in general, homeomorphisms between metric spaces do not have
to preserve completeness. See Exercise 1.42.
An easy application of Theorem 1.4.12 immediately produces a large
supply of isomorphic normed spaces.
PROOF. Let T be a linear operator from X onto Y. Since X and Y have the
same finite dimension, the operator T is one-to-one. By Theorem 1.4.12,
both T and T- 1 are bounded.
1.4.18 Corollary. For each nonnegative integer 71" the only norm topology
that IF" can have is its Euclidean topology.
bounded subset of IFn and ((a~j), ... , a~))) is a sequence in A, then the
bounded ness of each of the sequences (a~)) such that m = 1, ... , n as-
sures the existence of a subsequence (jk) of N and scalars al,"" an such
that (a~.)) --+ am for each m, which implies that (a~jkl, ... , a:~k)) --+
(aI, ... ,an)' It is trivially true that 1F0 has the Heine-Borel property.
Let X be a uormed space having finite dimension n and let T be an
isomorphism from X onto Euclidean n-space. If S is a closed bounded
subset of X, then T(S) is closed and bounded in IF n and therefore compact,
so S is itself compact.
A little more work shows that the Heine-Borel property actually charac-
terizes the finite-dimensional normed spaces among all normed spaces.
Exercises
1.40 Supply the details in Example l.4.6.
34 1. Basic Concepts
1.41 Suppose that Y is the vector space of finitely nonzero sequences, equipped
with the 1 norm. Show that B(l, Y) is not a Banach space.
1.42 Define d: RxR --> R by the formula d(x, y) = Itan- 1 (x)-tan- 1 (y)!. where
tan- 1 denotes the usual single-valued inverse tangent function from R
onto (-7r /2, 7r /2). Prove that d is an incomplete metric on R that induces
the usual topology of R. Conclude that it is possible for an incomplete
metric space to be homeomorphic to a complete one.
1.43 Suppose that T is a linear operator from a normed space X into a normed
space Y such that En TX n is a convergent series in Y whenever En Xn
is an absolutely convergent series in X. Prove that T is bounded.
1.44 For each positive integer j, define ej : Co --> IF by the formula ej(a n ) = aj.
Show that (ej) is a sequence in B(eo,lF) that does not converge to the
zero operator 0 of B(co,lF), even though limj ej(a n ) = O(an) whenever
(an) E Co.
1.45 Prove that no Banach space has a count ably infinite vector space basis.
(Suppose that X is a normed space with a countably infinite vector space
basis (b n ). For each positive integer n, let Yn = (b 1 , . , bn ). Select an Xl
in Yl such that Ilxlll = 1. Use an argument like that of Lemma 1.4.22 to
select an X2 in Y 2 such that d(X2, Yd = IIx2 - Xlii = 1/4. Select an X3
in Y3 such that d(X3, Y2) = IIx3 - X21! = 1/16. Continuing in this vein
yields a Cauchy sequence that cannot converge to anything in any Y n .)
1.46 Suppose that Y is a finite-dimensional normed space. Let T be a linear
operator from a normed space X onto Y, where X is not assumed to be
finite-dimensional and T is not assumed to be bounded. Prove that T is
an open mapping, that is, that T(U) is an open subset of Y whenever U
is an open subset of X. (Notice that it is enough to prove that T(Bx)
includes a neighborhood of 0.)
1.47 Let f be an unbounded linear functional on a normed space X. Prove
that if U is a nonempty open subset of X, then feU) = IF.
1.48 Let X and Y be normed spaces and let f: X --> Y be additive; that is,
let f(XI + X2) = f(xd + f(X2) whenever Xl,X2 EX.
(a) Prove that f(rx) = rf(x) for each rational number r and each X
in X.
(b) Prove that if X and Yare real normed spaces and f is continuous,
then f is linear.
(c) Show that the conclusion of (b) might not hold if X and Y are
complex normed spaces and f is continuous.
1.49 Let c be the Banach space of convergent sequences of scalars defined in
Exercise 1.25.
(a) Prove that c is isomorphic to co.
(b) Suppose that x E Seo ' Prove that there exist Xl and X2 in SeQ such
that Xl # X2 and X = ~(Xl + X2).
(c) Conclude from (b) that c is not isometrically isomorphic to eo.
1.5 Baire Category 35
1.51 Here is another proof of Corollary 1.4.18 that has a more geometric flavor.
Let n be a nonnegative integer. Let II I/o be a norm on lF n , let'! be the
topology induced by this norm, and let Uo and Uc be the open unit balls
for 1/. I/o and the Euclidean norm respectively. In the following, any term
preceded by "e-" refers to the Euclidean topology or norm. Do not use
Theorem 1.4.12 or any results based on it in your arguments. You may use
the fact that Euclidean n-space has the Heine-Borel property, since the
demonstration of that fact in the proof of Corollary l.4.21 is elementary.
(a) Using the fact that e-convergence implies coordinatewise conver-
gence when n ::::: 1, show that each e-convergent sequence in lF n is
'I-convergent to the e-limit. Conclude that every 'I-open subset oflF n
is e-open.
(b) Let C be an e-closed e-unbounded convex set in lF n that contains
the origin. Show that C includes a ray emanating from the origin,
that is, a set of the form {tx : t ::::: O} where x is a nonzero element
of lFn.
(c) Let C be an e-open e-unbounded convex set in lF n that contains the
origin. Show that C includes a ray emanating from the origin.
(d) Show that sUo ~ U e for some positive s. Conclude that every e-open
subset of lF n is 'I-open, and that'! must therefore be the Euclidean
topology. This concludes the proof of Corollary 1.4.18.
(e) Obtain Corollary l.4.16 from Corollary l.4.18. (Let T be a linear
operator from an n-dimensional normed space X onto lF n and define
a norm 1/. I/o on lF n by letting lIyl/o = I/T-1yl/. Use this norm to show
that X is isomorphic to Euclidean n-space.)
Baire category should spend some time with this section, since Baire cate-
gory is an important component of the working analyst's toolkit.
It is often useful to be able to prove that a set is too large to be written
as a union of countably many other sets that are in some sense very small.
For example, facts about the real line that depend on its uncountability
are true because the real line is "large" enough that it cannot be written
as a union of countably many one-point "very small" sets. Similarly, many
results in measure theory are obtained from knowing that a set of nonzero
measure is too "large" to be a union of countably many "very small" sets
of measure zero.
The concept of Baire category, due to Louis Rene Baire, gives a topolog-
ical meaning to the notion of the size of a set. Baire's approach is based
on density. A subset A of a topological space X is considered to be very
small in Baire's sense if there is no nonempty open subset U of X such that
An U is dense in U, that is, if it has empty interior. Baire's large sets are
those that are not unions of countably many of these very small sets.
1.5.2 Example. If a is any real number, then the set {a} is obviously of the
first category, in fact nowhere dense, in the real line, and is equally obviously
of the second category in itself. Since the set Q of rational numbers can be
written as the union of countably many singleton subsets of JR., it follows
that Q is of the first category in JR. while being at the same time dense in 1Ft.
Here are some simple but useful facts about Baire category, followed by
the main result of this section.
(d) The union of a countable collection of sets each of the first category
in X is of the first category in X.
(e) If a subset of X is of the second category in X, then it is of the
second category in itself.
PROOF. Clearly, subsets of nowhere dense subsets of X are themselves
nowhere dense in X. It follows that if A is the union of a countable collec-
tion It of nowhere dense subsets of X and B ~ A, then B is the union of
the countable collection {B n C : CElt} of nowhere dense subsets of X,
which proves (a).
Now suppose that A is an arbitrary subset of X. Then (At = 0 if and
only if X \ A intersects every nonempty open subset of X, which happens
if and only if X \ A is dense in X. Since X \ A = (X \ A)O, this proves (b).
Part (c) then follows from a straightforward application of De Morgan's
laws.
Part (d) is obvious. For (e), notice that if A ~ X, then a subset of A
that is nowhere dense in A is nowhere dense in X, from which it follows
that if A is of the first category in itself, then it is also of the first category
inX.
1.5.4 The Baire Category Theorem. (L. R. Baire, 1899 [9]). Let X be
a nonempty complete metric space.
(a) If it is a countable collection of open subsets of X each of which is
dense in X, then n{ U : U E it} is dense in X.
(b) Every nonempty open subset of X is of the second category in X
and hence in itself. In particular, the entire space X is of the second
category in itself.
PROOF. For (a), suppose that it is a countable collection of dense open
subsets of X. It can be assumed that it =I- 0. Then the members of it can
be written as a sequence (Un), which might require that some member of it
be repeated infinitely often in the list. Let U be a nonempty open subset
of X. To prove (a), it is enough to show that Un (nn Un) =I- 0. Since U1 is
dense in X, there is a closed ball Bl of radius no more than 1 included in
Un U1 Now suppose that m 2 2 and that closed balls B 1 , . .. ,Bm - 1 have
been found such that these conditions are satisfied when 1 ::; j ::; m - 1:
(1) B j - 1 2 B j if j ;::: 2;
(2) B j ~ un U1 n .. n Uj
;
satisfied for each positive integer j. The centers of these closed balls form a
Cauchy sequence that converges to some x in X. This x must lie in each Bn
and therefore in U and each Un, which shows that Un (nn Un) -I- 0 and
proves (a).
For (b), notice that the complement of a nonempty open subset of X
obviously cannot be dense in X, and therefore, by (a), cannot be the in-
tersection of a countable collection of sets each of whose interiors is dense
in X. It follows from Proposition 1.5.3 (c) that every nonempty open subset
of X is of the second category in X, and therefore of the second category
in itself by Proposition 1.5.3 (e).
Of course, the metric space formed from the empty set with its only
possible metric is a complete metric space of the first category in itself.
Nonempty incomplete metric spaces can be of either the first or second
category in themselves. The rationals with the metric inherited from the
reals form an incomplete metric space of the first category in itself, while
the interval (0,1) in R with the metric inherited from R is an incomplete
metric space that is of the second category in itself by part (b) of the Baire
category theorem.
Suppose that P is a property defined for the elements of some nonempty
complete metric space X. It follows from the Baire category theorem that
one way to prove that some member of X has property P is to show that the
members of X lacking property P form a set of the first category in X. If
this can be done, then the collection of elements of X having property P is
a set of the second category in X whose complement is of the first category
in X, so it can even be said that in some sense the "typical" member of X
has property P. The following is a classical application of this idea.
From Newton's time through the early part of the nineteenth century,
most mathematicians assumed that a continuous real-valued function de-
fined on an interval in the real line must be differentiable over most of
its domain. In 1834, Bernhard Bolzano gave an example of a real-valued
function continuous on an interval though differentiable nowhere on that
interval, 4 but for almost a century afterward mathematicians treated such
functions as pathological. However, in 1931 Stefan Banach showed that,
in a sense, the vast majority of continuous scalar-valued functions whose
domain is a given interval in IR are not differentiable anywhere.
1.5.5 Theorem. (S. Banach, 1931 [12]). Let D+ be the collection of all
members f of C[O, 1] for which there is a point xf in [0,1) at which f has
a finite right-hand derivative. Then D+ is of the first category in C[O, 1].
4Weierstrass is usually given credit for finding the first everywhere continuous,
nowhere differentiable function, but his example was first presented in lectures in 1861
and in a paper to the Berlin Academy in 1872. See pages 577 and 627 of [38] for a
discussion of Bolzano's priority.
1.5 Baire Category 39
FIGURE 1.1. A member of e[O, 1] with norm and large right-hand derivative.
PROOF. For each positive integer n, let Un be the collection of all mem-
bers f of e[O, 1] such that for each x in [0,1- n- l ],
sup { =
If(Y~ ~(x) I : x < Y < x + ~ } > n.
It will be shown that each Un is a dense open subset of e[O, 1]. To this end,
fix an no in N. Suppose that (fm) is a sequence in e[O, 1] \ Uno converging
to some fo in e[O,I]. To show that Uno is open, it is enough to show
that fo E e[O, 1] \ Uno. For each positive integer m, let Xm be an element
I
of [0,1 - noll such that (JTn(Y) - fm(xm))/(y - xm)1 ::;
no whenever
XTn < Y < XTn + nolo By thinning the sequence (fm) if necessary, it can be
assumed that there is an Xo in [0,1 - noll such that Xm ---+ Xo. If Y is such
that Xo < Y < Xo +nol, then for m large enough that XTn < Y < Xm +nol,
I
and so (Jo(Y) - fo(xo))/(y - xo)1 ::;
no It follows that fo E e[O, 1] \ Uno'
and therefore that Uno is open.
The proof that Uno is dense in e[O, 1] uses the Weierstrass approximation
theorem, which says that the polynomials are dense in e[O, 1]; see, for
example, [19] or [201]. If to and M are positive numbers, then Figure 1.1
shows how to construct a sawtooth function that is in e[O, 1], has norm to,
and has a right-hand derivative with absolute value greater than M at each
point of [0,1). Since each polynomial on [0,1] has a bounded right-hand
derivative on [0,1), it follows that if p is a polynomial on [0,1] and to> 0,
then a sawtooth function can be added to p to yield a member u of Uno
such that Ilu - pll= ::; . The set Uno is therefore dense in e[O, 1].
Thus, each Un is a dense open subset of e[O, 1], and so, by Proposi-
tion 1.5.3 (c), the set e[O, 1] \ (nn Un) is of the first category in e[O, 1].
Since a member of e[O, 1] with a finite right-hand derivative at some point
of [0, 1) cannot lie in every Un, the set of all members of e[O, 1] that are dif-
ferentiable from the right anywhere on [0, 1) is included in e[O, 1] \ ( nn
Un)
and therefore is of the first category in e[O, 1].
40 1. Basic Concepts
Since the Baire category theorem prevents the nonempty complete metric
space C[O, 1] from being the union of two sets of the first category in C[O, 1],
the next corollary follows immediately from the preceding one.
1.5.7 Corollary. The collection of all members of C[O,l] that are not
differentiable anywhere on (0, 1) is of the second category in C[O, 1].
Exercise 1.45 at the end of the preceding section asks for a proof that Ba-
nach spaces never have countably infinite vector space bases. The argument
outlined in that exercise involves the careful construction of a certain non-
convergent Cauchy sequence. The result can also be obtained very easily
from the Baire category theorem.
Exercises
1.52 Derive Theorem 1.3.14 from the Baire category theorem.
1.53 Prove that the union of finitely many nowhere dense subsets of a topo-
logical space must itself be nowhere dense in the space.
1.54 Prove that a topological space is of the first category in itself if and only if
it is the union of countably many closed sets each having empty interior.
1.55 Prove that if Xl and X 2 are topological spaces at least one of which is of
the first category in itself, then thc topological product Xl x X 2 is of the
first category in itself.
1.6 Three Fundamental Theorems 41
1.56 The closure of a nowhere dense set is obviously nowhere dense. Must the
closure of a set of the first category in a topological space X be of the
first category in X?
1.57 Prove that the boundary of a closed subset of a topological space must
be nowhere dense in the space. Conclude that a subset of a topological
space is nowhere dense in that space if and only if it is a subset of the
boundary of some closed set.
1.58 A subset A of a topological space X is nearly open if there are subsets
Ml and M2 of the first category in X such that (A \ M 1 ) U M2 is open.
Prove that every Borel subset of a topological space is nearly open. (The
preceding exercise might be helpful.)
1.59 An element x of a subset A of a metric space is an isolated point of A if
there is a positive f such that no point of A besides x itself has distance
to x less than f. A subset of a metric space is perfect if it is closed and
has no isolated points. Prove that every nonempty perfect subset of a
complete metric space is uncountable. Use this to give a proof that the
reals are uncountable without using the usual diagonalization argument.
Prove in a similar way that the Cantor set is uncountable.
1.60 Prove that the rational numbers are not a G6 subset of the reals. (Recall
that a G6 set is a set that is the intersection of countably many open
sets.)
1.61 Suppose that f is any function from R into R Prove that the points of
continuity of f form a G6 subset of R. Conclude that no function from IR
into R is continuous precisely on the rationals. Is the same true for the
irrationals? (The preceding exercise may help.)
1.62 Show that the real line can be partitioned into two subsets, one of which
is of the first category in IR and the other of which has Lebesgue measure
zero. Thus, though the real line is large in both Baire's topological sense
and Lebesgue's measure-theoretic sense, it can be written as the disjoint
union of a topologically small set and a measure-theoretically small set.
1.63 Prove that every nonempty locally compact Hausdorff space is of the
second category in itself.
in many analysis texts. See, for example, [67] and [200], as well as Sec-
tion 2.3 of this book, all of which contain more general versions of these
three results that are proved without using ZabreYko's lemma.
Let G = {x : x E X, p(x) < I}, the "open unit ball" for p. If t > 0,
then tG = {x : x E X, p(x) < t} by an application of property (1) in
Definition 1.6.1, and so x E tG whenever x E X and t > p(x). Thus, the
set G is absorbing. If x, y E G and 0 < t < 1, then
1.6.5 The Open Mapping Theorem. (J. Schauder, 1930 [208]). Every
bounded linear operator from a Banach space onto a Banach space is an
open mapping.
44 1. Basic Concepts
Therefore P(2:n Yn) ::; 2:n P(Yn) since E is an arbitrary positive number,
and so p is countably subadditive. This also implies that P(YI + Y2) ::;
p(Yd + P(Y2) whenever Yl, Y2 E Y, as can be seen by letting Yn = 0 when
n 2: 3. Thus, the function P is a countably subadditive seminorm on Y, and
so is continuous by Zabrelko's lemma. Finally,
so T(U) is open.
A one-to-one map from one topological space onto another is a homeo-
morphism if and only if it is both continuous and open. Combining this
with the preceding theorem immediately yields the following result.
1.6.6 Corollary. (S. Banach, 1929 [11]). Every one-to-one bounded linear
operator from a Banach space onto a Banach space is an isomorphism.
1.6.7 Definition. Two norms on the same vector space are equivalent if
they induce the same topology.
It follows from Corollary 1.6.6 that if 11111 and 11112 are two Banach
norms on the same vector space X and the identity operator, viewed as a
linear operator from (X, 11111) onto (X, 11112), is continuous, then a subset
of X is open with respect to one of the norms if and only if it is open with
respect to the other. Restating this in the language of Definition 1.6.7 gives
the following result.
1.6.8 Corollary. Suppose that 11111 and 11112 are two Banach norms on
a vector space X and that the identity map from (X, 11111) to (X, 11112) is
continuous. Then the two norms are equivalent.
IIT(Lxn
n) II = IILTXnl1 : ; L
n n
IITxnl1 ::; LP(x
n
n ),
46 1. Basic Concepts
A linear operator from one Banach space into another is sometimes called
a closed mapping if it satisfies the hypotheses of the next theorem (but see
Exercise 1.74). For that reason, the following result is often called the closed
mapping theorem.
1.6.11 The Closed Graph Theorem. (S. Banach, 1932 [13]). Let T be
a linear operator from a Banach space X into a Banach space Y. Suppose
that whenever a sequence (x n ) in X converges to some x in X and (Txn)
converges to some y in Y, it follows that y = Tx. Then T is bounded.
PROOF. Let p(x) = IITxl1 for each x in X. It is enough to prove that
p is continuous, for then there would be a neighborhood U of 0 such
that the set p(U) is bounded, which would in turn imply that T(U) is
1.6 Three Fundamental Theorems 47
IIT(2: Xn ) =
n
I II2: Txn l ~ 2:IITxnll,
n n
which shows that p is countably subadditive and finishes the proof.
Exercises
1.64 Suppose that X is a Banach space and T: X -> loo is a linear operator.
For each n in N, let (Tx)n be the nth tenn of Tx and let fn be the linear
functional on X that maps x to (T:!:)n. Prove that T is bounded if and
only if each In is bounded.
1.65 Repeat the preceding exercise, replacing loo by 1.
1.66 Prove that if T is a bounded linear operator from a Banach space X onto
a Banach space Y, then whenever a sequence (Yn) converges to a limit Y
in Y, there is a sequence (x .. ) converging to an x in X such that Tx = Y
and TX n = Yn for each n.
1.61 Suppose that X and Y are metric spaces and that I is a function from X
into Y. Prove that the graph { (x,J(x) : x EX} of f is closed in X x Y
if and only if Y = f(x) whenever a sequence (x n ) converges to x in X and
(I(x .. ) converges to y in Y. (This is the source of the name of the closed
graph theorem. Suppose that X and Yare Banach spaces and that T is a
linear operator from X into Y. Then the closed graph theorem says that
T is bounded if its graph is closed in X x Y. In fact, the operator T is
bounded if and only if its graph is closed in X x Y. See the next exercise.)
48 1. Basic Concepts
1.68 Prove the following converse of the closed graph theQrem: If / is a contin-
uous function from a topological space X into a Hausdorff space Y, then
the graph of / is closed in X x Y.
1.69 Let X be a normed space and let X = B(X, JF). Suppose that (Xn) is
a sequence in X such that Ln n
x'x converges whenever x E X'. Show
that the mapping x f-7 Ln n
x'x is a bounded linear functional on X'.
1. 70 Suppose that 1 ::; P ::; 00 and that T is a linear operator from Lp [0, 1]
into itself with the property that if (fn) is a sequence in Lp[O, 1] that
converges almost everywhere to some / in Lp[O, 1], then (T/n) converges
almost everywhere to T/. Prove that T is bounded.
1.71 Suppose that X and Yare Banach spaces and that ;J is a family of
continuous functions from Y into a Hausdorff space such that whenever
Y1, Y2 E Y and Y1 "I Y2, there is an /YI,Y2 in ;J for which /YI,Y2(Y1) "I
/YI,Y2(Y2). Prove that if T is a linear operator from X into Y such that
/ 0 T is continuous for each / in ;J, then T is bounded.
1. 73 Obtain the uniform bounded ness principle directly from Theorem 1.3.14.
(With all notation as in the statement of the uniform boundedness prin-
ciple, let C = {x: x E X, IITxlI :S 1 for each T in ;J }.)
1. 74 The term closed mapping is used both for a linear operator between Ba-
nach spaces that satisfies the hypotheses of the closed graph theorem and
for a mapping from one topological space to another that always maps
closed sets onto closed sets. The two meanings are not equivalent for linear
operators between Banach spaces, as will be shown in this exercise.
(a) Suppose that X and Yare normed spaces and T is a linear operator
from X into Y that is neither one-to-one nor the zero operator. Find
a closed subset F of X such that T(F) is not closed in Y.
(b) Find a linear operator T that satisfies the hypotheses of the closed
graph theorem even though there is a closed subset F of the domain
of T such that T(F) is not closed in the range of T.
n-l
IITnxnll2 n+ E II TnXj II (or 21 ifn = 1);
j=1
all translates of M along with the vector space operations given by the
formulas
(x + M) + (y + M) = (x + y) + M
and
+ M) = (ex x) + M.
ex (x
For each x in X, the tran~late x + M is called the coset of M containing x.
It is thus all right to treat the vector space operations of Definition 1. 7.1
as being the set operations of Section 1.1 as long as this one discrepancy
is kept in mind. The context should always prevent confusion about the
meaning of O(x + M).
Now suppose that M is a subspace of a normed space X. It is reasonable
to ask if the norm of X induces a norm on XI M in some natural way.
Such a norm clearly cannot come from the "formula" Ilx + Mil = Ilxll when
M =I- {O}, for y + M = 0+ M whenever y E M even if Ilyll =I- 11011. This is
one situation in which it is helpful to think first about distance and then
recover the norm from the notion of distance. There is a natural way to
define the distance between two cosets x + M and y + M, namely, the
way in which the distance between subsets of a metric space is defined in
Section 1.1:
Also, the formula from Section 1.1 for the distance between a point and a
set in a metric space can be used to find the distance between an element x
of X and a member y + M of X/M:
the set M, or as the distance from the origin of X to the set x + M, since
d(x, M) = d(x + M,O + M) = d(O, x + M). Therefore
Ilx + Mil = inf{ Ilx - zll : z EM} = inf{ Ilx + zll : z EM}
whenever x EX.
II(x+M)+(y+M)11 = II(x+y)+MII
S Ilx + y + Zl + z211
s Ilx + zlll + Ily + z211
Taking appropriate infima shows that
of each such line being its distance from the origin, that is, the absolute
value of its a-intercept. Some of the proofs in this section are based on
geometric ideas that are easier to visualize if this example is kept in mind.
1.7.13 Theorem. Suppose that X and Yare normed spaces and that
T is a linear operator from X into Y, not assumed to be bounded. Suppose
further that M is a closed subspace of X such that M ~ ker(T) and that 7r
is the quotient map from X onto X/M. Then there is a unique function 8
from X/M into Y such that T = 80 7r, that is, such that the following
diagram commutes.
This map 8 is linear and has the same range as T. The operator 8 is an
open mapping if and only if T is an open mapping, and is bounded if and
only if T is bounded. If T is bounded, then 11811 = IITII.
56 1. Basic Concepts
an open set. This shows that 5 is an open mapping and completes the
proof.
The preceding theorem may have a familiar ring to it, for it is the normed-
space analog of an important result from abstract algebra: If N is a normal
subgroup of a group G and f is a group homomorphism from G into a
group H such that N c::: ker(f) , then there is a unique group homomor-
phism fa from the quotient group GIN into H such that f(g) = fo(gN)
for each 9 in G. This result is used to prove the first isomorphism theorem
for groups, which says that if f: G --t H is a group homomorphism, then
G/ker(f) and f(G) are isomorphic as groups. Theorem 1.7.13 can be used
to obtain an analogous result for Banach spaces.
T = S7r,
This section concludes with two applications of quotient maps. The first
gives a nice test for the continuity of a finite-rank linear operator between
normed spaces.
For the second application, suppose that M and N are closed subspaces
of a normed space X. It is easy to see that M + N is also a subspace of X.
However, the set M + N might not be closed; see Exercise 1.84. There is
one situation, however, in which it must be.
Exercises
1.77 Prove that II(an)+eol! = lim sUPnlan I for each element (an)+eo ofioo/eo.
1.78 Let M = {f : f E e[O, 1], f(O) = O}, a closed subspace of e[O, 1]. Find
a simple expression for the quotient norm of e[O, 1]IM. What familiar
normed space is isometrically isomorphic to this quotient space?
1. 79 Give an example to show that the isomorphism guaranteed by the first
isomorphism theorem need not be an isometric isomorphism. Exercise 1.49
might help.
1.80 Display a bounded linear operator from a Banach space onto an incom-
plete normed space. What does this say about trying to extend the first
isomorphism theorem to incomplete normed spaces?
1.81 Suppose that T is a bounded finite-rank linear operator from a normed
space X into a normed space Y. Prove that XI ker(T) ~ T(X), whether
or not either X or Y is complete.
1.82 Show that the conclusion of Theorem 1.7.15 does not always hold for
linear operators not having finite rank. Exercise 1. 75 (c) may be helpful.
1.83 Prove that finite-dimensionality is a three-space property.
1.84 The purpose of this exercise is to show that the sum of two closed sub-
spaces of a Banach space need not be closed. Let M and N be the closed
subspaces of Co defined by the formulas
1.86 The purpose of this exercise is to prove this theorem: Suppose that X
and Yare normed spaces and that M is a closed subspace of X such
that X/M is finite-dimensional. If T E B(M, Y), then there is a To in
B(X, Y) such that To agrees with T on M. That is, every bounded linear
operator with domain M has a bounded linear extension to X.
(a) Show that the theorem is true if the dimension of X/M is zero.
Now assume that the dimension of X/M is a positive integer n. Let
Xl + M, ... ,Xn + M be a basis for X/M and let Z = (Xl, ... ,x n ).
Show that for each X in X there is a unique m(x) in M and a unique
z(x) in Z such that X = m(x) + z(x).
(b) Show that the mappings x f-> m(x) and x f-> z(x) are bounded linear
operators from X onto M and Z respectively.
(c) Prove the theorem.
1.87 Find the second and third isomorphism theorems for groups in your fa-
vorite abstract algebra text; for example, see [106].
(a) State and prove a Banach-space analog of the second isomorphism
theorem. You might need to impose some conditions on the pertinent
subspaces beyond just requiring them to be closed.
(b) Do the same for the third isomorphism theorem.
If Xl"'" Xn are normed spaces, then there is a way to norm their vector
space sum that is suggested by the norm of Euclidean n-space.
The subscripts that appear on the norms of Xl, ... ,Xn in the preceding
definition were included only for clarity, and will usually be omitted when
there is no possibility of confusion.
Of course, it must be shown that the formula in Definition l.8.1 actually
does define a norm on the vector space sum of Xl, ... , X n . The triangle
inequality follows from the triangle inequality for the norm of Euclidean
n-space, for if (Xl, ... ,X n ), (Yl, ... ,Yn) E XIX .. X X n , then
are often used instead of the norm given in Definition 1.8.1. Fortunately,
these three norms turn out to be equivalent; see Exercise 1.88. The reason
for choosing the particular norm used here is made clear in Section 1.10.
Just as the symbol L:7=1 aj is often used to represent the sum of scalars
001, ... ,an, there are compact "sigma" notations to represent the direct
sum of normed spaces Xl' ... ' X n . Unfortunately, there seem to be about
as many such notations in use as there are people to invent them. For
example, the notations L:ffio
n
Xj,
n
Be 2 Xj, and
(
L:n Xj
)
are all used,
J=l J=l J=l /0
where the 2 indicates in an obvious way the nature of the direct sum norm
and is sometimes omitted or replaced by "2 or just 2.
When a normed space direct sum has only one summand Xl, the no-
tation of Definition 1.8.1 does not distinguish between the summand Xl
and the direct sum X 1. This is no problem, for in this situation it is com-
mon practice to identify the direct sum with the summand. Since the map
1.8 Direct Sums 61
x ~ (x) from the summand onto the direct sum is clearly an isometric
isomorphism, the convention is justified.
When (Xl, dd, . .. , (Xn , dn ) are metric spaces, there is a standard way
to define a product metric on Xl x ... X X n , namely, by the formula
The next result says that the direct sum of normed spaces X I, ... , Xn
always includes isometrically isomorphic copies of the Xj's in the places
where one would most expect to find them.
1.8.3 Proposition. Let Xl, ... , Xn be normed spaces. For each integer j
such that 1 :s: j :s: n, let
to Xl ffi ... ffi Xn' An example will tell most of the story. Suppose for the
moment that n = 5, and consider the direct sum (X4ffi(X5ffiX2))ffiX3ffiXI
formed from Xl ffi X 2 ffi X3 ffi X 4 ffi X5 by permuting and associating its
terms. It is easy to check that the map
from (X4 ffi(X 5 ffiX2 )) ffiX3 ffiX I onto X I ffiX2 ffiX3ffiX4 ffiX 5 is an isometric
isomorphism. For any n and any direct sum Y formed by permuting and
associating the terms of Xl ffi ... ffi Xn, the obvious analog of the above
map is clearly an isometric isomorphism from Y onto Xl ffi ffi X n .
1.8.5 Proposition. Let Xl, .. ' ,Xn be normed spaces. Let {I, ... ,n} be
partitioned into two nonempty sets {jl,' .. , jp} and {k 1, ... , k q} and let
Then (Xl EI) .. 'ffiXn )/ Xj" .. ,jp is isometrically isomorphic to Xk, ffi ffiXkq .
PROOF. It is clear that Xj" ... ,jp is a subspace of Xl ffi ... ffi X n . If an
element (Zl, ... , Zn) of Xl ffi ffi Xn is the limit of a sequence of elements
of Xj" ... ,jp, then Zk j , ... ,Zk q must all be zero, so the subspace Xl! ,... ,jp is
closed. Thus, the quotient space (Xl ffi ... ffi Xn) / X j , ''''ljp can be formed.
Define T: (Xl ffi ffi Xn)/Xh, ... ,jp -> X k , ffi ffi Xk q by the formula
T((XI, .. " xn) +Xjj, ... ,jp) = (Xk 1 , , Xk q )' Since (Xl"'" Xn) +Xj" ... ,jp =
(Yl,"" Yn) + Xj".,jp if and only if Xkm = Yk m when m = 1, ... , q, the
map T is well-defined. It is easy to check that T is linear, one-to-one, and
onto X k , ffi .. ffi X kq . If (Xl, ... , Xn) + X]" .,jp E (Xl ffi .. ffi Xn) / X j , , ... ,]p'
then
II(Xl,''''Xn ) +Xj.,.,jpll
= d((O, ... , 0), (Xl, ... , Xn) + Xh,.jp)
= inf{ C~IIYmI12 r/ 2
: (YI, ... , Yn) E Xl ffi .. ffi Xu,
Yk , = Xk
"
... ,Ykq = Xk q }
= II(Xkl'""Xkq)11
= IIT((Xl, ... ,Xn)+Xj" .. ,jp)ll,
which is why the proposition can be called a cancellation law. This cancel-
lation law would be more appealing visually if the symbol 0 were used to
separate the terms of a direct sum, but that symbol is reserved for tensor
products. Because of their product-like behavior and their relationship to
Cartesian products, direct sums are sometimes called direct products, but
then so are tensor products. For this reason, the term direct product will
not be used in this book.
A metric space is called topologically complete if its metric is equivalent
to a complete metric, that is, if the topology of the space is induced by some
complete metric, even if the given metric is not complete. If Xl, ... ,Xn are
metric spaces, then the product topology of Xl x ... X Xn is topologically
complete if and only if each Xj is topologically complete; see, for exam-
ple, [65]. This suggests the following theorem, though it does not prove it.
As was shown in Exercise 1.42, it is quite possible for an incomplete metric
space to be topologically complete. 5
The internal direct sums that are of the most importance in the theory
of normed spaces have an additional restriction on the subspaces used to
form them.
That is, a normed space X is the internal direct sum of its subspaces
M I , ... ,Mn if and only if each of these subspaces is closed and X is their
algebraic internal direct sum.
1.8.9 Example. In e[O, 1], let Ml be the collection of all constant func-
tions, let M2 be the collection of all linear functions 1 such that 1(0) = 0,
and let M3 be the collection of all 1 such that 1(0) = 1(1) = o. It is easy to
check that M I , M 2 , and M3 are closed subspaces of e[O, 1] and that e[O, 1]
is the internal direct sum of the three.
Notice that the use of the words external and internal is optional when
referring to normed space direct sums. This means that a statement such
as "the normed space X is the direct sum of Y and Z" could be ambiguous.
Normally, the context does make it clear which type of direct sum is being
considered. Even when it does not, the two types of direct sums are so
closely related that problems usually do not arise. This relationship is the
subject of the next proposition.
1.8.10 Proposition.
(a) If XI, ... ,Xn are normed spaces and X =XIffiffiXn , then X has
closed subspaces X~ , ... ,X~ such that X is the internal direct sum of
X~ , ... ,X~ and each Xj is isometrically isomorphic to the corre-
sponding Xj.
(b) If X is a Banach space that is the internal direct sum of its closed
subspaces M I , ... , M n , then X ~ MI ffi ffi Mn
PROOF. For (a), let X~, ... ,X~ be as in Proposition 1.8.3. It is easy to check
that these closed subspaces of X do what is needed. For (b), suppose that
X is a Banach space that is the internal direct sum of its closed subs paces
M I , ... , Mn. Since each M j is a Banach space, so is MI ffi .. ffiMn . The map
T: MI ffi . ffiMn --> X given by the formula T(ml' ... , m n ) = ml + .. +mn
is clearly linear and onto X. By Corollary 1.4.18, the "1 and 2 norms on lF n
are equivalent, from which it follows that there is a positive constant c such
that
for each member (ml, ... , m n ) of Ml ffi ... ffi Mn Thus, the operator T
is bounded. If T(ml, .. ,mn ) = 0, then m] = ... = mn = 0 by Proposi-
tion 1.8.7. It follows that T is one-to-one. By Corollary 1.6.6, the operator T
is an isomorphism from MI EEl ffi Mn onto X.
Part (b) of the preceding proposition is not in general true for incomplete
normed spaces. See Exercise 1.95.
66 1. Basic Concepts
The notion of the direct sum of normed spaces leads in a very natural
way to the notion of the direct sum of linear operators between normed
spaces.
1.8.11 Definition. Suppose that Xl,." ,Xn and Y I , ... , Y n are normed
spaces and that T j is a linear operator from Xj into Yj when j = 1, ... , n.
Then the direct sum of T I , ... , Tn is the map
TI EB ... EB Tn: X I EB ... EB Xn ---> Y I EEl . EB Yn
defined by letting
TI EEl .. EEl Tn (Xl, ... , Xn) = (TIXI, ... ,Tnxn)
whenever (Xl, ... ,Xn ) E Xl ... X n .
1.8.12 Theorem. Suppose that XI"",Xn and YI, ... ,Yn are normed
spaces and that T j is a linear operator from Xj into Yj when j = 1, ... , n.
Then TI EB ... EB Tn is bounded if and only if each T j is bounded. 1
TI EB ... EEl Tn is bounded, then IITI EEl .. EEl Tn II = max{ IITIII, ... , IITn II}
Furthermore, the operator TI EEl . EB Tn is one-to-one, or onto, or an iso-
morphism, or an isometric isomorphism, if and only if T I , ... , Tn all have
that same property.
PROOF. Suppose that TI is unbounded. Then
sup{ IITI ... EEl Tn(XI,"" xn)11 : (Xl, ... , Xn) E BX1(f1"'(f1Xn }
~ SUp{ IITI EB EB Tn(Xb 0, ... ,0)11 : Xl E BX, }
which implies that TI EB ... EB Tn is bounded and that IITI EB ... EB Tn II :::::
max{IITlll, ... , IITnll} Notice next that whenever Xl E B x"
so IITI EB ... EB Tnll ;::: IITIII It follows similarly that liT! EB ... EB Tn II ;::: IITj II
when j = 2, ... , n, so liT! EB ... EB Tnll ;::: max{IITIiI,, IITnll}. Therefore
IITI EB EB Tn II = max{IITIII,, IITnll}
It is easy to see that TI EB .. EB Tn is one-to-one, or onto, or an isometric
isomorphism, if and only if each T j has the corresponding property. All
that remains to be proved is that TI e ... EB Tn is an isomorphism if and
only if each Tj is so. The failure of one of the Tj's to be one-to-one would
prohibit both TI EB ... EB Tn and that Tj from being isomorphisms, so it
can be assumed that each T j is one-to-one. Since TI EB ... EB Tn can be
viewed as an operator onto the normed space T(X I ) EB ... EB T(Xn)' no
harm comes from also assuming that each T j is onto the corresponding 1j.
It is then easy to see that (TI EB EB Tn)-l = T I- 1 EB EB T;:l, and so
(TI EB ... EB Tn)-l is bounded if and only if each T j- l is bounded. Thus,
the operator TI EB ... EB Tn is an isomorphism exactly when each Tj is an
isomorphism.
Exercises
1.88 Suppose that Xl, ... ,Xn are normed spaces and that 1 ::: p ::: 00. To
avoid an impending notational conflict, let the norm of be denoted e;
by II Ilf~ instead of II lip Show that the formula
must define a norm on the vector space sum of Xl, ... , Xn equivalent to
the usual direct sum norm.
1.90 Suppose that X is Cu or lp, where 1 ::: p ::: CX>. Prove that X 8 IF "" X.
Conclude that X EB Y "" X whenever Y is a finite-dimensional normed
space.
1.91 Suppose that X is Co or l", where 1 ::: p ::: 00. Prove that X eX"" X.
1.92 Prove that L1'[O, 1] EEl L1'[O, 1] "=' Lp[O, 1] when 1 ::: p::: 00. One way to do
this requires some knowledge of how to change variables in the Lebesgue
integral. See, for example, [202, pp_ 153-155].
1.8 Direct Sums 69
1.93 (a) Suppose that X, Y, and Z are normed spaces such that Y ED Y ~ Y
and Z is the direct sum of X with a finite positive number of copies
of Y. Prove that Z ~ X ED Y.
(b) Suppose that X and Y are Banach spaces such that Y ED Y ~ Y
and X has a complemented subspace isomorphic to Y. Prove that
X ED Y ~ X, and therefore that every direct sum of X with a finite
number of copies of Y is isomorphic to X.
1.99 Let T be a bounded linear operator from a Banach space X into a finite-
dimensional normed space Y. The purpose of this exercise is to show that
'1' is, in a sense, the direct sum of a Banach space isomorphism and a zero
operator.
(a) Define a subspace M of X as follows. First show that XI ker(T)
and T(X) have the same finite dimension n. If T(X) = {OJ, then
let M = {OJ. If T(X) {OJ, then let Xl + ker(T), ... ,Xn + ker(T)
be a basis for XI ker(T) and let M = (Xl, ... , Xn). Show that in
the second case the vectors Xl, . . . ,X n are linearly independent, and
70 1. Basic Concepts
= aT(w) + T(x).
a scalar O:x such that icl:xl = 1 and O:xf(x) is a nonnegative real number.
Notice that lu(x)1 :S If(x)1 = f(o:xx) = u(O:xx) whenever x EX. Therefore
It is time for the main results of this section. There is a body of related
facts, collectively called the Hahn-Banach theorem, that includes Theo-
rems 1.9.5 and 1.9.6 and Proposition 1.9.15 from this section as well as
Theorems 2.2.19, 2.2.26, and 2.2.28 in the next chapter. The common theme
of all of them is that under certain conditions a vector space always has a
large enough supply of well-behaved linear functionals to accomplish cer-
tain tasks.
and so
It follows that
sup{ 10(Y) - p(y - Xl) : y E Y} ::; tl ::; inf{p(y + Xl) - 10(Y) : y E Y}.
Let lI(y + txd = 10(Y) + t tl for each y in Y and each real number t.
It is easy to check that II is a linear functional on YI whose restriction
to Y is 10. It follows from the definition of tl that for each y in Y and each
positive t,
and
lI(y - txd = t(Jo(Cly) - h) ::; tp(Cly - Xl) = p(y - tXI),
so II (x) ::; p(x) whenever X E Y I .
The second step of the proof is to show that, in effect, the first step
can be repeated until a linear functional on all of X is obtained that is
dominated by p and whose restriction to Y is fo. Let 21 be the collection
of all linear functionals 9 such that the domain of 9 is a subspace of X
that includes Y, the restriction of 9 to Y is 10, and 9 is dominated by p.
Define a preorder j on 21 by declaring that 91 j 92 whenever 91 is the
restriction of 92 to a subspace of the domain of 92. It is easy to see that
each nonempty chain I! in 21 has an upper bound in 21; consider the linear
functional whose domain is the union Z of the domains of the members
of I! and which agrees at each point z of Z with every member of I! that
is defined at z. Of course, the empty chain has 10 as an upper bound. By
Zorn's lemma, the preordered set 21 has a maximal element I. The domain
of I must be all of X, for if it were not then the argument of the first step of
this proof, applied to f and its domain instead of to fo and Y, would yield
an II in 21 such that I j II but II ii I. This I does all that is required
~rt.
The next result is sometimes called the analytic form of the Hahn-Banach
theorem. It was first proved by Hahn [99] in 1927 for real normed spaces.
Banach [11] independently published the same result with the same proof
in 1929, but later became aware of Hahn's earlier paper and acknowledged
Hahn's priority. Bohnenblust and Sobczyk [32] gave the extension to com-
plex normed spaces in 1938. In the same year, Soukhomlinoff [227] inde-
pendently published the same result, and showed that it even holds for the
generalization of normed spaces for which the scalars are quaternions.
1.9 The Hahn-Banach Extension Theorems 75
whenever y E Y, and so
Since d(x, Y) > 0, it follows that Ilfoll 2: 1, and so Ilfoll = 1. To finish, let f
be any Hahn-Banach extension of fa to X.
Letting Y = {O} in the preceding corollary yields the first of the following
two results. The second then follows from the first by replacing x by x - y.
aij ) f31 + .. +a~) i3n = Cj when j = 1, ... , m. It is a standard fact from linear
algebra that there is a one-to-one correspondence between (IF'n)# and IF'n
such that for each f in (IF'n) # and the corresponding element (131, ... , f3n)
of IF''',
h
wenever (al, ... ,a n ) E IF' n . L e t x) -- ( a1, h
(j) ... ,a n(j)) wen]. -- 1, ... ,m.
Then the problem of finding a solution to the given system becomes that
of finding a linear functional Ion IF'n such that f(x)) = Cj for each j.
Now suppose that Y is a subspace of a normed space X and that fa is
a bounded linear functional on Y. For each y in Y, let c y = 10(Y). The
normed space version of the Hahn-Banach extension theorem says that
there is a bounded linear functional I on X with the same norm as fo such
that fey) = c y for each y in Y. In a sense, a system of linear equations has
been solved. Of course, this system is infinite if Y =J {O}.
It is not surprising that the Hahn-Banach extension theorems have im-
portant applications to solving systems of linear equations. The following
one is due to Hahn himself.
1.9 The Hahn-Banach Extension Theorems 77
1.9.10 Theorem. (H. Hahn, 1927 [99]). Suppose that X is a normed space.
Let A be a non empty subset of X and let { Cx : x E A} be a corresponding
collection of scalars. Then the following are equivalent.
(a) There is a bounded linear functional I on X such that I(x) = Cx for
each x in A.
(b) There is a nonnegative real number M such that
t
I3=1 O!jCXj - f
3=1
f3jCYj I ::; Milt
3=1
O!jXj - f
3=1
f3jyjll = 0,
for each linear combination adl + ... + anfn of h, ... , In, that is,
for each element of ({h, .. , in}).
ff (b) holds and E > 0, then Xo can be chosen in (a) so that Ilxo II 0<:: M + E.
PROOF. If (a) holds, then whenever al,"" an E IF,
Conversely, suppose that (b) holds and that E > 0. If each Cj is zero,
then both (a) and the remark following (b) hold when Xo = 0, so it can
be assumed that some Cj is nonzero. It follows from the inequality in (b)
that some fj is nonzero, so it can be assumed, after rearranging the iJ's
if necessary, that there is an integer m for which h, ... ,fm is a maximal
linearly independent subcollection of h, ... , f n' Thus, for each k such that
(k) (k) _ ",m (k)
k = 1, ... ,n there are scalars 0: 1 , ... , O:m such that fk - u j = l O:j fj.
Suppose that it were known that (b) implies (a) and the remark fol-
lowing (b) under the additional assumption that the linear functionals
h, ... ,fn are linearly independent. In the current situation in which only
h, ... , fm are known to be linearly independent, it would still follow that
there is an Xo in X such that Ilxoll ::; M + E and iJ(xo) = Cj when
j = 1, ... ,m, which implies that fk(XO) = 2::;:1 o:}k)Cj for each k. It would
then follow that
that is, that h(xo) = Ck, when k = 1, ... ,n. Thus, it can be assumed that
h, .. , fn are linearly independent.
For each x in X, let T(x) = (hx, ... , fnx). Then T is a linear operator
from X into Fn. Suppose for the moment that n 2': 2. For each k such that
n
k = 1, ... , n, Lemma 1.9.11 implies that j # ker(fJ) %ker(h), so there
is a Yk in X such that fk(Yk) = 1 and fJ(Yk) = 0 when j =1= k. It follows
that T(X) includes the standard basis for Fn, so T maps X onto Fn. If
n = 1, then T obviously maps X onto Fn. In any case, there is a Yo in X
such that (hyo, ... , fnYo) = (C1,"" cn). Since Yo tf- n7=1
ker(fJ), it follows
from Corollary 1.9.7 that there is a bounded linear functional f on X such
that Ilfll = 1, f(yo) = d(yo, n7=1ker(fJ)), and n7=1
ker(Jj) S;;; ker(J).
Another application of Lemma 1.9.11 yields scalars (31, ... ,!3n such that
f = 2:7=1 (3jiJ. Therefore
so there is a Zo in n7=1
ker(iJ) such that IIYo - Zo I ::; M + E. It follows
that Yo - Zo is an Xo that docs what is needed in (a) and in the remark
following (b).
where under certain conditions it is even possible to assert that the in-
equality is strict. In a sense, the functional f separates 0 1 from C 2 , since
f(Od and f(02) are disjoint or nearly so. Such separation theorems are
based on a very close relationship that exists between nonnegative sublinear
functionals and convex absorbing sets.
L( an ) <
_ 11msup
a1 + ... + an
n n
and L(a n ) is the (C, I)-limit of (an) if it has one. (The functional L
is called a Banach limit function on 00, and the "limits" that L
assigns to the members of = are said to form a system of Banach
limits.)
(b) Prove that
. f an < 11mln
11m1n . f a1 + ... + an
n - n n
:::; L(an )
. cq + ... + an
:::; hmsup------
n n
:::; lim sup an
n
for each member (an) of 00. Conclude from this that the notion of
Banach limit is a generalization of the notion of (C, I)-limit, which
is in turn a generalization of the usual notion of limit. Show that
each generalization is proper, that is, that no two of the notions are
equivalent.
(c) Prove that L(an ) = L(ak+1, ak+2, ... ) for each member (an) of =
and each positive integer k, that is, that L is shift-invariant. There-
fore Banach limits of bounded sequences, like regular limits, depend
only on the tails of the sequences and not on their leading terms.
(d) Of the three notions of limit discussed in (b), for which ones is it
true that each subsequence of a convergent bounded sequence must
converge to the limit of the entire sequence?
(e) Show that the linear functional L found in part (a) is not unique,
that is, that there is more than one system of Banach limits for 00.
1.103 Suppose that (jn) is a bounded sequence of bounded linear functionals on
a real normed space X. Show that there is a bounded linear functional f
on X such that liminfn fn(x) :::; f(x) :::; limsuPn fn(x) whenever x E X.
Conclude from this that if limn fn(x) exists whenever x E X, then the
formula f(x) = limn fn(x) defines a bounded linear functional on X.
(Compare Corollary 1.6.10.)
1.104 Let Y be a subspace of a nOfmed space X and let 2l be the collection of
all bounded linear functionals f on X such that Y ~ ker(f). Prove that
17 = n{ker(j): f E 2l}.
1.105 Let X be the vector space IR2 and let Y be the subspace of X consisting of
all (a, (3) such that f3 = o. Define fa: Y - t lR by the formula fo(a, 0) = a.
Notice that fa is a bounded linear functional on Y with respect to any
norm given to X.
84 1. Basic Concepts
(a) Suppose that X is given the Euclidean norm. Show that fo has a
unique Hahn-Banach extension to X.
(b) Suppose that X is given the norm that makes it into fi. Show that
fo has an infinite number of different Hahn-Banach extensions to X.
1.106 Let X be a normed space and let W be a finite-dimensional subspace of
B(X, IF). Suppose that F is a linear functional on B(X, IF). Show that
there is an Xo in X such that F(w) = w(xo) for every w in W. If F is
bounded and E > 0, show that Xo can be chosen so that IIxoll :s; IIFII + E.
1.107 Which of the properties positive-homogeneity, subadditivity, and sublin-
earity guarantee that if a real-valued function p on a vector space has that
property, then p(O) = O?
1.108 (a) Prove that every norm is the Minkowski functional of the closed unit
ball of the corresponding normed space.
(b) Find a nonconvex absorbing subset of some vector space such that
the Minkowski functional of the set is a norm.
1.109 Let X be a normed space. If t E lR and f is a nonzero bounded linear
functional on X, then the subset of X defined by the formula
is called a closed half9pace. Show that if X i= {O}, then each closed convex
subset of X is the intersection of some collection of closed halfspaces.
(Don't forget special cases.) What can be said when X = {O}?
1.110 Suppose that C is a nonempty convex subset of a normed space X and
that Xo is an element of X \ C such that d(xo, C) = O. Proposition 1.9.15
might lead one to conjecture that there must be some nonzero bounded
linear functional f on X such that Re f(xol ~ sup{ Re f(x) : X E C}.
Find a counterexample. (One can be constructed using a proper dense
subspace of a normed space.)
1.10.1 Definition. (H. Hahn, 1927 [99]). Let X be a normed space. The
(continuous) dual space of X or dual of X or conjugate space of X is the
normed space B(X, IF) of all bounded linear functionals on X with the
operator norm. This space is denoted by X*.
1.10 Dual Spaces 85
1.10.2 Example. Let (O,~, f..l) be a a-finite positive measure space. Fix p
such that 1 :S p < 00 and let Lp denote Lp(O,~,f..l). Let q be conjugate
to p; that is, let q = 00 if p = 1 and let q be such that p-l + q-l = 1
otherwise. If x* is a bounded linear functional on Lp, then an argument
involving the Radon-Nikodym theorem yields a g in Lq such that
1.10.3 Example. Suppose that f..l is the counting measure on the collec-
tion ~ of all subsets of N and that 1 :S p < 00_ Let q be conjugate to p.
Since R.p is Lp(N, ~, f..l), it follows from the preceding example that R.; can be
identified with R.q . If the element x* of R.; corresponds to the element (;3n)
of R.q , then the action of x* on an element (an) of R.p is given by the formula
underlying set consists of all the bounded finitely additive scalar-valued set
functions on the subsets of N. See [58] for an excellent discussion of ~.
1.10.4 Example. The dual space of Co is 1 in exactly the same sense that
the dual space of p is q when 1 :'S p < 00 and q is conjugate to p. That is,
there is an isometric isomorphism T from 1 onto Co such that if (f3n) E 1,
then T(f3n) is the linear functional x' on Co given by the formula
(1.3)
n
Unlike the preceding example, this is not just a special case of Exam-
ple 1.10.2 since Co is not a Lebesgue space in any natural sense. However,
it is not difficult to prove directly.
If (f3n) E 1, then 2:nlc~nf3nl :'S 11(f3n)111 II(an)lloo for each (an) in Co,
from which it follows that (1.3) defines a member x* of Co for which Ilx' II :'S
11(/3,,)111. It is then easy to check that the function T: 1 --+ Co that maps
each (f3n) in i1 to the x' defined by (1.3) is a linear operator such that
IIT(f3n) II :'S IIU1n)lll whenever ((3n) E 1, and this operator is one-to-one
since T((3n) # 0 when (f3n) # o.
Now suppose that x* is an arbitrary element of co. For each standard
unit vector en of Co, let f3n = x' en and let In be a scalar such that hn I = 1
and l(3nl = In(3n Then
for each m. The sequence ((3n) is therefore in iI, and 11((3n)lh :'S IIx'll For
each member (an) of Co,
so x* and ((3n) satisfy (1.3) whenever (an) E Co. Since T(f3n) = x, the
operator T maps 1 onto Co.
Finally, suppose that ((3n) is an arbitrary member of l\. Let x = T((3n).
By the argument of the preceding paragraph, there is a member (f3~) of 1
such that T(f3~) = x* and 11((3~)lh :'S Ilx*ll. Since T is one-to-one, it follows
that ((3~) = (!3n) and therefore that 11(!3n)lll :'S IIT(!3n)ll. Since it has al-
ready been shown that IIT(!3n)11 :'S 1I(!3n)lll, the operator T is an isometric
isomorphism, which finishes the proof.
1.10.5 Example. Suppose that n is a positive integer and that 1 :'S p < 00.
Let q be conjugate to p. If Jl is the counting measure on the collection L:
of all subsets of {l, ... ,n}, then i~ is Lp({l, ... ,n},L:,p,), and so (f!~)*
can be identified with f!~ by the method of Example 1.10.2. Notice that if
1.10 Dual Spaces 87
({3l, ... ,(3n) is in f~ and x* is the linear functional that it represents, then
the action of x* on f; is given by the formula
n
The dual space of f;;' can be identified with f'1 in the same way. A proof
of this, using an argument similar to that of Example 1.10.4, is given in
Appendix C; see Theorem C.13. This will also follow from some of the
results of the next section, as will be shown in the comments following
Theorem 1.11. 9.
For the sake of completeness, it should be noted that if 1 ::; p ::; DO and
q is conjugate to p, then (fg)' and f~ are both zero-dimensional, so (fg)'
can be identified with ~ in a very obvious and trivial way.
x*(f) = L f dJ-L.
See [67] or [202] for details. It turns out that rca(K) and C(K)* both have
natural partial orders, and that the isometric isomorphism of the Riesz
representation theorem is order preserving in the sense that if xi and xi
are elements of C(K)* and J-Ll and J-L2 are the respective elements of rca(K)
with which xi and x:i are identified, then xi :; x; if and only if J-Ll :; J-L2.
See Exercise 1.121.
Because of this, one way to show that a normed space is a Banach space
is to show that it is the dual of some other normed space. For example,
if K is a compact Hausdorff space, then rca(K) is a Banach space since it
is isometrically isomorphic to the dual of C(K). See the remarks following
Example 1.2.11.
One of the main reasons for proving the Hahn-Banach extension theorems
before discussing duality is to obtain Corollary 1.9.9, which says that if x
and yare different elements of a normed space X, then there is an x*
88 1. Basic Concepts
in X' such that x'x i= x'y. That is, the dual space of X is large enough
to separate the points of X. Of course, this implies that X* i= {O} when
X i= {O}. As will be seen in Sections 2.2 and 2.3, there are large vector
spaces with metric topologies possessing many of the properties of norm
topologies, even though the spaces have no continuous linear functionals
on them besides the zero functional.
The other corollaries of the normed space version of the Hahn-Banach
extension theorem also have applications to the theory of dual spaces, as
is shown by the proof of the following result.
not contain x n +!. By Corollary 1.9.7, there is a sequence (x~) of elements
of X' such that X~Xj = when j < n, but x~xn is nonzero for each n.
If 2:7=1 D:jX; is a linear combination of the terms of (x~) that equals 0,
then applying this linear combination to Xl, ... ,Xk in that order shows that
D:1 = ... = D:k = 0, from which it follows that (x~) is a linearly independent
sequence in X'. The space X' is therefore infinite-dimensional.
Suppose that X and Yare normed spaces and that y* E Y*, x EX,
T E B(X, Y), and S E B(Y*, X*). The statement y*Tx = Sy*x can have
only one meaning, but that meaning might take a few moments to unravel
since the order in which the operations are to be performed differs on the
two sides of the equation. The use of parentheses helps, but can still lead to
expressions that are visually confusing. For this reason, the notation (x, f)
is sometimes used for Ix when I is a linear functional on X and x E X.
The statement y*Tx = Sy*x then becomes (Tx,y*) = (x,Sy*), which is
easier to grasp. This device is useful in the following result.
PROOF. For (a), suppose that T E L(X, V). Theorem 1.10.9 implies that
IITxl1 = sup{ I(Tx,y*)I: y* E By.} whenever x E X, so
Part (a) follows from this. For (b), suppose that T E L(X, Y*). Then
IITxl1 = sup{ I(y, Tx)1 : y E By} whenever x E X, so
1.10.12 Theorem. Suppose that X and Yare normed spaces such that
there is an isomorphism T from X onto Y. Then the map T*: Y* -+ X*
given by the formula T* (y*) = y*T, where y*T is the usual product of y*
and T, is an isomorphism from y* onto X*, and IIT* II = IITII. If T is an
isometric isomorphism, then so is T*.
PROOF. it is clear that T*(y*) E X* whenever y* E Y*, and equally clear
that T* is linear. Notice that (Tx, y*) = (x, T*y') whenever x E X and
y* E Y*. By Proposition 1.10.11 (a),
so Proposition 1.10.11 (b) implies that T* is bounded and that IIT* II = IITII.
If y* E Y* and T*y* = 0, then y*y = (T(T~ly), y*) = (T~ly, T*y*) = 0
whenever y E Y, so y* = O. It follows that T* is one-to-one. If x* is
a member of X*, then x*x = (Tx,x*T~l) = (x,T*(x*T~l whenever
x E X, so T*(x*T~l) = x*. Thus, the operator T* maps Y* onto X*.
Since T* is a one-to-one bounded linear operator from one Banach space
onto another, it is an isomorphism by Corollary 1.6.6.
1.10 Dual Spaces 91
Y*(XI,""Xn) = LxjXj
j=l
PROOF. For each member (xi, ... ,x~) of Xi EB EBX~, let T(xi, ... ,x~)
be the linear functional on Xl EB ... EB Xn given by the formula
n
that
n
1(X1, ... ,xn),T(xr, , x;,)) I ~ L Ilxillllxjll
j=l
whenever (Xl, ... , Xn) E Xl @ ... @ X n , the functional T(xr, ... , x~) is in
(Xl EB .. EBXn)* and has norm no more than II (xr, ... , x~) II It follows that
T is a linear operator from X; EB ... EB X~ into (Xl EB ... EB Xn)* such that
IIT(xi, . .. , x~)11 ::: II (xi, ... , x~) II whenever (xi, .. , x~) E Xi EB ... EB X~.
Now suppose that y* E (Xl EB . EBXn )*. For each j such that j = 1, ... , n,
it is clear that the function x; on Xj defined by the formula x;(x) =
y*(O, ... ,O,x,O, ... ,O), where x is the lh component of the n-tuple, is
a member of X;. Then T( xi, ... , x~) = y*, which shows that T maps
Xi EB EB X~ onto (Xl EB EB Xn)*
Suppose that (xi, .. , x~) E X; EB ... EB X~. All that remains is to
prove that IIT(xi, ... ,x~)11 = II(xi, .. ,x;JII, so it can be assumed that
(xi, ... , x~) i- 0. It is enough to prove that
It has already been shown that IIT(xi, ... ,x~)II::: lI(xi, .. ,x~)11 = 1, and
so it is enough to show that IIT(xi, ... , x~) II ::::: 1. For each j such that j =
1, ... , n, let (y)kl)k=l be a sequence in BXj such that limklx;yyl I = Ilxj II;
because there is for each j and each k a scalar o:jk) such that Io:;k) I = 1 and
xj(o:jklyyl) = IxjyYll, it can be 3.'l5umed that x;yyl 2:: 0 for each j and
each k and so that limk xjyyl = Ilx; II for each j. Let xjkl = Ilx.; IIYJkl for
each j and each k. Then II (X\k l, ... , x~kl) II ::: II (xi, ... ,x~) II = 1 for each k,
and 50
n
1 = L IIxill2
j=l
= I 1m
k
L x-x-
n
. J
(k)
J
j=l
-I
-
( (kl (k)) , T( Xl''
1m Xl , ... , Xn
Xn*))
k
The sets A-L and 1-B are usually called the annihilators of A and B
respectively, with the qualifying phrases "in X*" and "in X" dropped. In
theory, this could cause confusion when referring to annihilators of subsets
of dual spaces. Since the dual X* of a normed space X is itself a normed
space, both -LB and B-L are defined for each subset B of X*, and both have
the right to be called the annihilator of B. In practice, the context usually
prevents confusion. Where a misunderstanding might occur, the space in
which the annihilator is being taken should be made explicit, either by
adding the qualifying phrase or by using the left-hand or right-hand "perp"
notation.
See Exercise 1.120 for some observations about (.LB).L when B is a sub-
set of the dual space of a normed space. Characterizations of (.LB).L anal-
ogous to those obtained in the above proposition for .L(A.L) will be given
in Proposition 2.6.6.
Suppose that M is a subspace of a normed space X. The restriction of an
x* in X* to M always results in an element of M*. Moreover, if m* E M*,
then the normed space version of the Hahn-Banach extension theorem al-
lows m* to be extended to an element x;". of X*, and the restriction of x;".
to M is m*. Thus, the elements of M* are just the restrictions of the ele-
ments of X* to M, so in one sense M* can be identified with X*. Notice
however that different elements of X* might agree on M, so elements of M*
should really be considered to be equivalence classes of elements of X* , with
two elements of X' considered equivalent if they agree on M. A moments'
thought shows that the resulting equivalence classes are exactly the cosets
x' + M.L that form the elements of X* / M.L. The following result is now
not too surprising.
Exercises
1.111 Let c be the Banach space of all convergent sequences of scalars defined
in Exercise 1.25. Prove that c' is isometrically isomorphic to 1' Notice
from Exercise 1.49 that c and Co are not isometrically isomorphic, even
though their dual spaces are.
1.112 Let Y be a dense subspace of a normed space X. Prove that X' and Y'
are isometrically isomorphic. Use this to give an example of two normed
spaces X and Y such that X' and Y' are isometrically isomorphic, even
though X and Yare not even isomorphic.
1.113 Suppose that X is a complex normed space and that x EX'. Prove that
x is a norm-attaining complex-linear functional if and only if its real part
is a norm-attaining real-linear functional.
96 1. Basic Concepts
1.115 (a) Either prove that every element of t'i is a norm-attaining functional
or display one that is not.
(b) Repeat (a) with t'i replaced by t'~.
1.116 (a) Let the direct sum Xl EB ... EB Xn of normed spaces Xl, ... , Xn be
given one of the norms of Exercise 1.88. Denote this norm by II lip,
Find a norm on X~EB" 'EBX; that makes the map T: XiEB" 'EBX; - t
(Xl EB .. EB Xn)* in the proof of Theorem 1.10.13 into an isometric
isomorphism.
(b) Suppose that the norm II 111 had been used instead of 11112 for the
norm of all direct sums of normed spaces. Let X be the real Banach
space JR. Show that (X EB X EB X)* is not isometrically isomorphic
to X* EB X' EB X*. (Notice that this requires more than just showing
that the mapping T defined in the proof of Theorem 1.10.13 is not
an isometric isomorphism.)
1.117 Let X and Y be normed spaces.
(a) Suppose that x* E X' and y E Y. Define Tx',y: X - t Y by the
formula Tx"Y(x) = (xx)y. Prove that Tx',y E B(X, Y) and that
IITx',yll ~ Ilx*IIIIYII
(b) Suppose that Y =1= {a}. Prove that B(X, Y) has a closed subspace
isometrically isomorphic to X*.
(c) Suppose that X =1= {o}. Prove that B(X, Y) has a closed subspace
isometrically isomorphic to Y.
(d) Suppose that X =1= {a} and Y is not a Banach space. Prove that
B(X, Y) is not a Banach space. (See the comments following Theo-
rem 1.4.8.)
1.118 Let X and Y be normed spaces. Prove that B(X, YO) is isometrically
isomorphic to B(Y, X*).
1.119 Exchanging the roles of X and X' in Theorem 1.9.10 results in the fol-
lowing statement, which will be called statement 1.9.10*:
Suppose that X is a normed space. Let A be a nonempty subset of X*
and let {c x ' : x" E A} be a corresponding collection of scalars. Then the
following are equivalent.
(1) There is an Xo in X such that x'xo = Cx " for each x* in A.
(2) There is a nonnegative real number M such that
1.11.3 Proposition. Let X be a normed space and let (Q(x)) (x*) = x'x
whenever x E X and x* E X*. Then Q(x) E X'* whenever x E X,
and Q is an isometric isomorphism from X into X**. Furthermore, the
subspace Q(X) of X** is closed if and only if X is a Banach space.
There are other ways to prove that every incomplete normed space can
be completed to a Banach space. See Exercise 1.123 for a proof that requires
quite a bit more work than that of Theorem 1.11.5, but has the advantage
that it could have been given almost as soon as Banach spaces and isometric
isomorphisms had been defined.
Suppose that X is a finite-dimensional normed space. It is a standard
fact from linear algebra that the space of all linear functionals on X has the
same finite dimension as X itself. Since each of these linear functionals is
bounded, the spaces X and X* have the same dimension, so the dimension
of x(n) equals that of X for each positive integer n. Since the natural
map Q from X into X** is one-to-one, its range has the same dimension
as X, which implies that Q actually maps X onto X**.
Actually, Hahn called such spaces regular. The more descriptive term
reflexive was coined by Edgar R. Lorch [158] in 1939.
in Y* and each x" in X**. By Theorem 1.10.12, the maps T* and T**
are isomorphisms from Y* and X** onto X* and Y" respectively. Let Qx
and Qy be the natural maps from X and Y into X** and Y** respec-
tively. Fix an element y*' of Y** and let x be the element of X for which
T**Qxx = y**. If y* E Y*, then
(y*, y**) = (y*, T**Qxx) = (T*y*, Qxx) = (x, T*y*) = (Tx, y*),
whenever (a1, ... ,a r ,) E i'1 and (Pl, ... ,Pn) E i~. By Theorem 1.10.12,
the map T* given by the formula T*(x**) = x**T is an isometric isomor-
phism from (i'1)** onto (i~)*. Since i1 is reflexive, the natural map Q
from i1 into (il)** is onto (i1)**' so T*Q is an isometric isomorphism
from f onto (~)*. Whenever (aI, ... ,an) E i'1 and (PI, ... ,Pn) E i~,
((PI, ... , Pn), T*Q( aI, ... , an)) = (T(P1, ... , Pn), Q( aI, ... , an))
= ((a1, ... ,an ),T(P1, ... ,Pn)
n
as required.
As is suggested by the preceding proposition and Theorem 1.10.16, there
is a close relationship between the higher order duals of a subspace of
a normed space and the higher order annihilators of the subspace. See
Exercise 1.127.
When two normed spaces are isometrically isomorphic in some natural
way, it is common practice to treat the spaces as if they were the same space
with two different sets of labels for its elements, and then substitute one
of the spaces and its elements' labels for the other space and its elements'
labels in some expression or argument. Given this "substitution principle,"
Proposition 1.11.14 becomes rather obvious. When M is a subspace of a
normed space X, Theorems 1.10.17 and 1.10.16 provide natural ways to
identify M1..1 with (X* /M.1)* and X' /M1. with M*, so that M.1.1 can be
treated as the dual space of M* , with the action of the elements of M 1.1. on
those of M* given by the formula that shows how the elements of M 1.1. act
on those of X* / M.1. The proof of Proposition 1.11.14 amounts to nothing
more than combining the appropriate isometric isomorphisms in the correct
order to make this argument rigorous.
This substitution principle is often used without comment in the lit-
erature to shorten an argument by suppressing portions of the argument
1.11 The Second Dual and Reflexivity 103
M 1-1- is given its usual meaning as a subspace of X**. One of the most
important theorems about reflexive spaces follows almost immediately.
It follows that the natural map from Minto M** is onto M*', so M is
reflexive.
x - ("'Q
(x " ,x ***\1-- (Q xX,X ***) - (x,x **'Q) x **),
x,x
so x*" = Qx- (x'Qx). Since Qx- is onto X***, the space X* is reflexive.
Conversely, suppose that X* is reflexive. Then both X** and its closed
subspace Qx(X) are reflexive, so X is reflexive since it is isomorphic
to Qx(X).
1.11 The Second Dual and Reflexivity 105
1.11.20 Corollary. Suppose that XI, ... , Xn are normed spaces and that
X = Xl EEl EEl X n . Then X is reflexive if and only if each Xj is reflexive.
PROOF. It can clearly be assumed that n 2:: 2, and in fact that X =
Xl EEl X 2 ; an induction argument based on the fact that Xl ... Xn ~
(Xl EEl EEl X n - l ) EEl Xn when n 2:: 3 then gives the general case. Let M =
{(XI'O) : Xl E Xl}' a closed subspace of X isometrically isomorphic to Xl
by Proposition 1.8.3. The quotient space Xj M is isometrically isomorphic
to X 2 by Proposition 1.8.5, so it is enough to prove that X is reflexive if
and only if both M and X j M are reflexive. This follows from the preceding
theorem and corollaries.
1.11.24 Example. Suppose that (0,2:, f.L) is a positive measure space and
that 1 ::; p ::; 00. If Lp(O, 2:, j1) is infinite-dimensional, then Rp is isometri-
cally embedded in itj see Exercise 1.129. It follows immediately from the
nonreflexivity of Rl that L 1 (0,2:,j1) is reflexive if and only if it is finite-
dimensional, and similarly for L oo UJ,2:,f.L) because of the nonreflexivity
of Roo. In particular, the spaces LdO, 1] and Loo[O, 1] are not reflexive.
it follows that fO(Xl) = -1 and fo(xn) = +1 if n ::0: 2. But this cannot be,
since it would imply that there is a neighborhood of Xl on which the real
part of fo is negative and in which there are infinitely many points at which
fo takes on the value 1. Thus, if f E BC(K) , it must be that IJK f dill < 1,
so 11, viewed as a member of C(K)', is not norm-attaining. The space C(K)
is therefore not reflexive, and so the space rca(K) isomorphic to C(K)* is
also not reflexive.
This section ends with a word of caution about the definition of reflex-
ivity. Since the natural map from a normed space into its second dual is
1.11 The Second Dual and Reflexivity 107
Exercises
1.129 Suppose that (o,~, p,) is a positive measure space, that 1 ~ P ~ 00, and
that Lp(O,~, p,) is infinite-dimensional.
(a) Prove that there is a sequence of disjoint members of ~ each having
nonzero measure, and that if p ::/= 00, then the sets can be selected
so that each has finite measure.
(b) Prove that f.p is isometrically embedded in Lp(O,~, p,).
1.130 Complete the discussion begun in Example 1.11.25 by examining the re-
flexivity or nonreflexivity of C(K) and rca(K) when K is a compact
Hausdorff space with a finite number of points.
1.131 Let X and Y be Banach spaces. Prove that if X is reflexive, then X"", Y
if and only if X' "'" y'. Prove the corresponding statement for isometric
isomorphisms. Compare this to the result of Exercise 1.111.
1.132 Let X be a Banach space. Show that the kernels ofthe finite-rank bounded
linear operators with domain X are either all reflexive or all nonreflexive.
1.133 Prove or disprove: If X is a reflexive normed space and Y is a normed
space such that there is a bounded linear operator from X onto Y, then
the completion of Y must be reflexive.
1.134 Let X be a normed space and let statement 1.9.10' be as in Exercise 1.119.
Prove that statement 1.9.10* is true for X if and only if X is reflexive.
In particular, show that if X is not reflexive then there is a subset A
of X' and a corresponding collection of scalars satisfying (2) but not (1)
in statement 1.9.10'.
1.12 Separability
Recall that a topological space is separable if it has a countable dense
subset. In this section some of the special properties of separable normed
spaces are explored. The first order of business is to determine which of the
most commonly encountered normed spaces are separable. For this purpose,
the following result is useful.
1.12.1 Proposition.
rn=O
1.12.4 Example. The space Loo[O, 1] is not separable. To see this, suppose
1.12.5 Example. For each t in [0, 1], let Dt be the Borel measure on [0, 1]
such that Ot(A) = 1 if tEA and Ot(A) = otherwise. Then {Dt : t E [0,1] }
is an uncountable subset of rca[O, 1]. Since IIOtl - Dt211 = 2 if tl f= t2, the
space rca[O, 1] is not separable.
Most of the results about refiexivity in the preceding section have their
analogs for separability, and in fact more general results are often true.
The converse of the preceding theorem is false. For example, the space f.J
is separable, but its dual space is isometrically isomorphic to the nonsep-
arable space f.oo and so is not separable. The following, at least, can be
said.
1.12 Separability 113
Notice that the corollary gives new proofs of the nonreflexivity of the
spaces L1[0, 1], 1, and e[O,l], and therefore of Loo[O, 1J , Roc)) rca[O,l],
and co.
Suppose that M is a closed subspace of 1. Proposition 1.12.9 (e) guar-
antees that dM is separable. It may be somewhat surprising that, up to
isomorphism, such quotients of 1 are the only separable Banach spaces.
1.12.13 Lemma. (S. Banach and S. Mazur, 1933 [16]). Let X be a separa-
ble Banach space. Then there is a bounded linear operator from 1 onto X.
PROOF. Let {Xn : n EN} be a countable dense subset of B x. If (an) E 1,
then L:n anXn is absolutely convergent and therefore convergent, so the
formula T(a n ) = L: n anXn defines a map T from 1 into X. It is clear that
T is linear. Since IIT(an)11 ::; L:nlanl = lI(an)111 whenever (an) E 1, the
operator T is bounded.
All that remains is to show that T is onto X. Suppose that x E Ex.
It is enough to prove that x E T(l)' Select n1 so that Ilx - xn,ll < 1/2.
Select n2 so that n2 > n1 and 112(x - Xn,) - xn2 11 < 1/2, that is, so that
Ilx - x n , - 2- 1xn2 11 < r2.
Select n3 so that n3 > n2 and 1122(x - X n, - 2- 1 x n2 ) - xnJ < 1/2, that
is, so that
Ilx - Xn, - 2- 1x n - 2- 2xn3 11 < 2- 3 .
Continuing in the obvious way yields a subsequence (xnJ of (xn) such
that x = L: j 2 1 - j x n ]. Let (an) be the element of 1 obtained by letting
a nj = 2 1 - j for each j and letting an = 0 whenever there is no j such that
n = nj. Then T(a n ) = x, so x E T(l)'
1.12.14 Theorem. (S. Banach and S. Mazur, 1933 [16]). For every sep-
arable Banach space X, there is a closed subspace Mx of 1!1 such that
X~dMx.
This last set whose closed linear hull is being taken consists of the terms
and limit of a convergent sequence in X, and so is compact.
A normed space that is the closed linear hull of one of its compact subsets
is said to be compactly generated. The preceding theorem just says that a
normed space is compactly generated if and only if it is separable.
Exercises
1.135 Prove that every separable infinite-dimensional normed space has a lin-
early independent countable dense subset.
1.136 Prove that a normed space is separable if and only if its unit sphere is
separable.
1.137 Let M be a subspace of a normed space X. Prove that if M and M.l. are
both separable, then X is separable.
1.138 (a) Show that no subspace of Co is isomorphic to 1.
(b) Show that no quotient space of Co is isomorphic to 1.
1.139 Let X be the subspace of 1 consisting of the sequences in 1 that sum
to o. Is X separable? Explain.
1.140 Give an example of a one-to-one bounded linear operator from a nonsep-
arable normed space onto a separable normed space.
1.141 Here is a companion result for Lemma 1.12.13. Suppose that X is a Banach
space and that T is a bounded linear operator from X onto 1.
(a) Let (en) be the sequence of standard unit vectors in 1. Prove that
there is a bounded sequence (Xn) in X such that TXn = en for
each n.
(b) Conclude that X has a subspace isomorphic to e1
1.142 Let K be a compact Hausdorff space. Prove that rca(K} is separable if
and only if K is countable.
*1.13 Characterizations of Reflexivity 115
1.143 Suppose that (0, E, p,) is a positive measure space. Prove that Loo(O, E, p,)
is separable if and only if it is finite-dimensional.
1.144 Let X be a normed space and let Z be a separable subspace of X'.
Prove that there is a separable closed subspace Y of X such that Z is
isometrically isomorphic to a subspace of Y.
1.145 The evidence accumulated so far might lead one to conjecture that all
reflexive normed spaces are separable. Give a counterexample. (Consider
counting measures on uncountable sets.)
The portion of this section extending from here through Example 1.13.7
is devoted primarily to obtaining several closely related characterizations
of reflexivity, one of which is in terms of the following type of convergence.
It will be shown in Chapter 2 that every normed space has a topology 'I
such that a sequence in the space converges weakly to an element of the
space if and only if the sequence converges to that element with respect
to 'I. For the moment, the statement that a sequence converges weakly to
a certain limit should not be taken to imply anything more than is stated
in Definition 1.13.2. Notice that if a sequence (xn) in a normed space X
converges weakly to elements x and y of X, then x*x = limn xx n = x*y
for each X* in X*, and so x = y by Corollary 1.9.9. That is, no sequence
in a normed space has more than one "weak limit." Notice also that the
convergence of a sequence in a normed space in the usual sense implies
the weak convergence of that sequence to the same limit. Incidentally, the
converse is not true. The sequence (en) of standard unit vectors in 2 is
obviously weakly convergent to 0 and equally obviously not convergent
to anything with respect to the norm topology since the sequence is not
Cauchy.
The following lemma will be superseded by Theorems 1.13.5 and 1.13.6,
in which it will be shown that statements (a), (b), and (c) of this lemma
are actually equivalent.
1.13.3 Lemma. Let X be a normed space. Then (a) => (b) => (c) in the
following collection of statements.
(a) The space X is reflexive.
(b) Every bounded sequence in X has a weakly convergent subsequence.
(c) Whenever (en) is a sequence of nonempty closed bounded convex sets
in X such that Cn ;2 Cn+l for each n, it follows that nn C n =I- 0.
PROOF. Suppose that (b) holds. Let (en) be a sequence of nonempty closed
bounded convex subsets of X such that C 1 ;2 C 2 ;2 .... For each positive
integer n, let Xn be an element of en, and let x be the limit of a weakly
convergent subsequence (x nk ) of (xn). If x fI. em for some m, then by
Proposition 1.9.15 there is an x* in X* such that
This and the density of {y~ : n EN} in X* together imply that the se-
quence (x*xnk) is Cauchy and hence convergent whenever x E X*. The
map X* ~ limk x*xnk is clearly a linear functional on X', and this func-
s:
tional is in X** since Ilimk xx nk I IIx' II sup{ Ilx n II : n EN} for each x*
in X'. The reflexivity of X yields an x in X such that limk x*xnk = x*x
whenever x* E X*, so (x nk ) is a weakly convergent subsequence of (xn).
Therefore (a) =} (b) when X is separable, which finishes the proof.
Most of the theorems of this section are ultimately derived from the
following result.
Let Cn = (} and let Cl = ... = Cn-l = O. Then for each linear combination
alQx1 + .. : + a n -1Qx n -1 + anx'* of QXl,'" ,QXn-l, x**,
tajel =
I
J=l
ItajX*'x;1 <:::
J=l
IIX**lllltajx;ll,
1'=1
It is a corollary of Theorem 1.13.4 and its proof that (c) => (a) in
Lemma 1.13.3. To see this, suppose that X is a normed space that sat-
isfies part (c) of that lemma; that is, for which nnCn =1= 0 whenever
(Cn ) is a sequence of non empty closed bounded convex subsets of X such
that Cn 2 C n + 1 for each n. The goal is to prove that X is reflexive, so
by Proposition 1.13.1 it may be assumed that IF = JR. as in the proof
of Theorem 1.13.4. Let (Yn) be a Cauchy sequence in X and let Dn =
co{ Yj : j ::::: n} for each n. Then for each positive E there is a positive inte-
ger n, such that Dn lies in a closed ball of radius E when n ::::: n,. It follows
that nn Dn has exactly one element Yo and that Yn -- Yo. The space X is
therefore a Banach space. If X were not reflexive, then part (c) of Theo-
rem 1.13.4 would hold, and a peek at the first paragraph of the proof of
that theorem shows that X would have a sequence (Cn ) of nonempty closed
bounded convex subsets such that Cn 2 Cn+l for each nand nnC n = 0.
This contradiction proves that X is reflexive.
Since (c) => (a) in Lemma 1.13.3, statements (a), (b), and (c) in that
lemma are actually equivalent. This is the content of the next two theorems.
1.13.7 Example. Let llJ) be the closed unit disc in the complex plane and
let A(llJ)) be the disc algebra, that is, the subspace of G(llJ)) consisting of the
members of C(llJ)) analytic in the open unit disc. Since the uniform limit
of a sequence of functions analytic in the open unit disc is analytic in that
disc, it follows that A(llJ)) is a closed subspace of G(IID) and so is a Banach
space.
For each positive integer n, define Xn in A(IID) by the formula xn(z) = zn,
and let Gn = co({X n ,X n +1, ... }). If x is in CO({X1,X2, ... }), then IIxll oo = 1
since x(l) = 1 and Ix(z)1 :S 1 for each z in llJ). It follows that Gn S;; SA(HlI)
for each n.
If n E N and x E co( {x n , Xn+l' ... } ), then x and its first n - 1 derivatives
all have value 0 when z = O. Now if (Yj) is a sequence in A(llJ)) and Y is an
element of A(IID) such that IIYj - Ylloo --> 0, then a standard result about
the uniform convergence of analytic functions assures that yj(O) -- y'(O),
yj'(O) -+ yl/(O), yj"(O) __ y'I/(O), and so forth. It follows that if n E Nand
x E Cn, then x and its first n - 1 derivatives all have value 0 when z = O.
Suppose that x E nn Gn . Then x and its derivatives of all orders have
value 0 when z = 0, so a moment's thought about the power series ex-
pansion for x about 0 shows that x = 0, even though Ilxll= = 1. This
contradiction shows that nn C n = 0, so A(IID) is not reflexive.
120 1. Basic Concepts
This shows that (a) ==} (b). It is clear that (b) ==} (c).
Finally, suppose that 8 and (xn) are as in (c). For each positive integer n,
let en= co( {Xn+l' Xn+2""})' a subset of Bx Suppose that X E nn en.
Then there is an m in N and a y in co( {Xl"'" xm}) such that Ilx-yll < 8j2,
as well as a z in cO({X m +l,X m +2, ... }) such that Ilx - zll < ej2, which
implies that e s: s: e.
Iiy - zll Ily - xII + Ilx - zll < This contradiction shows
that nn en = 0, so X is not reflexive. This proves that (c) ==} (a).
*1.13 Characterizations of Reflexivity 121
The preceding theorem could have been stated with the convex hulls
replaced by closed convex hulls, since the distance between two nonempty
subsets of a metric space is the same as the distance between their closures.
Suppose that X is a Banach space. By Proposition 1.11.11, if X is reflex-
ive then each member of X* is norm-attaining. The converse is also true
and is known as James's theorem. This result has a long and interesting
history, most of it associated with the name of Robert C. James.
Stanislaw Mazur [163J was the first to ask if a Banach space must be
reflexive when all members of its dual space are norm-attaining. Though
Mazur's paper appeared in 1933, the first substantial progress toward an-
swering his question was not made until 1950, when James [108J showed
that if a separable Banach space X has a Schauder basis, a certain type of
sequence that will be discussed in Chapter 4, then X is reflexive if each Ba-
nach space Y isomorphic to X has the property that each element of Y is
norm-attaining. In the same year Victor Klee [132J used an argument based
on Theorem 1.13.6 to improve James's result by removing the requirements
that X have a Schauder basis and be separable. Incidentally, Klee's paper
contains a number of interesting characterizations of reflexivity that will
not be covered here. Most of the paper is accessible after reading Chapters
1 and 2 of this book.
In 1957, James [111J showed that a separable Banach space is reflexive if
each member of its dual space is norm-attaining. Though the same result
for nonseparable Banach spaces would not appear for another seven years,
the impact of James's 1957 result on Banach space theory was immediate
and substantial. It is often possible to prove from this result that a Banach
space is reflexive by showing that all of its separable closed subs paces are
reflcxivc. For example, it can be shown in this way that a Banach space is
reflexive whenever each of its nonempty closed convex subsets has a point
nearest the origin; see Exercise 1.153. It is not at all a coincidence that the
branch of approximation theory called Banach space nearest point theory,
dealing with points of sets in Banach spaces nearest other sets or points,
began to grow rapidly at about this time. The interested reader might want
to look at Section 4 of the 1958 paper by Ky Fan and Irving Glicksberg [77J
on spheres in normed spaces for another good example of an argument in
which James's 1957 result is used to prove that a possibly nonseparable
Banach space is reflexive.
James [112J finally completed his quest by showing in a 1964 paper that
the separability hypothesis in his 1957 result is unnecessary. In fact, in an-
other 1964 paper James [115J proved a stronger result called James's weak
compactness theorem, or often just James's theorem since the reflexivity
theorem that also goes by that name turns out to be a special case of the
weak compactness theorem. It would not be easy to state James's weak
compactness theorem in its full generality here, since the language needed
to do so is not developed until Chapter 2. However, it is already possible to
state the following version of it. Though the class of sets in the hypotheses
122 1. Basic Concepts
is smaller than in the general version, this version is still strong enough to
imply the reflexivity theorem. See Exercise 1.154.
Let C be a closed convex subset of a Banach space X. Suppose that
whenever x* E X*, the supremum of Ix* I on C is actually attained
by Ix*1 somewhere on C. Then every sequence in C has a weakly
convergent subsequence.
The general form of James's weak compactness theorem will be derived in
Section 2.9.
James's 1964 proof of his reflexivity theorem is based on Theorem 1.13.4
and is somewhat intricate. The proof given here is a greatly simplified one,
also by James [117], that appeared in 1972. The plan of attack is to prove
James's theorem for separable Banach spaces, then obtain the general case
from that. The following technical lemma is needed for the separable case.
This lemma is stated and proved in a bit more generality than is actually
needed in this section, for the only immediate application will be to the
case in which the set A mentioned in the lemma is the closed unit ball of
the space. However, the more general result is no more difficult to prove,
and will have an important application in Section 2.9.
f DO {3k E k 00
and
where
It !'3jY;IA
= I2::~:!'3j (;!'3jY; + (~!'3j)Y~) + 2::~:tj ;!'3jy;IA
<
Let Cj = 0 when j = 1, ... , n and let Cn+1 = O. Then for each linear
combination ll:lQXI + ... + ll:nQxn + Qn+lX'* of QXI,"" QXn' x'*,
(iv) Ilv;II::;M+E;
(v) (QXj)(V;) = Cj = 0 when j = 1, ... ,n; and
(vi) x"y; = Cn+l = O.
Letting x~ = Y; for a suitably small E yields an x~ satisfying (i), (ii),
and (iii).
If x* E cor { x~ : n EN}), then the validity of (ii) for each n implies that
Ilx' II 2: Ix**x* I = 0, and so
d(O,co({X~: n EN})) 2: O.
Iz*xl = If
1=1
{Jjy;XI
S; I~ {JjY;XI ~f-~jIY;XI
< lit 1=1
(JjY; II + a() f
J=n+1
{Jj
=a
= Ilz*lI,
so z* does not attain its norm at x. This proves that (c) => (d).
Finally, Proposition 1.11.11 assures that each bounded linear functional
on a reflexive normed space is norm-attaining, so (d) => (a).
and
This notation is needed only for the next few results and does not ap-
ply outside this section, with the only exception being that the nota-
tion for L(x~) will be temporarily reinstated in Section 2.9. Notice that
(x~) E V(x~), that each member of V(x~) is a sequence in the closed ball
centered at 0 of radius sup{ Ilx~ II : n EN}, and that V(y~) ~ V(x~) and
L(y~) ~ L(x~) whenever (y~) E V(x~). Notice also that if x' E L(x~),
then Ilx*11 S; sup{ Ilx~11 : n EN} since IX'xl S; sup{ Ilx~11 : n EN }llxll for
each x in X, and lim infn x~x ::; x'x whenever x E X since lim infn x~x =
-limsuPn x~( -x).
1.13.12 Lemma. Let X be a real normed space and let (x~) be a bounded
sequence in X*. Then L(x~) is nonempty.
PROOF. Let p(x) = limsuPn x;,x for each x in X. Then p is a sublinear
functional on X. Let Y' be the zero functional on the subspace {O} of X.
128 1. Basic Concepts
Then y*(O) = p(O), so the vector space version of the Hahn-Banach ex-
tension theorem implies that y* has a linear extension x* to X such that
x*x ::::: p(x) whenever x E X. Since Ix*xl : : : sup{ Ilx~1I : n E N}llxll when-
ever x EX, the linear functional x* is bounded, and so x* E L( x~).
The following technical lemma is the heart of the proof of James's the-
orem for arbitrary Banach spaces. Notice its similarity to Lemma 1.13.10.
As with Lemma 1.13.10, it is proved in a bit more generality than is needed
in this section, but the extra generality will be required in Section 2.9.
defines a nonempty subset of [a, 2], so the following formula defines a num-
ber in [a, 2]:
am-l <.
_ m f{ sup S rn-l (*
y, (Vj*)). Y* -_ {3m-l
,",00
*
,Ym-l
L..Jj=m-l (3)
2:r;m (3j
+ ,",00
L..Jj=rn-l (3)
,X ,
*
as claimed.
Claim 6: e ~ am ~ 2. To prove this, notice that because of Claims
2 and 5 it is enough to show that al :::: e. Let y* be an clement of
co({x; : j EN}) and (v;) an element of Vex;). It is enough to show
that SUPSl(Y*, (v})) :::: e, that is, that sup{ IY* - w*IA : w* E L(v})} :::: e.
Let w* be an element of L(vi). It is enough to show that IY* - w*IA :::: e.
But this is true by the hypotheses of this lemma, since L( vil c::: L( x;),
which proves Claim 6.
Using the definition of am, choose Y;" from co({m-lx:n, m-lx;"+l, ... })
and (mzil from V(m-lx,i) so that
am I
~ su p { ~l {3jyj + (~{3j )Y;;" - w*L: w* E L(Tnzj) } (1.6)
< am(l + Em),
then choose w;" from L( mzil so that
(1.7)
*1.13 Characterizations of Reflexivity 131
It follows from Claim 1 that llim infj(mzixm) I :::; 1. Let (mxi) be a subse-
quence of (mzi) such that limj(mxixm) = liminfj(mzixm)' It then follows
from the fact that (m-1xi) lies in B x " along with Claims 1, 5, and 6, that
am, y;", (mzi), and (mxi) do what is required of them in (1) through (6)
when n = m. This completes the induction.
Claim 7: L(y;) c::::: n~=o L(nxi) c::::: n~=1 L(nzi). To see that the second
inclusion holds, suppose that n E N. It is enough to show that L(nxi) c:::::
L(nzj), for which it is enough to show that (nxi) E v(nzi), and this
follows from the fact that (nxi) is a subsequence of (nzi). For the proof
that L(yi) c::::: n~=o L(nxi), suppose that n is a nonnegative integer. Notice
that if j EN, then (2), (3), and (4) together imply that
Yj* E co ({J-l'
Xj'
j-l Xj+l, . . . }) C
_ co ( { j - lZj'
'
J-l Zj+l,' .. })
C
_co ({j-2'
Xj' j-2 Xj+l,'" }) C
_co ({j-2'
Zj' j-2Zj+l,'" })
C
c::::: co( {Ox;, Xi+l' ... }).
Therefore Yi E co( {nxi, nX;+I' ... }) when j > n, from which it follows
that lim SUPj (y}x) :::; limsuPj(nxjx) for each x in X and thus that L(y}) c:::::
L(nx}). This finishes the proof of Claim 7.
Claim 8: If w* E L(y}) and mEN, then (1.7) holds for the same
element X Tn of A when w;" is replaced by W. To prove this, notice from
Claim 7 that w E L(mxi), and so the way that (mxi) was obtained from
(mz;) assures that
Claims 5 and 6 together assure that limn an exists and lies in [B,2]. Call
this limit a. Taking limits as n tends to infinity in (1.8) shows that
1.13.14 Theorem. (R. C. James, 1964 [112], 1972 [117]). Let X be a real
Banach space. Then the following are equivalent.
(a) The space X is not reflexive.
(b) If 0 < 8 < 1, then there is a closed subspace M of X and a se-
quence (x~) in Bx< such that d(M.1.,co({x~ : n EN})) 2:: 8 and
limn x~x = 0 for each x in M.
(c) If 0 < 8 < 1 and ({3n) is a sequence of positive numbers with sum 1,
then there is an a such that 8 :::; a :::; 2 and a sequence (y~) in Bx<
such that whenever w* E L(y~),
(1) 11~~1 {3J(yj - w*)11 = a; and
(2) 11~;'=1 {3J(yj - w*)11 < a(1 - 8 ~~n+l {3j) for each positive in-
teger n.
(d) There is a z* in X* that is not a norm-attaining functional.
PROOF. For the proof that (a) =} (b), suppose that X is not reflexive
and that 0 < 8 < 1. It follows from Theorem 1.13.8 that some separable
closed subspace M of X is not reflexive. By Theorem 1.13.11, there is
a sequence (m~) in B M , such that d(O,co({m~ : n EN})) 2:: 8 and
limn m~x = 0 for each x in M. For each positive integer n, let x~ be a Hahn-
Banach extension of m~ to X. If x E co( {x~ : n EN}) and y' E M.1.,
then the restriction of x' - y* to M is a member m* of co( { m~ : n EN}),
and so
~ Ilm'll
2:: d (0, co ( { m~ : n EN} ) )
2:: 8.
It follows that
Then (3n) is a sequence of positive scalars that sums to 1. Let a and (y~)
be as in (c). Let w* be any member of L(y~) and let z* = 2:;:1 (3j(yJ -w*).
Then Ilz11 = a. It will be shown that z* is not norm-attaining. Suppose
that x E Bx. Since liminfj yjx :<::; w*x and (;I :<::; a, there is a positive
integer n such that
Since w*y :<::; limsuPj yjy :<::; 1 whenever y E B x , it follows that Ilw*11 :<::; 1.
Therefore
00
j=1
+ I: (3j(yJ
00
< a( f 1- 0 (3j)
j=n+1
+ (aO - 2D.)(3nH +2 f (3j.
j=n+2
I:
00
I:(3j
00
= a - (a(;l- 2D.)
j=n+2
<a
= Ilz*ll
Since -x E Bx, it also follows that -z'x = z*( -x) < Ilz'lI, so Iz*xl < IIz*lI.
Therefore z' is not norm-attaining, which proves that (c) => (d).
Finally, Proposition 1.11.11 assures that each bounded linear functional
on a reflexive normed space is norm-attaining, so (d) => (a).
134 1. Basic Concepts
The big theorem of this section is now an easy corollary of the result just
proved.
(i) The following does not hold: For each (} such that 0 < (} < 1 there
is a sequence (x~) in Sx' and a sequence (xn) in Sx such that
Rex~xj :::: (} if n ~ j and Rex~xj =0 if n > j.
(j) The following does not hold; For some () such that 0 < () < 1 there is
a sequence (x~) in Sx' and a sequence (xn) in Sx as in (i).
(k) The following does not hold: For each () such that 0 < e < 1 there is
a sequence (xn) in Sx such that
for each n.
(1) The following does not hold: For some () such that 0 < () < 1 there is
a sequence (xn) in Sx as in (k).
Another proof of the equivalence of (a), (c), and (e) will be obtained in
Section 2.8 from the Eberlein-Smulian theorem.
Exercises
1.146 Use an argument based on Proposition 1.13.1 to show that the nonreflex-
ivity of complex Co follows from that of real Co. (This is a trick question.
To avoid the trap, think carefully about what is obtained from complex Co
by restricting multiplication of vectors by scalars to IR x co.)
1.148 Find a sequence (Cn ) of nonempty closed convex subsets of 8f= such that
C n "2 C n + 1 for each nand nn
C n = 0.
1.149 In theory, Theorem 1.13.6 gives a test for the reflexivity of a normed
space requiring no knowledge of its dual. In practice, it often happens
that information obtained from the dual space is used to show that The-
orem 1.13.6 can be applied. This is true in the proof of the nonreflexivity
of A(lIli) given in Example 1.13.7. In that proof, what information comes
from A(lIli)*?
1.150 Suppose that X is an incomplete normed space such that every element
of X' is norm-attaining. Prove that the completion of X is reflexive.
1.152 Prove that a Banach space is reflexive if and only if it has this property:
Whenever A and Bare nonempty closed convex subsets of the space such
that d(A, B) = 0 and at least one of the two sets is bounded, the two sets
intersect. Exercise 1.151 might help.
136 1. Basic Concepts
1.153 Let X be a Banach space. Prove that the following are equivalent.
(a) The space X is reflexive.
(b) Whenever C is a non empty closed convex subset of X, there is a
point of C nearest the origin; that is, there is at least one x in C
such that Ilxll = d(O, C).
(c) Whenever Y is a closed separable subspace of X and y* E Y*, there
is a point of {y : y E Y, y*y = Ily*11 } nearest the origin.
If James's theorem is necessary in your argument, use only the version
for separable spaces. If it is not, please send a copy of your proof to the
author of this book!
1.154 Show that James's theorem follows from the form of James's weak COIIl-
pactness theorem stated in the discussion preceding Lemma 1.13.10.
1.155 Give a single example that shows that the conclusions of Theorems 1.13.4
and 1.13.9 do not necessarily hold if the normed space X in the statements
of those theorems is not required to be complete.
1.156 This exercise uses the result of Exercise 1.151. Theorem 1.13.8 might
lead one to wonder if a normed space must be separable whenever each
of its reflexive subspaces is separable. The purpose of this exercise is to
disprove this conjecture by showing that each reflexive subspace of Roo
is separable. Let X be a reflexive subspace of Roo. Justify each of the
following statements.
(a) Let II(Xj)lla = L::) TJlxjl whenever (x)) EX. Then 11lla is a norm
on X. Furthermore, if (Xj) E X and (x;n))~=l is a sequence in X
bounded under the 00 norm, then limnll(xjn)) ~ (Xj)lla = 0 if and
only if limn x;n) = Xj for each j.
(b) The set Ex (meaning E(x, 11.1100)' not E(x, 11'lIa)) is a compact, hence
separable, metric space under the metric induced by 11lla' Let D be
a countable subset of Ex dense in Ex with respect to this metric.
For the rest of this exercise, the normed space language and symbols refer
to the space (X,IIlloo), not (X,IIlla).
(c) For each x in Ex there is a sequence from D converging weakly to x.
(d) It follows that [D] = X, so X is separable.
2
The Weak and Weak* Topologies
2.1.2 Definition. Let (X, 'I) be a topological space and let x be an element
of X. A (local) basis for 'I at x is a collection ~x of open sets containing x
such that every open set containing x includes a member of ~x.
2.1.4 Definition. Let X be a set and let 6 (Fraktur S; notice its resem-
blance to u) be a collection of subsets of X. Let ~6 be the collection of all
sets that are intersections of finitely many members of 6. Then the topology
generated by the subbasis 6 is the topology generated by the basis ~6.
The following properties of bases and subbases for topologies follow di-
rectly from the above definitions. Every basis for a topology is also a sub-
basis for that topology. Every member of a basis or subbasis for a topology
belongs to that topology. If 'I' is a topology for a set X, then a collection 23
of subsets of X is a basis for 'I' if and only if every member of 23 is open
and every open set is a union of members of 23, which happens if and only
if the following holds: For each x in X, the family {B : B E 23, x E B}
is a basis for 'I' at x. If 6 is a subbasis for a topology 'I'6 for a set X and
'I' is another topology for X such that 6 ~ 'I', then 'I'6 ~ 'I'; that is, the
topology 'I'6 is the smallest topology for X that includes 6.
Notice that the basis generated by the sub basis in the preceding defini-
tion consists of all sets of the form TIa.EI Ua., where each U'" is open and
{ a : a E I, Un =1= X", } is finite. Henceforth, when the Cartesian product of
topological spaces is treated as a topological space without the topology being
specified, the product topology is implied.
It is worth noting what happens when the index set I in the preceding
definition is empty. In that case, the Cartesian product TInEI X", has as
its lone element the empty set, viewed as a function from I into U"'EI X",;
see, for example, [65, p. 22]. If 6 and 23 are, respectively, the subbasis and
basis for the product topology of TIo:EI X", discussed above, then it is easy
to check that 6 = 23 = {TI"'E! X",} = {{0}}.
(c) The space X is a To space if, for each pair of distinct points in X, at
least one has a neighborhood not containing the other.
(d) The space X is a T 1 space if, for each pair of distinct points in X,
each has a neighborhood not containing the other.
(e) The space X is a Hausdorff or separated or T2 space if, for each pair
of distinct points x and y in X, there are disjoint neighborhoods Ux
and Uy of x and y respectively.
140 2. The Weak and Weak* Topologies
IThe sharp-eyed reader will notice that continuity is mentioned here even though its
definition does not appear until later. It is assumed that the reader is already familiar
with continuity in topological spaces, as well as with most of the other terms cataloged
in the early part of this section for reference.
2.1 Topology and Nets 141
The conditions defined in (c) through (h) are sometimes called separation
axioms. It is not hard to show that a topological space is a T} space if
and only if each of its one-element subsets is closed; see Exercise 2.2 (a).
From that and an application of Urysohn's lemma, it follows easily that
T4 =} T31 =} T3 =} T2 =} T} =} To.
2
Here arc some relationships between the various types of compactness
mentioned above. Most general topology texts, such as [65J or [172J, will
have proofs of the less obvious ones. In a metric space, the properties of
compactness, countable compactness, limit point compactness, and sequen-
tial compactness are equivalent, as are the corresponding relative prop-
erties. Compactness obviously implies countable compactness. Countable
compactness implies limit point compactness; the properties are equivalent
in Hausdorff spaces. Even in Hausdorff spaces it is not always true that a set
is relatively sequentially compact exactly when its closure is sequentially
compact, or that a set is relatively limit point compact exactly when its
closure is limit point compact; see Exercise 2.15. However, the equivalences
do hold in metric spaces; see Exercise 2.l.
2Since two quantities equal to the same quantity should be equal to each other, it is
best to avoid the notation limn Xn = X when (xn) might have more than one limit. A
similar comment applies to the notation for net convergence that will be introduced in
Definition 2.1.14. This problem does not arise in Hausdorff spaces; see Proposition 2.1.17.
142 2. The Weak and Weak* Topologies
(b) continuous if, for each open subset V of Y, the set f- 1 (V) is open;
(e) open if, for each open subset U of X, the set feU) is open;
The theory of nets can be developed either with or without this additional
axiom. See [172, pp. 187-188] for a development that uses it.
The reason that the term Moore-Smith sequence is sometimes used for
a net is that E. H. Moore and H. L. Smith [171] introduced nets in 1922
as the basis for a general theory of limits. Mauro Picone [190] devised the
same theory independently in a book that appeared the next year. The
term net was actually first used by J. L. Kelley [131] in a 1950 paper on
topological convergence. 3 More on the early history of nets and the general
theory of limits can be found in the survey articles by E. J. McShane [167]
and R G. Bartle [18].
If f: I -+ X is a net, then for each Q in I the Qth term f(a) of the net.
is often denoted by X a , and the entire net is often denoted by (Xa)aEI or
just (xa). By analogy with sequences, it is said that Xa precedes x{3 in a
net when Q ::5 {J. In general, the familiar language of sequences is extended
to nets whenever the meaning is clear.
2.1.9 Example. Every sequence is a net, with the directed set being N in
its natural order.
2.1.10 Example. The set lR with its natural order is a directed set, so this
order makes every function with domain lR into a net. Notice that a term
in a net can be preceded by infinitely many others and that nets need not
have first terms.
3The terminology was not Kelley's invention, though. Kelley had wanted to call
such an object a way. However, nets have subnets, which Kelley would have dubbed
subways. Norman Steenrod talked him out of it. After some prodding by Kelley, Steenrod
suggested the term net as a substitute for way. See [204].
144 2. The Weak and Weak* Topologies
2.1.11 Example. The set JR2 can be made into a directed set by declaring
that (O'l,!'h) j (0'2,jJ2) whenever 0'1 :::; 0'2. If X(a,j3) = O'+jJ for each (0',jJ)
in JR2, then (X(a,j3)) is a net in lR. Notice that (1,2) j (1,3) and (1,3) j
(1,2) even though (1,2) i- (1,3).
illustrates several important ways in which nets can differ from sequences.
(a) The index set for a net can be finite.
(b) Nets can have last terms.
(c) Nets can have multiple "first" terms, that is, terms not preceded by
other terms.
(d) The index set for a net need not be a chain.
xo: ---+ x if and only if the following is true: For every member U of 6 that
contains x, there is an au in I such that Xo: E U whenever au :::5 a.
PROOF. The forward implication follows immediately from the definition
of net convergence. For the converse, suppose that every member U of 6
that contains x satisfies the stated condition. It follows from conditions (2)
and (3) in Definition 2.1.8 that every finite subset of I has an upper bound
in I, which implies that if.j is a finite subset of 6 each of whose members
contains x, then there is an a;; in I such that Xo: E n{ U : U E .j} whenever
a;; ::S a. Therefore Xo: ---+ x because of the way that 6 determines a basis
for the topology of X.
Since a point x is a limit point of a set S if and only if xES \ {x}, this
next corollary follows immediately from the proposition.
2.1.22 Corollary. If two topologies on the same set result in the same
convergent nets with the same limits for those nets, then the two topologies
are the same.
PROOF. Under the hypotheses of the corollary, the identity map on the
space, treated as a map between the two topological spaces in question, is
continuous in each direction and so is a homeomorphism.
1 3 5 7 9
2.1.27 Example. Let (xn) be any sequence and let J be the natural num-
bers with their usual order except that each relation between two integers
that requires an integer to be properly less than an odd integer has been
discarded. The directed set J is represented visually in Figure 2.1, where
Q -> f3 means that Q ~ (3. If g is the "identity" map from J into the natu-
ral numbers with their usual order, then (Xg(n) is a subnet of (xn). Thus,
subnets of sequences do not have to be sequences, or even have chains for
index sets.
2.1.28 Example. Suppose that (xn) is any sequence. Let the subset [1, (0)
ofR have its natural order and let g: [1, (0) -> N be the function that maps
each r in [1, (0) to the greatest integer less than or equal to r. Then (x r )
(that is, (Xg(r))) is a sub net of (xn). While a subsequence is often thought of
as being obtained from a sequence by thinning the index set, this example
shows that a subnet can be formed from a net by thickening the index set
(though of course the range of a subnet is never any larger than that of the
original net). Incidentally, notice that in this particular case the net (xn)
can also be viewed as a subnet of (xr) in an obvious way.
2.1.29 Example. Let (x",) be the net of Example 2.1.25. Let J be a one-
element set {(3} made into a directed set in the only way possible, and
define g: J -> No by letting g(3) = O. Then (Xg(,8) is a subnet of (x",).
More generally, suppose that J is a directed set and g: J -> No is a function
to be used to construct a subnet of (x",). Then the range of 9 must contain 0
but does not have to contain anything else.
150 2. The Weak and Weak* Topologies
Here is a collection of facts about subnets and the ways in which their
convergence is related to that of the main net.
2.1.32 Technique. Suppose that (xa) and (YfJ) are nets with respective
index sets I and J. It is often useful to be able to find subnets (xl') and (Yl')
of (xa) and (YfJ) respectively that have the same index set K. To do this, let
K = I x J, directed by declaring that (a 1,(3d ::S (a2' (32) when a 1 ::S a2 and
(31 ::::: (32. Let g: K -+ I and h: K -+ J be the projection mappings, that is,
the mappings defined by the formulas g(a, (3) = a and h(a, (3) = (3. Then
(xg(a,fJ and (Yh(a,fJ are subnets of (xa) and (YfJ) respectively having
the same index set. Notice that these subnets are, in a sense, formed by
thickening the index sets of the corresponding nets. Notice also that if (xa)
lies in a topological space, then (xg(a,fJl) converges to some x if and only if
(xa) converges to x, and similarly for (Yh(a,fJl) and (YfJ)
One more notion for sequences needs to be extended to nets before pro-
ceeding. Recall that an accumulation point of a sequence in a topological
2.1 Topology and Nets 151
Two of the most basic facts about net accumulation are contained in the
next result. The proof follows very easily from the relevant definitions.
The following result is the analog for arbitrary topological spaces of the
equivalence of compactness and sequential compactness in metric spaces.
as required.
See Proposition D.lO in Appendix D for a characterization of relative
compactness in topological groups in terms of the convergence of ultranets.
2.1 Topology and Nets 155
logical group has its topology induced by a metric, for it is quite possible
for a net in such a space to be metrically Cauchy or topologically Cauchy
without being Cauchy in both senses; see Exercise 2.17. However, this dif-
ficulty does not arise in the situations of most importance in this book, as
will be shown by Proposition 2.1.44 and its corollary.
and
Since every normed space is an abelian topological group when the group
operation is vector addition, and since the metric induced by the norm is in-
variant for this operation, the following result is an immediate consequence
of the proposition just proved.
so (xc:.) is Cauchy.
2.1.48 Definition. An abelian topological group is complete if each Cau-
chy net in the group converges.
158 2. The Weak and Weak* Topologies
Exercises
Here are some examples of vector topologies that are not in general
induced by norms.
JL[ lanfn - afl > f] ::::: JL [Ianln - anfl > ~] + JL [Ianl - all > ~]
::::: JL [ Blfn - II > ~] + f-t [bnlfl > ~]
= JL [Ifn - ]
II > 2~ + f-t [If I > 2: ].n
Letting n tend to 00 shows that f-t[ lanln - ail> f 1 -> 0 whenever e > 0,
so anln converges to af in measure. (Notice that the finiteness of JL(n) is
used here to assure that JL[ If I > e/2b n ] -> 0.) Multiplication of vectors by
scalars is therefore a continuous operation, so the topology of convergence
in measure is a vector topology for Lo(n, I:; , f-t) whenever (0" r:, JL) is a finite
positive measure space.
Now suppose that 0, is the interval [0,1]' that I:; is the u-algebra of
Lebesgue-measurable subsets of 0" and that A is Lebesgue measure on
164 2. The Weak and Weak* Topologies
(n, ~). Let Lo[O,l] denote Lo(n,~, .\). Let C be a convex neighborhood
of in Lo[O, 1] and let 10 be such that the open ball of radius 10 centered
at lies in C. Suppose that f E Lo[O, 1]. Let n be a positive integer such
that l/n < 10/2. For each integer j such that 1 ~j ~ n, let I[.t..;;-!.*) be the
indicator function of the interval [~, *)
and let the member h of La [0, 1]
be defined by the formula fj(t) = nf(t)I[.t..;;-!.-!;)(t). Then A[ Ihl > E/2] ~
l/n < 10/2 for each fj, so each h is in C. Since f = 'L.j=l n- 1h, the convex
combination f of members of C is itself in C. It follows that C = Lo[O, 1].
By Proposition 2.1.39 (b), every nonempty convex open set in Lo[O, 1] is a
translate of a convex neighborhood of the origin, so it also follows that the
only nonempty convex open subset of Lo[O, 1] is Lo[O, 1] itself. Thus, the
topology of convergence in measure for Lo[O, 1] is a vector topology that is
not locally convex.
2.2.6 Example: Lp(n,~, '""),
< p < 1. Suppose that'"" is a positive
measure on a O"-algebra ~ of subsets of a set n and that < p < 1. Define
Lp(n,~, '"") to be the collection of all ,",,-measurable scalar-valued functions
on n such that 10
Ifl P d,"" is finite, with the usual convention that functions
that agree almost everywhere are considered to be the same. If s ::::: and
s(t) = sP + t P - (s + t)P whenever t ::::: 0, then s is nondecreasing and
hence nonnegative on [0,(0), from which it follows that
,",,[If-g'>c 1 =1 [I/-gl><]
l
1d,",,:St- p
[1/-gl><1
If-gIPd,",,~cPdU,g),
from which it follows that every Cauchy sequence in Lp(rl, L;, '"") is Cauchy
in measure and so is convergent in measure to some measurable function.
Let Un) be a Cauchy sequence in Lp(o',"2'" J-l) and let f be the function
to which (fn) converges in measure. It is a standard fact from measure
theory that there is a subsequence Unj) of Un) that converges to f almost
everywhere. Suppose that 10 > O. Let the positive integer N. be such that
2.2 Vector Topologies 165
It follows that I = (f - INJ + IN, E Lp(D., E, fl) and that d(f, In) -+ 0,
which establishes the completeness of d.
The metric d is invariant, so vector addition is continuous for Lp(D., E, fl)
since the space is a topological group under this operation. Furthermore,
if (an) and (fn) are sequences in IE" and Lp(D.,E,fl) respectively and have
the respective limits a and I, then
for each integer j such that 1 ::; j ::; n. For each such j, let I[tj_l,t J ) be the
indicator function of the interval [tj-l, tj) and let the member fJ of Lp[O, 1]
be defined by the formula /jet) = nl(t)I[tJ_l,tj)(t). Then
d(fJ, 0) = n P1 .I/IP
[tj-l,t)
dA = nP-ll
[0,1]
I/I PdA = np-1d(f, 0) < E
2.2.7 Example: i p , < p < 1. Let p be a rcal number such that < p < 1
and let J.l be the counting measure on the collection E of all subsets of N.
166 2. The Weak and Weak* Topologies
By analogy with what is done when 1 :::; p :::; 00, the space fp is defined to
be Lp(N,~,f1,). Notice that fp is the space of all sequences (an) such that
2:: n la n lP is finite, and that if (an), (On) E fPl then
operations of X implies that the maps x f-+ x + Xo and x f-+ aox are contin-
uous, as are their respective inverses x f-+ x - Xo and x f-+ ao1x, so the two
maps of part (c) are homeomorphisms from X onto itself. If A is a subset
of X that is open, or closed, or compact, then Xo + A and aoA have that
same property since these properties are preserved by homeomorphisms.
If A and U are subsets of X and U is open, then A + U is open as the union
of open sets, since A + U = U{ a + U : a E A}. This proves (c).
Suppose that A, B, Xo, and ao are as in the hypotheses of (d). Let Yo
and Zo be elements of A and B respectively, and let (y"J and (zj3) be nets
in A and B respectively such that y" ....... Yo and zj3 -> Zo0 There are subnets
(Yi) and (zi) of (Yo,) and (zj3) respectively having the same index set; see
Technique 2.1.32. Then Y"I + z"I ....... Yo + Zo, so Yo + Zo E A + B. This proves
that A + B <;;: A + B. Since A + BO is an open subset of A + B, it follows
that AO + BO <;;: (A + B)o. Straightforward arguments based on (a) and (c)
yield the rest of (d).
Part (e) follows immediately from Proposition 2.1.39 (b) and the fact
that X is a topological group under addition of vectors. For (f), suppose
that U is a neighborhood of 0 in X. If x E X, then the continuity of
multiplication of vectors by scalars yields a positive r x such that tx E U
when 0 < t < r Xl so x E tU when t > r;l. Thus, the set U is absorbing,
which proves (f).
To prove (g), suppose that U is a neighborhood of 0 in X. The con-
tinuity of vector addition yields neighborhoods U1 and U2 of 0 such that
U1 + U2 <;;: U. Let Ua = U 1 nU2 n( -Udn( -U2 ). Then U3 is a neighborhood
of 0 such that U3 = -U3 and U3 + U3 <;;: U. The same procedure applied
to U3 instead of U yields a neighborhood U4 of 0 such that U 4 = -U4 and
U4 + U4 + U4 + U4 <;;: U. It follows that U4 + U4 + U4 does not intersect
X \ U, so the fact that U4 = -U4 implies that U 4 + U4 does not intersect
(X \ U) + U4 . Since (X \ U) + U4 is open, it follows that U4 + U4 does not
intersect (X \ U) + U4, so U4 + U4 <;;: U. The continuity of multiplication of
vectors by scalars produces a positive 8 and a neighborhood U5 of 0 in X
such that aU5 <;;: U4 whenever lal < 8. Let V = U{ aU5 : lal < 8}. Then
V is a balanced neighborhood of 0 lying in U4 , and
This proves all of (g) except for the convexity assertion, which uses (i) in
its proof and will be obtained below.
Suppose that A is a bounded subset of X, that Xo is an element of X,
and that ao is a scalar. Let U be a neighborhood of 0 in X and let V be a
balanced neighborhood of 0 in X such that V + V <;;: U. Let Sv be a positive
number such that A <;;: tV and Xo E tV whenever t > Sv. If t > sv, then
Xo + A <;;: tV + tV <;;: tU, from which it follows that Xo + A is bounded. If
t> (Iaol + l)sv, then aoA <;;: t(laol + 1)- l aoV <;;: tV <;;: tU, which shows
that aoA is bounded and finishes the proof of (h).
2.2 Vector Topologies 169
and
6. Claims: If s, t > 0, then B(s) + B(t) ~ B(s + t), and if 0 < s < t,
then B(s) ~ B(t). These claims follow in an obvious way from Steps
3 and 5.
7. Claim: For each positive t, the set B(t) is a balanced neighborhood
of O. To see this, first observe that if U and V are balanced neighbor-
hoods of 0, then so is U + V. It then follows from Step 1 that B(2n)
is a balanced neighborhood of 0 for each integer n, from Step 2 that
B(r) is a balanced neighborhood of 0 for each positive dyadic ratio-
nal r, and from Step 5 that B(t) is a balanced neighborhood of 0 for
each positive real number t.
8. Claim: X = U{ B(t) : t > O}. To see this, suppose that x E X.
Let n be a positive integer such that x E 2n B. An easy argument
2n
based on Step 1 shows that 2n B ~ Lj=l B = B(2 n ), from which the
claim follows.
9. For each x in X, let f(x) = inf{ t : t > 0, x E B(t)}. It follows from
Step 8 that f is finite-valued and so has its range in [0, +00). It is
clear that f(O) = 0 and that if t > 0, then f(x) ~ t if x E B(t)
and f(x) 2 t if x E X \ B(t). In particular, it follows that f(x) 2 1
whenever x E X \ B.
10. Claim: If x E X, then f(x) = f(-x). This is a straightforward
consequence of the definition of f and the fact that B(t) = -B(t) for
each positive t.
11. Claim: If x, y E X, then f(x + y) ~ f(x) + f(y) and If(x) - f(y)1 ~
f(x - y). To see this, suppose that x, y E X and that to > O. Then
there are positive reals s and t such that s < f(x) + to, t < f(y) + to,
X E B(s), and y E B(t). Then x + y E B(s) + B(t) ~ B(s + t), and
so f(x + y) ~ s + t < f(x) + f(y) + 2tO. It follows that f(x + y) ~
f(x) + f(y) Also,
f(x) - f(y) = f(x - y + y) - f(y)
~ f(x - y) + f(y) - f(y)
= f(x - y),
and similarly
f(y) - f(x) ~ f(y - x) = f(x - y).
It follows that If(x) - f(y)1 ~ f(x - y), which finishes the proof of
the claim.
12. Claim: The function f is continuous. For the proof of this, suppose
that Xo E X and that to > O. If x is in the neighborhood Xo + B(f)
of xo, then x - Xo E B(f), and so If(x) - f(xo) I ~ f(x - xo) ~ f. It
follows that f is continuous at xo, which proves the claim and finishes
the construction.
174 2. The Weak and Weak* Topologies
PROOF. Suppose that X is a TVS whose topology is To. The first order
of business is to show that the topology is in fact T 1. Let x and y be
distinct elements of X. Then there is a neighborhood U of a such that
either x ~ y + U or y ~ x + U; it can be assumed without loss of generality
that x ~ y + U. Then y is not in the neighborhood x - U of x, from which
it follows that the topology of X is T 1.
Now let Xo be an element of X and let F be a closed subset of X not
containing Xo. Since a is not in the closed set -Xo + F, there is a balanced
neighborhood B of 0 such that Bn (-xo +F) = 0. By Construction 2.2.13,
there is a continuous function f: X -. [a, +(0) such that f(O) = 0 and
f(x) 2: 1 whenever x E X \ B; notice that f(x) 2: 1 for each x in -Xo + F.
Let g(x) = min{l,f(x - xo)} whenever x E X. Then 9 is a continuous
function from X into [0,1] such that g(xo) = 0 and g(x) = 1 whenever
x E F. The topology of X is therefore completely regular.
There is a slight technical conflict between this definition and the one
previously given for the dual space of a normed spacc. If X is a normed
space, then by Definition 2.2.15 the dual space of X should be the vector
space of continuous linear functionals on X, while by Definition 1.10.1 it
should be the corresponding normed space. The only time this could cause
any real problem is when the vector space X underlying a normed space is
given a topology 'I different from its usual norm topology, and "X*" could
refer to either (X, 1111)* or (X, T)*, dual spaces that might not even have
the same underlying vector space; see Exercise 2.30. In this situation, care
will be taken to assure that no confusion results.
2.2 Vector Topologies 175
Since p(x) 2: 0 for each x in X, it is also true that x;;(y+o:xo) :s; p(y+o:xo)
whenever y E Y and 0: :s; 0, so Xo is dominated by p on Y + ({ xo}).
By the vector space version of the Hahn-Banach extension theorem, the
functional xi) can be extended to a linear functional x* on X such that
x'x :s; p(x) whenever x E X. Now Co includes a balanced neighborhood U
of 0, and x* (U) is a bounded subset of IR. since
Now drop the assumption that 0 E Co. Let Xl be an element of Co. Then
the interior -Xl + Co of the convex set -Xl + C contains 0 and does not
intersect the flat subset -Xl + F of X, so there is an x in X* such that
xx = 1 when X E -Xl +F, xx ~ 1 when X E -Xl +C, and xx < 1 when
X E -Xl + Co. It follows that xx = XXI + 1 when X E F, xx ~ X*XI + 1
when X E C, and xx < X*Xl + 1 when X E Co, and again the conclusion
of the theorem holds.
Finally, suppose that the scalar field is C. Let Xr be the real TVS ob-
tained by restricting multiplication of vectors by scalars to IR x X. Since
every subspace of X is also a subspace of Xn the set F is flat in X r . It
follows that there is a continuous real-linear functional z* on X and a real
number s such that zx = s when X E F, zx ~ s when X E C, and
z*x < s when X E Co. Let xx = zx - iz*(ix) for each x in X. It follows
from Proposition 1.9.3 that x* is a complex-linear functional on X with
real part z*. The continuity of z* and of the vector space operations of X
and C implies that x E X, so x has all the required properties.
The following three corollaries are analogs for locally convex spaces of the
normed space version of the Hahn-Banach extension theorem and of Corol-
laries 1.9.7 and 1.9.9 of that theorem. Notice that the third one requires
the topology to be Hausdorff.
The conclusion of the preceding corollary does not in general hold if the
vector topology is not required to be Hausdorff or is not required to be
locally convex. The examples about to be given to illustrate this depend
on the fact that if X is a TVS whose only convex open subsets are the
empty set and X itself, then X* contains only the zero functional. To see
this, suppose that X is a TVS having only those two convex open subsets
and that x* E X*. For each open ball U centered at the origin of IF, the
subset (x*) -1 (U) of X is a nonempty convex open subset of X and so must
be X. Since x*(X) is either {O} or IF, it follows that x* = O.
s:
2.2.24 Example. Suppose that 0 p < 1. It was shown in Examples 2.2.5
and 2.2.6 that the only nonempty convex open subset of Lp[O, 1J is Lp[O, 1]
itself, so (Lp[O,l])* = {O}. Therefore Lp[O,I] is a Hausdorff TVS having
more than just the zero element in which no two distinct elements can be
separated by a member of the dual space in the sense of Corollary 2.2.22.
Let f be a nonzero element of Lp [0,1] and let Y = ({J}), a one-dimen-
sional subspace of Lp[O, 1]. Let y*(af) = a whenever a E IF. Then y*
is a linear functional on Y that is continuous since its kernel is the closed
subspace {O} ofY. Since (Lp[O, 1])* = {a}, there is no member of (Lp[O, 1])*
whose restriction to Y is y*. Also, there is no member of (Lp[O, 1])* whose
kernel includes the closed subspace {O} of Lp[O, 1] and whose value at f
is 1. This shows that the assumption of local convexity in Corollaries 2.2.20
and 2.2.21 cannot in general be omitted.
The fact that (Lo[O, 1])* = {O} is a result of Otton M. Nikodym [175],
while the corresponding property of Lp[O, 1] when 0 < p < 1 is due to
Mahlon M. Day [46].
2.2.25 Example. Suppose that 0 < p < 1. For each m in N, the map x;"
that sends each member (an) of fp to its mth term am is clearly a continuous
linear functional on fp, so f; has enough members to separate the elements
of fp in the sense of Corollary 2.2.22 even though fp is not an LCS.
Mazur's and Eidelheit's separation theorems both require one of the con-
vex sets to have nonempty interior. The third and final separation theorem
of this section does not, but instead requires one of the convex sets to be
closed and the other to be compact, and also requires that the TVS in
question be locally convex.
The following lemma gives another example of a phenomenon previously
encountered in Proposition 2.2.11: In many ways, a compact subset of a
TVS behaves as if it were a singleton.
2.2.27 Lemma. Suppose that X is a TVS and that A and K are subsets
of X such that A is closed and K is compact. Then A + K is closed.
180 2. The Weak and Weak* Topologies
PROOF. Suppose that (a",) and (k",) are nets in A and K respectively
such that (a", + k",) converges to some x in X. It is enough to show that
x E A + K. From the compactness of K, there is a subnet (k/3) of (k",)
converging to some k in K. Since a/3 = (a/3 + k/3) - k/3 -+ x - k, it follows
that x - k E A, so x E k + A ~ A + K.
2.2.28 Theorem. (J. W. Tukey, 1942 [233]; V. L. Klee, 1951 [133]). Let
K and C be disjoint nonempty convex subsets of an LCS X such that K
is compact and C is closed. Then there is a member x* of X* such that
max{Rex*x: x E K} < inf{Rex*x: x E C}.
as required.
Tukey proved the preceding theorem for only one special type of locally
convex topology, the weak topology of a normed space to be defined later in
this chapter. The general case of the theorem is due to Klee. The hypotheses
on the two convex sets in the theorem cannot be relaxed very much, as
is illustrated by several examples in Tukey's paper and by an example
of Dieudonne [61], who showed that i!1 has two disjoint nonempty closed
bounded convex subsets that cannot be separated by a bounded linear
functional in the sense of the preceding theorem. Klee [134) later extended
Dieudonne's result from i!1 to all nonreflexive separable Banach spaces.
Also, the hypotheses of the theorem cannot be relaxed by only requiring X
to be a TVSj see Exercise 2.32.
One important consequence of Theorem 2.2.28 is that if it is known that
the topology of a vector space X is locally convex, and the dual space of X
under that topology is known, then the closure of each convex subset of X
is completely determined just by that information. This is the content of
the following corollary, which will have several major applications later in
this chapter.
2.2 Vector Topologies 181
2.2.29 Corollary. Suppose that a vector space X has two locally convex
topologies'rl and'r2 such that the dual spaces of X under the two topolo-
gies are the same. Let C be a convex subset of X. Then the 'rl-closure
of C is the same as its 'r 2-closure. In particular, the set C is 'r I-closed if
and only if it is 'r2-closed.
PROOF. It may be assumed that C =1= 0. Let X* represent the dual space
-:I,
of X under each of'r l and 'r2. For each x in X \ C ,use Theorem 2.2.28
to produce an x;
in X* such that
whenever ai, ... ,an E IF. It is easy to check that 1111 is a norm on X and
so is a Banach norm by Corollary 1.4.19. Now let 'I' be a Hausdorff vector
182 2. The Weak and Weak* Topologies
topology on X and let X'! and XII' II be X equipped with the topology 'I' and
the topology induced by 1111 respectively. Let xi, ... ,x~ be the coordinate
functionals for {Xl, ... ,xn}j that is, let X; (O:lXI + ... +O:nXn} = O:j when
Q: 1, ... ,Q: n E F and j = 1, ... , n. Then the kernel of each of the linear
functionals x; has dimension n- 1 and so is closed in X,! by Pn - 1 and the
lemma preceding this theorem. It follows from Theorem 2.2.16 that each x;
is continuous on X'r, and from the defining formula for 1111 that each is x;
continuous on XII II Since x = (XiX}Xl + ... + (x~X}Xn whenever x E X,
the continuity of each x; and of the vector space operations of X'r and XII' II
assures that the identity operator on X, viewed as a map from X'! onto XII'!!
or from XII'!! onto X,!, is continuous. The topologies of X'r and XII'!! are
therefore the same, which proves Pn and finishes the induction.
It follows from the continuity of the vector space operations of Y and of the
maps O:lXl + ... + Q:nXn I--> O:j when j = 1, ... ,n that T is continuous.
Exercises
p(j,g)= n 1 +lj_gld
Ij-gl
1 lL
"
d((al, ... ,an),(,81, ... ,,8n)) = Llaj -,8jIP.
j=l
Prove that e; is an LCS. Now consider the scalar field to be ~ and sketch
ei
the graph of the closed unit ball of j 2, that is, the set of points of j 2 ei
no more than one unit from the origin.
2.21 Suppose that 0 < p < 1 and that x and yare two different elements of fp.
Show that there are disjoint convex neighborhoods C x and C y of x and y
respectively. Notice that fp has many more convex open sets than does
Lp[O,I] (which has only two), even though it does not have enough to
make it an LCS.
2.22 Suppose that (\1,~, J-t) is a finite positive measure space and 0 < p :s: 00.
Prove that the set underlying Lp(\1, E, IL) is a dense subset of Lo(\1,~, IL).
2.23 Show that equality need not hold in the inclusions A + Ii c;;;; A + Band
AD + B O C;;;; (A + B)O in Theorem 2.2.9 (d). Also, find a balanced subset
of a TVS whose interior is not balanced.
2.24 Prove that if A and B are bounded subsets of a TVS, then A + B is
bounded.
184 2. The Weak and Weak* Topologies
2.29 Suppose that 0 <p < 1. If (an) E Cp and ({3n) E Coo, let
Prove that the map ({3n) f--> x({3n} is a vector space isomorphism from Coo
onto C;.
2.30 Give an example of a vector space X with a norm 1111 and a Hausdorff
vector topology '1" such that (X, 1111) * and (X, '1")* do not have the same
underlying vector space.
2.32 Find two disjoint. nonempty convex subsets K and C of a TVS, with K
compact and C closed, such that K and C cannot be separated by a
continuous linear functional in the sense of Theorem 2.2.28.
2.33 Let X be a TVS and let S be the collection of all x in X such that each
open set containing either x or 0 contains both.
(a) Prove that S is a subspace of X.
(b) Prove that every open set that intersects S actually includes S.
(c) Prove that if X is not completely regular, then S i= {a}. Thus, a TVS
that is not completely regular has very weak separation properties.
(d) Let 7r be the quotient map from X onto the quotient vector space
XIS; that is, let 7r(x) = x + S for each x in X. Define the quotient
topology of X/ S by letting a subset U of X/ S be open if and only
if 7r- 1 (U) is open in X. Show that this defines a completely regular
vector topology for X/ S. (This is one way to start with a TVS that
is not To and construct a completely regular TVS from it.)
*2.3 Metrizable Vector Topologies 185
(e) Let (0, L, p,) be a positive measure space and let X be the vector
space of all scalar-valued p,-integrable functions on O. In this exer-
cise, do not consider functions to be the same if they agree almost
everywhere but differ somewhere on O. For each I in X and each
positive E, let U(f, t) = {9 : 9 E X, fol9 - II dp, < t}. Prove that
{U(f, t) : I E X, E > O} is a basis for a locally convex topology
for X. Describe S. Describe XIS and the quotient topology of XIS.
2.34 The purpose of this exercise is to study vector topologies for finite-dimen-
sional vector spaces without assuming that the topologies are Hausdorff.
Part (a) of Exercise 2.33 is needed for this. Let 'I be a vector topology
for a finite-dimensional vector space X and let S be the subspace of X
consisting of all x in X such that each open set containing either x or 0
contains both. Let Y be a vector space complement of S in X, that is,
a subspace of X such that S + Y = X and S n Y = {O}. (For example,
let lEI be a vector space basis for S, let 1E2 be a subset of X such that
lEI n 1E2 = 0 and lEI U 1E2 is a basis for X, and let Y = (1E2)')
(a) Prove that the relative topology that Y inherits from X is the unique
Hausdorff vector topology for Y.
(b) Prove that 'I = {U + S: U is an open subset of Y}.
(c) Prove that 'I is a locally convex subtopology of the unique Hausdorff
vector topology of X.
This exercise will be continued at the end of Section 2.4 as Exercise 2.51,
in which will be developed a characterization of vector topologies on finite-
dimensional vector spaces in terms of linear functionals on the space.
The reader should be warned that some sources include local convexity
in the definition of an F-space, while others omit it from that of a Frechet
space. It is also common to require that the topology of a TVS be induced
by some complete invariant metric before the TVS is called an F-space.
Banach noticed that all of the topologically complete TVSs he studied ac-
tually do have their topologies induced by complete invariant metrics, and
he 3.'lked in [13] if this must always be the case. It is a remarkable fact due
to Victor Klee that the answer is yes, as will be shown in Theorem 2.3.20.
The definition of an F-space that requires invariance is therefore equivalent
to the one given above that does not.
2.3.4 Example. Suppose that 0 < p < 1. It was shown in Example 2.2.6
that if J.L is any positive measure, finite or not, on a a-algebra L. of subsets of
a set D, then Lp(D, L., J.L) is a TVS whose topology is induced by a complete
metric, so Lp(D, L., J.L) is an F-space. It was also shown in Example 2.2.6
that Lp[O, 1] is not locally convex, and so is an F-space that is not a Frechet
space. The space fp is a further example of an F-spaee that is not a Frechet
space, 3.'> was seen in Example 2.2.7.
As will be shown by Example 2.3.25, not every Frechet spaee has its
topology induced by a norm.
This section contains two major results linking metrizability of a vector
topology to metrizability by an invariant metric. The first, Theorem 2.3.13,
*2.3 Metrizable Vector Topologies 187
2.3.6 Definition. A local basis for a vector topology is a basis for the
topology at O.
That is, a local basis for a vector topology is a collection 1)30 of neigh-
borhoods of 0 such that every neighborhood of 0 includes a member of 1)30.
Notice that every vector topology has a local basis, namely, the collection
of all neighborhoods of O.
The following proposition contains a few basic facts about local bases
that all follow easily from the appropriate definitions and from various
parts of Theorem 2.2.9.
2.3.7 Proposition. Let 1)30 be a local basis for the topology of a TVS X.
(a) The collection of all translates of members of 1)30 is a basis for the
topology of X.
(b) The space X has a local basis 1)3~ for its topology with cardinality
no more than that of 1)30 such that every member of 1)3~ is balanced.
If X is locally convex, then 1)3~ can be selected so that each of its
members is balanced and convex.
(c) If U is a neighborhood of 0 in X, then there is a member V of 1B0
such that V c: V c: V + V c: U.
2.3.8 Corollary. Every TVS has a local basis for its topology consisting
of balanced sets, and every LCS has a local basis for its topology consisting
of balanced convex sets.
2.3.9 Corollary. A TVS is locally convex if and only if its topology has a
local basis consisting of convex sets.
2.3.10 Corollary. If a vector space has two vector topologies with a com-
mon local basis, then the two topologies are the same.
If the topology of a TVS is induced by a metric, then the open balls cen-
tered at the origin whose radii are reciprocals of positive integers form a
countable local basis for the topology. This, together with the following re-
sult, shows that in fact the existence of a countable local basis characterizes
metrizable vector topologies among all Hausdorff vector topologies.
188 2. The Weak and Weak* Topologies
2.3.11 Theorem. (G. Birkhoff, 1936 [27]; S. Kakutani, 1936 [124]). SUJr
pose that X is a Hausdorff TVS whose topology has a countable local basis.
Then the topology of X is induced by an invariant metric such that the
open balls centered at the origin are balanced. If X is locally convex, then
its topology is induced by a metric with the preceding properties and for
which all open balls are convex.
PROOF. Let 'I be the given vector topology for X. This proof is based
on Construction 2.2.13, so the local basis about to be defined is indexed
to match the notation of that construction. As an easy consequence of
parts (b) and (c) of Proposition 2.3.7, the topology 'I has a local basis
{B(2-n) : n = 0,1,2, ... } such that, for each n, the set B(2-n) is balanced
(and, if X is an LCS, convex) and B(2- n- 1 ) + B(2- n- 1 ) ~ B(2-n); call
this local basis 113 0. Let B = B(l), then let Band {B(2-n) : n EN} be
used in Step 1 of Construction 2.2.13 to get the construction started. In
that construction, the map t f--7 B(t) from {2- n : n = 0,1,2, ... } into 'I
is extended to (0,00) in such a way that each B(t) is a 'I-neighborhood
of 0 that is balanced (and convex if each member of 113 0 is convex) and
B(s) ~ B(t) whenever 0 < s < t. It is then shown that the formula
f(x) = inf{ t : t> 0, x E B(t)}
defines a 'I-continuous nonnegative-real-valued function on X such that
f(O) = 0 and such that f(x) = f( -x) and f(x+y) ::; f(x) + fey) whenever
x, y EX. If x is a nonzero member of X, then the fact that 113 0 is a
local basis for the Hausdorff topology 'I implies that there is a nonnegative
integer n such that x rt. B(2- n ) and therefore that f(x) ::::: 2- n > O. It
follows from all of this that the formula d(x, y) = f(x - y) defines an
invariant metric on X.
For each positive t, let Vet) = U{ B(s) : 0 < s < t}. Then each Vet) is
a 'I-neighborhood of 0 that is balanced (and convex if each member of 113 0
is convex), and {V (t) : t > O} is a local basis for 'I. It is easy to see that
Vet) = {x : x E X, f(x) < t} whenever t > 0, that is, that each Vet) is the
d-open ball of radius t centered at the origin, so the d-open balls centered
at 0 are all balanced (and convex if each member of 1130 is convex). Now
let U(x, r) denote the d-open ball ofradius r centered at x. Then for each x
in X and each positive r,
U(x, r) = {y : y E X, fey - x) < r}
= x + {y: y E X, fey) < r}
= x + VCr).
Combining Theorem 2.3.11 with its corollary yields the first of the two
major results of this section connecting metrizability of vector topologies
to metrizability by invariant metrics.
Notice that the preceding theorem does not say that the original metric
inducing the topology has to have the nice properties assured by the con-
clusions of the theorem, but rather that some metric inducing the topology
has those properties.
The second major result of this section linking metrizability of vector
topologies to metrizability by invariant metrics is an easy consequence of
a more general result about topological groups obtained by combining the
results of three lemmas, each of which is interesting in its own right.
j = 1, 2 and observing that (J (VI, n . V2, n)) must converge shows that there
is no ambiguity in the definition of Yl . Y2. Multiplication is therefore a
well-defined operation on Y.
Suppose that Yl, Y2, Y3 E Y and that sequences (Xl,n), (X2,n), and (X3,n)
in X are such that limn f(xj,n) = Yj when j = 1,2,3. Then
(Yl'Y2)'Y3 =limf((Xl
n n 'X2n)'X3n) J J J
= Yl . (Y2 . Y3),
so multiplication of elements of Y is associative. Let e be the identity of X
and let en = e for each positive integer n. Then
and similarly f(e) . Yl = Yl, so f(e) is an identity for Y. Notice next that
the invariance of dx implies that dx(xl,;' xl,;") = dX(Xl,m, Xl,n) whenever
m,n E N, so (J(xl,;)) is a Cauchy sequence in Y. Let Yo denote its limit.
Then
Yl . Yo = limf(xl,n
n
. xl,;) = f(e) = limf(xl,; . Xl,n) = Yo . Yl,
n
f (u . v) = lim
n
f (un . V n ) = f (u) . f (v),
which together with the fact that f is one-to-one shows that f is a group
isomorphism onto a subgroup of Y.
(2.1)
It follows that Xn ~ Xo in Y.
Let E be a positive number, and choose a positive integer N such that
2jN < E. Let m be a positive integer such that
(2.2)
The next lemma is not one of the three important lemmas used di-
rectly in the proof of Theorem 2.3.18, but is instead just a "sublemma"
for Lemma 2.3.17. The result is almost obvious, but needs a moment's
arguing.
The following lemma is due to Victor Klee, and the proof given here
is from Klee's paper. As Klee points out, the argument is essentially the
same as one used by S. Mazur and L. Sternbach [165] to show that every Go
subspace of a Banach space is closed.
In [135], Klee used an example due to Dieudonne [62] to show that the
conclusion of the preceding theorem can fail if the metric d x is only as-
sumed to be left-invariant.
As an easy consequence of Proposition 1.4.14 (c), a normed space is a
Banach space if there is some complete norm that induces its topology, and
Corollaries 2.1.50 and 2.1.51 provide a strengthening of that statement: A
*2.3 Metrizable Vector Topologies 193
The main consequence of Theorem 2.3.18 for the purposes of this section
is the following one, which gives an affirmative answer to Banach's question
mentioned after Definition 2.3.2.
A bit more can be said. The invariant metric used in the proof of the
preceding theorem can, by Theorem 2.3.13, be selected to have some special
properties. This is summarized by the following result.
Combining Theorems 2.3.13 and 2.3.20 with Corollary 2.1.50 yields the
following result immediately.
For example, every norm topology is locally bounded, since the open unit
ball of the space is bounded. The following result is a partial converse of
that fact.
There are, however, metrizable vector topologies that are not locally
bounded.
It is easy to check that d is an invariant metric and that a sequence (x( n))
of members of X converges to some x if and only if limn xjn) = Xj for
each j, for which reason the topology induced by d is called the topology
of termwise convergence. By Proposition 2.1.43, the invariance of d implies
that addition of vectors is continuous. Also, if a sequence (x(n)) of members
of X converges to some x and O:n -+ 0: in IF, then limn O:nxjn) = O:Xj for
each j, and so limn O:nx(n) = O:X, from which it follows that multiplication
of vectors by scalars is continuous. The topology of termwise convergence
is therefore a vector topology for X. Since every Cauchy sequence of mem-
bers of X is termwise Cauchy, hence termwise convergent and therefore
convergent to some member of X, the metric d is complete. The space X
with the topology of termwise convergence is therefore an F-space. Now
define a collection of neighborhoods of the origin of X by letting
Vn = {x : x EX, Ix j I < n- 1 when j = 1, ... , n }
*2.3 Metrizable Vector Topologies 195
for each positive integer n. It is easy to check that each Vn is convex and
that each open ball centered at 0 includes some Vn , from which it follows
that {Vn : n EN} is a local basis for the topology of X consisting of
convex sets and thus that X with the topology of termwise convergence is
a Frechet space.
It will now be shown that the topology of X is not locally bounded,
for which it is sufficient to show that no member of {Vn : n EN} is
bounded. Fix a positive integer n and a positive real number t and let x
be the member of X for which Xn+l = t and all other terms are O. Then
x E Vn \ tVn+l, which shows that Vn is not included in any positive scalar
multiple of Vn+l and therefore is not bounded.
Notice that since no neighborhood of the origin of X is bounded, the
topology of X is not induced by a norm, so X is an example of a Frechet
space whose topology is not compatible with any norm.
As is true for the theory of linear operators between normed spaces, the
theory of linear operators between TVSs is quite rich, especially when the
topology of the domain space is metrizable. The following definition and
theorem extend Definition 1.4.1 and Theorem 1.4.2 from normed spaces
into these more general settings.
It is not always true that part (c) of the preceding theorem implies parts
(a) and (b) when the topology of X is not metrizable. See Exercise 2.57.
*2.3 Metrizable Vector Topologies 197
The open mapping theorem, closed graph theorem, and uniform bound-
edness principle all have natural extensions to F-spaces that will now be
obtained. To make the statement of each extension conform as closely as
possible to the statement of the corresponding result for normed spaces
given in Section 1.6, reference will be made in the statement of each to
bounded linear operators. In each case, the domains of the linear operators
will be F -spaces, so for these operators boundedness will be equivalent to
continuity.
Several of the following results cite Banach's 1932 monograph [13] in
addition to earlier sources. When this occurs, the earlier references are
to proofs for Banach spaces, while the references to [13] are to Banach's
extensions of the results to F-spaces by substantially the same arguments.
that is, the set T(N) includes the neighborhood (T(V)) 0 _ (T(V)) 0
of Oy. It will therefore follow that Oy E (T( N) ) 0 once it is shown that
(T(V)) 0 is not empty. Since T(X) = Y and V is absorbing, it follows
that Y = Un T(nV), so by the Baire category theorem there must be a
positive integer no such that T(noV) is not nowhere dense in Y, that is,
such that (T( no V) ) Q =1= 0. It follows that (T(V) Q =1= 0 and therefore
that Oy E (T(N)r
198 2. The Weak and Weak* Topologies
L
m2
< dx(xj,Ox)
j=ml+1
L
00
< 2- j (
j=ml +1
= 2- m ,(,
from which it follows that the partial sums of the formal series 2:n Xn form
a Cauchy sequence and therefore that 2:n Xn converges. Let Xo = 2:n Xn
Since limn dY (Yo,T(2:7=1 Xj)) = 0, it follows that
Finally,
Then d xxY is a complete invariant metric that induces the product topol-
ogy of X x Y; see the discussion of product metrics in Appendix B. It
is easy to check that X x Y is an F -space when given its product topol-
ogy and the usual vector space operations for a vector space sum. Let
G = {(x, Tx) : X EX}; that is, let G be the graph of T in X x Y. It
follows from the hypotheses of this theorem that G is a closed subspace
of X x Y and so is itself an F-space. Since the map (x, Tx) f-> x from G
onto X is a one-to-one bounded linear operator, its inverse is bounded by
Corollary 2.3.30, so the map x f-> (x, Tx) f-> Tx is itself a bounded linear
operator.
and H. Steinhaus, 1927 [17]; S. Mazur and W. Orlicz, 1933 [164]). Let ~
be a family of bounded linear operators from an F-space X into a TVS Y.
Suppose that { Tx : T E ~} is bounded for each x in X. Then ~ is uniformly
bounded. In short, the pointwise boundedness of ~ implies its uniform
boundedness.
PROOF. To avoid having to think about a special case throughout this
proof, notice that it can be assumed that ~ i= 0. Let B be a bounded
subset of X and U a neighborhood of the origin Oy of Y. The theorem will
be proved once a positive s is found such that T(B) c:;: tU when T E ~ and
t > s. Let V be a balanced neighborhood ofOy such that V + V c:;: U and let
S = n{ T- 1 (V) : T E ~}, a closed subset of X because of the continuity
of each T in ~. If x E X, then the boundedness of {Tx : T E ~} assures
that {Tx : T E ~} c:;: nx V for some positive integer nx and therefore
that x E nxS. It follows that X = U{ nS : n EN}. By the Baire category
theorem, one of the closed sets nS, and hence S itself, must have nonempty
interior. Let Xo be a point in So and let W = Xo - So, a neighborhood of
the origin of X. For each T in ~,
PROOF. The continuity of the vector space operations of Y and the linearity
of each Tn together imply the linearity of T. Let B be a bounded subset
of X. The proof will be complete once it is shown that T(B) is bounded. The
convergence of the sequence (Tnx) for each x in X forces {Tnx : n EN}
to be bounded whenever x E X, which by the preceding theorem implies
that U{ Tn(B) : n EN} is bounded. As a subset of the bounded set
U{Tn(B) : n E N}, the set T(B) is bounded.
Some of the most important casualties are the normed space version of
the Hahn-Banach extension theorem and its main corollaries about con-
tinuous linear functionals. Most of these results do extend in some way to
Frechet spaces, as is shown by Corollaries 2.2.20-2.2.22 of Mazur's separa-
tion theorem, but this is due to the local convexity of Frechet spaces (and,
in the case of Corollary 2.2.22, to the fact that their topologies are Haus-
dorff) rather than to the metrizability of their topologies. These corollaries
do not in general extend to F -spaces that are not locally convex. For in-
stance, it was shown in Examples 2.2.24 and 2.2.25 that if 0 < p < I, then
Corollary 2.2.22, which says that Hausdorff LCSs have enough continuous
linear functionals to separate points, does extend to the non-Iocally-convex
F-space Ep but not to Lp[O, 1].
The situation is less ambiguous for Corollary 2.2.21, which says that
every LCS has the property that for each subspace of the space, every
continuous linear functional on that subspace has an extension as a con-
tinuous linear functional to the entire space. For F-spaces, this property
is called the Hahn-Banach extension property. In 1969, Duren, Romberg,
and Shields [69] noted that all of the classical non-Iocally-convex F-spaces,
as well as other examples of such spaces that they had studied, lack the
Hahn-Banach extension property, and asked if it might be the case that an
F-space has the Hahn-Banach extension property if and only if it is locally
convex. Nigel Kalton showed in 1974 [126] that the answer is yes. For an
exposition of Kalton's result, see [128], which is an excellent source for the
reader wishing to learn more about F -spaces.
Exercises
2.35 Let X be the vector space of all continuous functions from ]F into ]F, and
let
whenever f, g EX.
(a) Prove that d is a complete invariant metric.
(b) Prove that a sequence (fn) in X converges to an f in X if and
only if the following holds: For each compact subset K of ]F, the
sequence Un) converges uniformly to f on K. (Because of this, the
topology induced by d is called the topology of uniform convergence
on compact sets.)
(c) Prove that X with its d-topology is a Fnkhet space that is not locally
bounded. Conclude that this topology is not induced by a norm.
2.36 Show that in each of the two statements in Theorem 2.3.13, the invariant
metric in the conclusion of the statement can be selected to have the nice
properties of that conclusion and the further property that the diameter
202 2. The Weak and Weak* Topologies
of the space with respect to that metric is at most 1. (Recall that the
diameter of a nonempty subset A of a metric space with metric d is
sup{ d(x, y) : x, yEA }.)
2.37 (a) Suppose that fL is a positive measure on a a-algebra I: of subsets of a
n
set and that 0 < p < 1. Prove that Lp(n,~, fL) is locally bounded.
(b) Prove that Lo [0, 1] is not locally bounded.
2.38 Prove that no F-space has a count ably infinite vector space basis.
2.39 Prove that Zabrelko's lemma extends to F-spaces, that is, that every
countably subadditive semi norm on an F-space is continuous.
2.40 Prove that if a vector space has two topologies under which it is an F-
space, and one of the topologies includes the other, then the two topologies
are the same.
2.41 (a) Prove the following generalization of Theorem 1.3.14: Suppose that
C is a closed, convex, absorbing subset of an F-space. Then C in-
cludes a neighborhood of the origin. (While this can be done by
copying the proof of Theorem 1.3.11 verbatim, there is a shorter
proof available. Find it.)
(b) Suppose that X is a vector space with a topology. A barrel in X
is a closed, convex, balanced, absorbing subset of the space. The
space X is barreled if each of its barrels includes a neighborhood of
the origin. Conclude from (a) that every F-space is barreled.
(c) For some F-spaces, the observation in (b) is not very significant. To
see why this is so, list all of the barrels in Lp [0, 1] when 0 -s: p < 1.
2.42 Suppose that X is an infinite-dimensional metrizable TVS and that Y
is a TVS with more open sets than just 0 and Y (which happens, for
example, when Y is a Hausdorff TVS such that Y i {O}). Prove that
some linear operator from X into Y is unbounded. Conclude that every
infinite-dimensional metrizable TVS has a linear functional on it that is
unbounded.
2.43 An F-space is often said to have the Hahn-Banach extension property if,
for every closed subspace of the space, every bounded linear functional
on that subspace has an extension as a bounded linear functional to the
entire space. Show that this is equivalent to the definition of the Hahn-
Banach extension property for F -spaces given near the end of this section,
in which the word "closed" is omitted.
2.44 Some texts define an F-space to be a vector space X with a topology
ind uced by a complete invariant metric such that multiplication of vectors
by scalars is continuous in each variable separately; that is, which has
the property that for each scalar 0'0 and each vector Xo, the mappings
fao: X --> X and 9"'0: IF ---+ X given by the formulas fa.o(x) = aox and
9"'0 (a) = axo are continuous. The purpose of this exercise is to prove that
this definition is equivalent to the one given in Definition 2.3.2. For the
moment, let an Fa-space be a vector space satisfying the above alternative
definition.
2.4 Topologies Induced by Families of Functions 203
The term weak topology is included in the preceding definition for ref-
erence, since it is often given this meaning. However, there is a specific
topology for normed spaces called the weak topology whose study will be
taken up in the next section. To avoid confusion, the term weak topology
will not be used here in the more general sense.
the topological product rrfE~ Y f ; notice that if ~ is empty, then the nota-
tion (J(x)) fE~ is being used to represent the unique element of rrfE~ Y f
lt follows immediately that the map x I-> (J(x)) fE~ is a homeomorphism
from X onto a topological subspace of rrfE~ Y f provided the map is ane-
ta-one. A moment's thought shows that the property needed to assure that
the map is one-to-one is exactly the following one.
or X2 must contain both, so the same holds for each member of the standard
basis for the ~ topology, from which it follows that the ~ topology is not
even To.
d(XI,X2) = L min{1,dn(Jn2~d,fn(X2))}.
n
is a sub basis for the topology of X that generates a basis for that topology
consisting of convex sets, so X is an LeS.
Let fo be a continuous linear functional on X. There is some neigh-
borhood of 0 in X that is mapped by fo into the open unit ball of IF,
from which it follows that there is a nonempty finite collection h,, fn
of members of X' and a corresponding collection U 1 , .. , Un of neigh-
borhoods of 0 in IF such that fll(Ut} n ... n f;; 1 (Un) is mapped by fo
into the open unit ball of IF. Suppose that x E kerCh) n ... n ker(fn).
Then mx E fl-l(Ut} n ... n f;;l(Un ) for each positive integer m, and so
mlfo(x)1 = Ifo(mx)1 < 1 whenever mEN, which implies that x E ker fo.
208 2. The Weak and Weak* Topologies
Let
The final two re!:>ults of this section concern boundedness with respect
to the topology induced on a vector space X by a !:>ubspace of X#. The
first gives a useful test for such boundedness, while the second shows one
dramatic difference between topologies of this sort and norm topologies
when the topologizing subspace of X# is infinite-dimensional.
PROOF. Throughout this proof, the topology of X is the X' topology. Let A
be a subset of X. Suppose first that A is bounded. Let f be a member of X'
and let U be the open unit ball of IF. Then there is a positive t such that
A ~ tj-I(U), which implies that f(A) ~ tU and therefore that f(A)
is bounded. Conversely, suppose that f(A) is bounded whenever f E X'.
Let Uo be a neighborhood of 0 in X and let iI, ... , fn be members of X' and
VI, ... , Vn neighborhoods of 0 in IF such that fll (VI) n ... n f;; 1 (Vn) ~ Uo.
The boundedness of each h(A) yields a positive S i:iUch that fJ(A) ~ tVj
foreachj when t > s. It follows that A ~ t(J1I(VJ)n nf;;I(Vn)) ~ tUo
when t > s, so A is bounded.
Exercises
2.45 Give an example of a set X and an uncountable separating topologizing
family ~ of functions for X such that the ~ topology of X is not metrizable
even though each of the topological spaces into which the members of ~
map X is a metric space. Exercise 2.14 might be helpful.
2.46 If X is an infinite-dimensional normed space, then some linear functional
on X is not continuous; see Theorem 1.4.11. Show that this result does
not generalize to infinite-dimensional Hausdorff locally convex spaces.
2.47 Prove that if a vector space with a topology has a linearly independent
sequence in it that converges to 0, then some linear functional on the space
is not continuous. Use this fact and your example from the preceding
exercise to give an example of an infinite-dimensional Hausdorff LeS in
which no linearly independent sequence converges to O.
2.48 Let X be a set and ~ a separating topologizing family of functions for X.
Suppose that for each f in ~, the topological space Yf into which f maps X
is Hausdorff. Prove that X is compact with respect to its J topology if
2.5 The Weak Topology 211
the dual space of X with respect to the norm topology of X, except where
explicitly stated otherwise, even in contexts in which another topology for X
is being discussed. Recall also the convention from Section 1.2 concerning
topological terms when a normed space is involved: Whenever reference is
made to some topological property in a normed space without specifying the
topology, the norm topology is implied. For example, if a subset of a normed
space is said without further qualification to be open, then it is meant that
it is open with respect to the norm topology of the space.
It is time to introduce the first of the two topologies that are the main
topics of this chapter.
That is, the weak topology of a normed space is the smallest topology
for the space such that every member of the dual space is continuous with
respect to that topology. See Proposition 2.4.1 and Definition 2.4.2.
Since the dual space of a normed space X is a separating family of
functions for X by Corollary 1.9.9, and each x* in X* maps X into the
completely regular space IF, it follows from Proposition 2.4.8 that the weak
topology of X is itself completely regular. By Theorem 2.4.11, this topology
is also locally convex, and the dual space of X with respect to this topology
is exactly X*. As the smallest topology for X with respect to which every
member of X* is continuous, the weak topology of X must be included in
every topology with respect to which the members of X* are all continuous,
and in particular must be included in the norm topology of X. These
observations are summarized in the following two results.
such that x E X and A is a finite subset of X*; see Proposition 2.4.12 for the
development of this basis and some other information about it. The reader
might also wish to review the properties of a TVS given in Theorem 2.2.9,
since those properties will be used extensively in what is to follow.
As a special case of Definition 2.2.8, a subset A of a normed space X
is weakly bounded if, for each weak neighborhood U of 0 in X, there is a
positive Sf} such that A ~ tU whenever t > Sf}. By Proposition 2.4.14, this
is equivalent to requiring that x* (A) be a bounded set of scalars for each x*
in X'. As Example 2.5.4 shows, the weak topology of a normed space can
be a proper subtopology of the norm topology, so it might seem easier for
a subset of a normed space to be weakly bounded than to be bounded.
Perhaps surprisingly, this is not the case.
as required.
214 2. The Weak and Weak* Topologies
such as the way Co is usually identified with lI, then the weak topologies
of the two spaces are preserved by the same isometric isomorphism.
If X is an infinite-dimensional normed space, then Corollary 2.5.9 implies
that its open unit ball cannot be weakly open, so the norm and weak topolo-
gies of X must differ. This could not happen if X were finite-dimensional,
for then the norm and weak topologies of X would both be the unique Haus-
dorff vector topology of X and so would be the same; see Theorem 2.2.31.
Summarizing these observations yields the following useful characterization
of normed spaces whose norm and weak topologies are identical.
2.5.16 Theorem. (S. Mazur, 1933 [162]). The closure and weak closure
of a convex subset of a normed space are the same. In particular, a convex
sllbset of a normed space is closed if and only if it is weakly closed.
Mazur actually showed that every closed convex subset of a normed space
is weakly sequentially closed, but his method can be used to prove the more
general result.
2.5.17 Corollary. (S. Banach, 1932 [13]). The closllre and weak closure
of a subspace of a normed space are the same, so a sllbspace of a normed
space is closed if and only if it is weakly closed.
PROOF. It follows from Theorem 2.2.9 (i) that coCA) = coCA) = coCA) W =
co W(A).
2.5 The Weak Topology 217
One of the basic properties of normed spaces is that the norm function
x ~ Ilxll is continuous. However, it does not have to be weakly continuous.
For instance, the norm function on 2 is not weakly continuous since, by
Example 2.5.4, there are sequences in the unit sphere of 2 that converge
weakly to o. In fact, it can be shown that the norm function is weakly
continuous if and only if the norm and weak topologies of the space are the
same, that is, if and only if the space is finite-dimensional. See Exercise 2.54.
Thus, it is not always true that Ilxall ---> Ilxll when a net (xa) in a normed
space converges weakly to some x. Something can be said in this situation,
but saying it requires the definition of the limit inferior of a net of real
numbers.
Suppose that (ta)aEI is a net of real numbers. For each 0:: in I, let ia =
inf{ t{3 : 0:: ::S f3} and Sa = sup{ t{3 : 0:: ::S f3}; notice that ia and Sa could
be infinite. If 0::1 ::S 0::2, then i al ::; ia2 and Sa,;::: Sa2l from which it follows
that lima ia and lima Sa both exist, provided that the notion of the limit
of a net of real numbers is extended in the obvious way to nets of extended
real numbers, that is, to nets in the ordered set lRu { -00, +oo}. By analogy
with sequences of real numbers, it is natural to call these limits the limit
inferior and limit superior, respectively, of (ta).
2.5.20 Definition. Let (ta)aEI be a net in R Then lima inf{ t{3 : 0:: ::S f3}
and lim" sup{ t{3 : 0:: ::S f3} are, respectively, the limit inferior and limit su-
perior of (t,,), and are denoted, respectively, by liminfa t" and limsuPa ta.
See Exercise 2.55 for some properties of the limit inferior and limit su-
perior that can be extended from sequences to nets.
The rest of this section deals with sequential properties of the weak
topology and ways in which those sequential properties can mimic those of
the norm topology.
It follows from Proposition 2.5.15 that every infinite-dimensional normed
space contains a weakly Cauchy net with no weak limit, but that does not
eliminate the possibility that all of the weakly Cauchy sequences in such a
space could be weakly convergent. This sometimes happens.
Llo:~k)l:S; M.
11,=1
n=l
2.5 The Weak Topology 219
Llanl::; M,
n=l
for each j. Let (3~) = ta~kj) - tan for each j and each n. Then (jJ~) ';:1
is a weakly Cauchy sequence in f1 such that 11(3$/)111 : :;:. 1 for each j and
limj (3$/) = 0 for each n. After thinning ((3$/))';:1 if necessary, it may be
assumed that there is a sequence (nj) of nonnegative integers such that
0= n1 < n2 < ... and
m=nj+1
for each j. For each positive integer m, let 'Ym be a scalar of absolute
value 1 such that if j is the positive integer for which nj < m ::; njH, then
'Ym(3';/,) = (-lFI(3';/,\ Let x* be the member of i that is represented in
the usual way by the member bn) of f oo . For each positive even integer j,
L
nj+l
Re x* (3$/) = 'YmjJ';/,) +
L
nj+l
> 1(3';/,)1-
By Example 2.5.4, the space 2 does not have Schur's property. As will
be shown, it does have the following weakened version of the property.
n n n n
=0.
The reason that the Radon-Riesz property is named after J. Radon and
F. Riesz is that they proved that the spaces Lp(O,~,J-l), where J-l is a
positive measure on a O"-algebra ~ of subsets of a set 0, have it when
1 < p < 00, with Radon's proof coming in 1913 [192] and Riesz's in 1928~
1929 [196, 197]. M. 1. Kadets and V. L. Klee used versions of the Radon-
Riesz property to develop the ingredients for the proof that all infinite-
dimensional separable Banach spaces are homeomorphic; see [119], [138],
and [122]. In addition, Kadets used it in [120] to prove that every sepa-
rable normed space has an equivalent locally uniformly rotund norm; the
definition of this rotundity property will be given in Chapter 5. Because
2.5 The Weak Topology 221
of the work that Kadets and Klee did with the Radon-Riesz property, it
is sometimes named after them. The Radon-Riesz property is also some-
times called property (H) since a strengthened form of it was given that
name and studied in a 1958 paper by K. Fan and I. Glicksberg [77], and
M. M. Day later adopted that notation for the Radon-Riesz property in
his book [56]. Incidentally, the letter H in the notation does not stand for
anything. The Fan and Glicksberg paper has an alphabetically-labeled list
of properties for normed spaces that starts with (A) and ends with (H), and
their strengthened version of the Radon-Riesz property happens to fall last
in the list. Such is often the way mathematical notation gets established!
Exercises
2.53 Suppose that X is CQ or p, where 1 < p < 00. Let (,B~")aEI be a net
in X and (,Bn) an element of X.
(a) Show that if (,B~a ~ (,Bn), then ,B~a) --+,Bn for each n.
(b) Show that if the net (,B~Q)QEI is bounded and ,B~u) --+ ,Bn for
each n, then (,B~" ~ (,Bn).
(c) Show that the conclusion of (b) can fail if the net (,B~C<)aEI is not
required to be bounded.
(d) Suppose that (,B~a)OEl is a net in 1 and (,Bn) an element of fl.
Does the result stated in (a) still hold? What about (b)?
2.54 Suppose that X is an infinite-dimensional normed space. Show that there
is a net in Sx that converges weakly to 0. (Notice that this implies that
the map x r--+ Ilxll from a normed space into IF' is not weakly continuous if
the normed space is infinite-dimensional.)
2.55 For the purposes of this exercise, the notions of infimum, supremum, and
net limit in ~ have been extended in the obvious way to the extended
real numbers. Suppose that (t a ) and (Ua) are nets in ~ having the same
index set I. Prove that the following hold.
(a) liminfutu = sup{inf{ttl : Q ~,B}: Q E I} and limsuPata =
inf { sup{ ttl : a ~ ,B} : Q E I }.
(b) liminf" ta ~ limsupo t".
(c) If s ;::: 0, then liminf,,(st,,) = sliminfata and limsuPa(st,,)
s lim sup" ta.
(d) If s ~ 0, then lim inf,,(st,,)
slim infa t".
(e) liminf"t" + lim infa U a ~ liminfa(ta+ua) and limsuPa(ta+ua ) ~
limsupu ta + limsuPa u" except in cases in which a sum of liminf's
or limsup'6 i6 formally +00 - 00 or -00 + 00 and therefore not
defined.
(f) There arc subnets (t"() and (to) of (ta) such that lim"( t"( = lim inf a ta
and limo to = lim sup" ta.
222 2. The Weak and Weak* Topologies
(g) If (td is a convergent subnet of (t",), then lim inf" t", ::; lim, t, ::;
lim sup" t",. (Notice that lim, t, could be oo.)
(h) If (td is any subnet of (t,,) whatever, then lim inf" t" ::; lim inf, t, ::;
lim sUPe t( ::; lim sup" to.
(i) lim inf" t" = lim sup" to if and only iflimo to exists. If lima to exists,
then lima t" = lim inf a t" = lim sup" ta.
2.56 Prove that a normed space with its weak topology is of the second category
in itself if and only if the space is finite-dimensional. (The fact that an
infinite-dimensional normed space with its weak topology is of the first
category in itself is a 1938 result of J. V. Wehausen [241].)
2.57 Find a one-to-one linear operator from a Hausdorff TVS onto a normed
space such that the operator is bounded but not continuous. (Compare
this to Theorems 1.4.2 and 2.3.28.)
2.58 Prove that if 1 is isomorphic ally embedded in a normed space X, then
some bounded sequence in X h8.5 no weakly Cauchy subsequence. (Com-
pare Exercise 2.66.) This is the easy part of Rosenthal's J\ theorem (H. P.
Rosenthal, 1974 [199]): Each bounded sequence in a Banach space X has
a weakly Cauchy subsequence if and only if 1 is not isomorphically em-
bedded in X. See Joseph Diestel's book [58] for an extensive discussion
of this theorem and its consequences.
2.59 Show that Co is not weakly sequentially complete.
2.60 Suppose that K is a compact Hausdorff space. Show that C(K) is weakly
sequentially complete if and only if K is finite.
2.61 Suppose that 1 < p < 00. Show that p is weakly sequentially complete.
(Do not use any results from later sections of this book to do this. As will
be seen in Section 2.8, every reflexive normed space is weakly sequentially
complete, but that result is a corollary of a fairly deep theorem of that
section.)
2.62 Show that the spaces Co and p such that 1 < p < 00 all lack Schur's
property.
2.63 Show that if a normed space X h8.5 Schur's property, then so does every
subspace of X, but in general the same cannot be guaranteed of every
quotient space X / M such that M is a closed subspace of X.
2.64 Show that a normed space X h8.5 the Radon-Riesz property if and only
if it satisfies this condition: Whenever (Xn) is a sequence in Sx and x is
an element of Sx such that Xn ~ x, it follows that Xn --> x.
2.65 A normed space X has the semi-Radon-Riesz property if it satisfies the fol-
lowing condition: Whenever (Xn), x, and (x;') are, respectively, a sequence
in Sx, an element of Sx, and a sequence in Sx such that x;'x n = 1 for
each nand Xn ~ x, it follows that x;'x --> 1.
(a) Show that the Radon-Riesz property implies the semi-Radon-Riesz
property.
(b) Show that Co does not have the semi-Radon-Riesz property.
2.6 The Weak* Topology 223
2.6.1 Definition. Let X be a normed space and let Q be the natural map
from X into X**. Then the topology for X* induced by the topologizing
family Q(X) is the weak* (pronounced "weak star") topology of X* or the
X topology of X* or the topology a(X*, X).
That is, the weak* topology of the dual space of a normed space X is the
smallest topology for X* such that, for each x in X, the linear functional
x* f-+ x*x on X* is continuous with respect to that topology.
By analogy with the weak topology, a topological property that holds
with respect to the weak* topology is said to hold weakly*5 or to be a weak*
property. Whenever w* is attached to a topological symbol, it indicates that
the reference is to the wea~* topology. Examples of this would be x~ .:!'.;+ x*,
* = X * , an d -A w .
W * - I'lma Xa
Let X and Q be as in Definition 2.6.1. If x* and y* are different elements
of X*, then there is an x in X such that x'x oF y*x, so Q(X) is a separating
family of functions for X*. It therefore follows from Proposition 2.4.8 and
Theorem 2.4.11 that the weak* topology of X* is a completely regular
locally convex topology and that the dual space of X' with respect to this
topology is Q(X). Since Q(X) ~ X** and the weak topology of X* is the
smallest topology for X* with respect to which all members of X** are
continuous, the weak* topology of X* is included in the weak topology
of X*, and the two topologies are the same if and only if Q(X) = X",
that is, if and only if X is reflexive. The following theorem and proposition
summarize these observations.
5 "Weakly'" might sound somewhat awkward, but the alternatives are at least equally
unpleasant. "Weak*ly" is perhaps grammatically correct, but too outrageous to consider.
"Weak*" is sometimes used as both an adjective and adverb, but consistency then
demands that "weak" be used as an adverb, and calling a function "weak continuous"
also seems awkward. "Weakly'" was suggested to the author by Mahlon Day.
224 2. The Weak and Weak* Topologies
2.6.3 Corollary. Let X be a normed space. Then the weak* and norm
topologies of X* are the same if and only if X is finite-dimensional.
PROOF. If X is infinite-dimensional, then the weak topology of X* is a
proper subtopology of the norm topology by Proposition 2.5.13, and there-
fore so must be the weak* topology. If X is finite-dimensional, then the
weak*, weak, and norm topologies of X' are each the unique Hausdorff
vector topology of X*.
Suppose that X is a normed space and that A c:;: X*. In accordance with
Definition 2.2.8, the set A is weakly* bounded if, for each weak* neighbor-
hood U of 0 in X*, there is a positive Su such that A c:;: tU whenever
t > su. By Proposition 2.4.14, the set A is weakly* bounded if and only if
{x*x : x* E A} is bounded in II? for each x in X. The following result is
the weak* analog of the fact that a subset of a normed space is bounded if
and only if it is weakly bounded. However, notice the requirement that X
be a Banach space.
Neither the preceding theorem nor any of its three corollaries remain true
if X is only required to be a normed space; see Exercise 2.68. However, the
following result does hold for every normed space, whether or not it is
complete.
The following two results are the weak* analogs of Propositions 2.5.14
and 2.5.15. The first of the two has a completeness hypothesis that cannot
be omitted; sec Exercise 2.70.
and
so sllxllb :s: Ilxll :s: tllxllb. Therefore II lib is equivalent to the original norm
of X.
228 2. The Weak and Weak* Topologies
Let II lie be the dual norm on (X, 11'llb)*' All that remains to be proved is
that 11lIa = II lie. Fix an x* in X*. If Ilx*lla :S 1, then Ix*xl :S 1 whenever
Ilxllb :S 1, so Ilx*lle :S 1. It follows that Ilx*lle :S Ilx*lla no matter what the
value of Ilx*lla is. All that is left to be shown is that Ilx*lle ~ Ilx*lla.
Let E be a fixed positive number. Suppose that n E N and that Xl, ... , Xn
is a basis for an n-dimensional subspace Mo of X. Let (T(x)(x*) = x*x
whenever X E X and x' E X'. Then T is an isometric isomorphism from
(X, II lib) into (X*, 11lla)*. If aI, ... , an ElF, then
Notice that xMo and x' agree on Mo. It follows that for each finite-dimen-
sional subspace M of X there is an x M in X* such that x M and x* agree
on M and IlxM Iia :S Ilx* lie + . Preordering the collection of all finite-
dimensional subspaces of X by declaring that Ml :::S M2 when Ml 0;;;; M2
makes (x M) into a net. Since x M :!!.* x*, it follows that
So far, most of the results of this section have emphasized the similarities
between the weak and weak* topologies, especially when the weak* topol-
ogy is for the dual space of a Banach space. As the following two examples
show, there are also some fundamental ways in which the two topologies
differ.
211(a n )lh and IIT-l(an)lIl :::: 211(a n )lIl whenever (an) Eco,
the operator T
is an isomorphism from Co onto itself.
Now let (e~) be the sequence of standard unit vectors of 1, viewed as
members of Co. Then (e~) converges weakly* to 0, but the sequence (Te~)
does not since (Te~)(l, 0, 0, 0, ... ) = 1 for each n. The operator T is there-
fore not weak*-to-weak* continuous, even though it is a norm-to-norm iso-
morphism from Co onto itself. Thus, the weak* analogs of Theorem 2.5.11
and Corollary 2.5.12 both fail.
As will be seen in Corollary 3.1.12, it is true that if X and Yare normed
spaces and T is a weak*-to-weak* continuous linear operator from X*
into Y*, then T is norm-to-norm continuous.
The next theorem illustrates another way in which the weak* topology
is strikingly different from both the norm and the weak topology. As was
mentioned in the introduction to this chapter, many of the difficulties that
arise when trying to extend familiar facts about finite-dimensional normed
spaces to the infinite-dimensional case come about because of the loss of
the Heine-Borel property, that is, because the closed unit ball of a normed
space X is not compact unless X is finite-dimensional. When X is infinite-
dimensional, the fact that the weak topology of X is a proper subtopology
of the norm topology makes it easier for Bx to be weakly compact than
compact, and in fact Bx is weakly compact if and only if X is reflexive, as
will be seen in Theorem 2.8.2. This leads naturally to the question of what
conditions on X assure that Bx. is weakly* compact.
One of the major results of the theory of normed spaces is that B x'
is always weakly* compact. The general form of this result first appeared
in a 1940 paper by Leonidas Alaoglu, and for that reason is often called
Alaoglu's theorem. However, there is a result on page 123 of Banach's 1932
book [13] that easily implies the theorem when X is a separable Banach
space; see Exercise 2.73 at the end of this section. For that reason, Banach
is often given joint credit with Alaoglu for discovering the theorem, and
that is what is done here.
Let (x;) be a net in Bx- such that F(x;) converges to some (OX)XEBx
in JBx . The goal is to find an x* in Bx- such that F(x*) = (Ox)XEBx '
For each nonzero x in X, let x*(x) = IlxlloCllxll-1x), and let x*(O) = O.
Since x* (x) = Ilxlllim,a xHllxll-1x) = lim,a x;x whenever x E X \ {O}, and
thus x* (x) = lim,a x~x for every x in X, it follows that x* is a linear func-
tional on X, and furthermore that x* E Bx- since Ix*xl = lim,alx.exl ::; Ilxll
whenever x EX. For each x in B x,
PROOF. Let K be Bx- with the topology it inherits from the weak* topol-
ogy of X, and define T: X -+ C(K) by the formula (T(x))(x*) = x*x. It
is easy to check that T really does take its values in C(K) and that T is
linear. If x EX, then it follows from Theorem 1.10.9 that
{O} = nUn =
n
Bx- n { x* : x EX, xx = 0 for each x in UAn }.
n
The second dual X" of a normed space X is the dual of X' and therefore
has its own weak* topology. The rest of this section concerns that topology.
2.6.24 Proposition. The natural map Q from a normed space X into X"
is weak-to-weak* continuous, and in fact is a weak-to-relative-weak* ho-
meomorphism from X onto Q(X).
PROOF. Just notice that if (xa) is a net in X and x is an element of X,
then Xa :!!!., x if and only if x'x a --+ x'x for each x* in X', which happens
if and only if Qx", :'!'... Qx.
2.6.25 Corollary. Let X be a normed space and let Q be the natural map
from X into X**. Then the topologies that Q(X) inherits from the weak
and weak* topologies of X** are the same.
PROOF. It follows from the preceding proposition and Corollary 2.5.12
that Q is both a weak-to-relative-weak* homeomorphism from X onto the
subset Q(X) of X** and a weak-to-weak homeomorphism from X onto
the normed space Q(X). By Proposition 2.5.22, the weak topology of the
normed space Q(X) is the same as the relative weak topology of Q(X) when
Q(X) is viewed as a subset of X**, from which the corollary follows.
---w'
Q(Bx) is convex and weakly* closed, an application of Theorem 2.2.28
produces an Xo in X* such that
Ix~'x~1 ~ Rex~'x~
> sup{ Re x*'x~ : x" E Q(Bx) w}
~ sup{Rex~x: x E Bx}
= IIRex~1I
= Ilx~ll
2.6.28 Corollary. Let X be a normed space and let Q be the natural map
from X into X**. Then Q(X) is weakly* dense in X**.
PROOF. If x** is a nonzero element of X*', then Goldstine's theorem im-
plies that there is a net (x",) in Ex such that Qx", ---- Ilx"
w
11- 1x**, from
which it follows that Q(llx**llx",) ~. x**.
Exercises
2.61 Suppose that X is Co or f p , where 1 :S p < 00, and that X' is identified
in the usual way with the appropriate fq such that 1 :S q :S 00. Let
((,B~<))<>El be a net in X' and (,Bn) an element of X'.
(a) Show that if (,B~Q) ~. (,Bn), then ,B~< ----> (3n for each n.
(b) Show that if the net ((,B~"'))"'El is bounded and ,B~"') ----> ,Bn for
each n, then (,B~Q) ~. (,Bn).
(c) Show that the conclusion of (b) can fail if the net ((,B~Q) tEl is not
required to be bounded.
2.68 Let X be the vector space of finitely nonzero sequences equipped with the
f1 norm. For each positive integer m, let x;" : X ----> IF be defined by the
formula x;" ((an)) = m . am. Let A = {x;" : mEN}.
(a) Show that A is a weakly* bounded subset of X' that is not norm
bounded, and therefore that the conclusions of Theorem 2.6.7 and
Corollary 2.6.8 do not follow when X is only required to be a normed
space.
234 2. The Weak and Weak* Topologies
(b) Show that the conclusions of Corollaries 2.6.9 and 2.6.10 do not
follow when X is only required to be a normed space.
2.69 Let X be the vector space of finitely nonzero sequences equipped with
the 1 norm. Find a weakly* Cauchy sequence in X that is not weakly*
convergent.
2.70 Give an example of an infinite-dimensional normed space X such that the
weak* topology of X is metrizable.
2.71 Suppose that X is a normed space and that D is a dense subset of S x.
(a) Show that a bounded net (x~) in X converges weakly* to an ele-
ment x" of X" if and only if x~x --> xx for each x in D.
(b) Give an example to show that the requirement in (a) that the net
be bounded cannot be omitted.
(c) Let 0[0,1]* be identified with rca[O, 1] in the usual way. Show that
a bounded net (1-",) in 0[0,1] converges weakly* to an element 1-
of 0[0,1]* if and only if f[o.I) t n dl-",(t) --> f[o.I) t n dl-(t) for each
nonnegative integer n.
2.72 Suppose that X is a separable normed space. The goal of this exercise
is to show that there is a norm 11110 on X" such that for each bounded
subset A of X, the topologies induced on A by the weak* and 11110
topologies of X" are the same. It may be assumed that X #- {O}. Let
{x n : n EN} be a countable dense subset of S x.
(a) Define 11110: X" ..... IR by the formula Ilx"llo = Ln Tnlxxnl. Show
that 11110 is a norm on X', and that if 1111 is the usual norm of X,
then IIx"llo :::; Iix"1I for each x" in X .
(b) Let A be a bounded subset of X. Show that the topologies that
A inherits from the weak* and 11110 topologies of X" are the same.
Exercise 2.71 might help.
Do not use the Banach-Alaoglu theorem or any of its corollaries in your
arguments. Notice that this provides a proof of Corollary 2.6.20 that is
not based on the Banach-Alaoglu theorem.
2.73 The following result appears on page 123 of Banach's book [13], using
slightly different notation and terminology: If the Banach space X is
separable, then every bounded sequence in X" has a weakly* convergent
subsequence. Prove that this implies the conclusion of the Banach-Alaoglu
theorem when X is a separable Banach space. Of course, you should not
use the Banach-Alaoglu theorem or any of its corollaries in your argument.
You might find Exercise 2.72 helpful.
2.74 For this exercise, it will be useful to know that if K is a compact metric
space and P is the Cantor set, then there is a continuous map from P
onto K. (I. Rosenholtz has given a particularly nice proof of this that is
built up from a sequence of elementary lemmas; see [198].) Let X be a
separable normed space.
(a) Prove that X is isometrically isomorphic to a subspace of O(P).
(b) Prove that X is isometrically isomorphic to a subspace of 0[0, 1].
2.7 The Bounded Weak* Topology 235
2.15 (a) Prove that if X is a separable Banach space, then every weakly*
compact subset of X* is weakly* sequentially compact.
(b) Find a subset of f~ that is weakly* compact but not weakly* se-
quentially compact.
2.76 Use the Banach-Alaoglu theorem to prove that every bounded net in a
normed space has a weakly Cauchy subnet. (See Exercise 2.66 for another
line of proof that uses ultranets.)
2.11 Corollary 2.6.28 can be easily derived from results of this section much
more basic than Goldstine's theorem. Do so.
2.78 The purpose of this exercise is to generalize the notion of a weak* topology
to a more abstract setting. Let X be a topological vector space and let Y
be the subset of (X')# consisting of all linear functionals f on X for
which there is an xf in X such that fx' = X'Xf for each x' in X'.
Then the weak* topology of x' is the topology induced on X' by the
topologizing family Y.
(a) Prove that Y is a subspace of (X')# that is a separating family of
functions for X', and therefore that the weak* topology of X' is
a completely regular locally convex topology for X such that the
dual space of X' with respect to this topology is Y.
Suppose that A C;;; X. Then {x' : x EX', Ix'xl :S 1 for each x in A} is
called the absolute polar set for A.
(b) Let U be a neighborhood of 0 in X. Prove that the absolute polar
set for U is weakly* compact.
(c) Derive the Banach-Alaoglu theorem from (b).
The statement proved in (b) is itself sometimes called the Banach-Alaoglu
theorem.
B(x*, (xn)) = {y* : y' EX', I(y* - x*)xnl < 1 for each n}.
The topology for X* having the basis consisting of all such sets B(x*, (x n ))
is the bounded weak* topology of X'.
236 2. The Weak and Weak* Topologies
Of course, it must be shown that the collection of sets B(x*, (xn)) really
is a basis for a topology for X*. This is done in the proof of the next
theorem.
For notational convenience, properties related to the bounded weak*
topology will often be referred to as "b-weak*" properties; for example,
a set that is closed with respect to the bounded weak* topology will be
said to be b-weakly* closed. Convergence of a net (x~) to an x* with re-
spect to this topology will be denoted by writing x~ ~ x*.
is a subbasis for 'I w', and therefore that 'Iw' ~ 'Ibw" Notice that this
implies that 'I bw ' is Hausdorff. Now suppose that (X~)O:EI is a net in X*
that is norm convergent to an x* in X*. Then for each sequence (x n ) in X
converging to there is an a(x n ) in I such that x~ E B(x*, (x n )) when
a(x n } :5 a. This implies that x~ ~ x*, so 'Ibw' ~ 'In.
The next order of business is to show that 'Ibw' is locally convex. Suppose
that x*, y* E X*, that a E IF, and that U and V are b-weak* neighborhoods
of x* + y* and ax* respectively. Then there are sequences (un) and (v n )
in X converging to such that x* + y* E B(x* + yo, (Un)) ~ U and
ax' E B(ax*, (v n )) ~ V. It is easy to check that
An B(x', (x n ))
= An {y': y' E X*, ICy* - x*)xnl < 1 when n = 1, ... , no}.
Since An {y* : y* E X*, I(y* - x*)xnl < 1 when n = 1, ... , no} is open
with respect to the relative topology that A inherits from 'I w ', it is clear
that the relative topologies that A inherits from 'I bw ' and 'Iw' are the
same.
Let 'I be a topology for X* such that the relative topologies inherited
by each bounded subset of X* from 'Iw' and 'I are the same, and let U
be a member of 'I. The proof of the theorem will be finished once it is
shown that U E 'I bw " Let x* be an element of U. The proof that U is
b-weakly* open is based on the construction of a sequence Fo, F l , F 2 , . . .
of finite subsets of X such that for each nonnegative integer n,
238 2. The Weak and Weak* Topologies
(1) {y* : y* E X*, Ily* - x*11 ::; n + 1, I(y* - x*)xl ::; 1 when x E Fa U
... U Fn } ~ U; and
(2) Ilxll ::; n- l if n ;?: 1 and x E Fn.
The construction begins with the observation that since the closed ball
of radius 1 centered at x* inherits the same relative topologies from 'T
and 'T w " there are elements Xl, ... ,xno of X such that
{y* : y* E X*, Ily* - x*11 :so 1, I(y* - x*)xjl < 1 when j = 1, ... , na} ~ U.
Let Fa = {2XI, ... , 2x no }. Now suppose that mEN and that finite subsets
Fa, .. . , Fm - 1 of X have been found such that (1) and (2) are satisfied when
n = 0, ... ,m - 1. Suppose further that there is no finite subset Fm of X
such that (1) and (2) are satisfied when n = m. Let I be the collection
of all finite subsets of m- 1 B x , pre ordered by declaring that F =5 G when
F ~ G. For each F in I, let
Notice that each K(F) is nonempty and that K(F) :;2 K(G) when F =5 G.
As a straightforward consequence of the Banach-Alaoglu theorem and the
fact that 'T and 'Tw' induce the same relative topology on bounded subsets
of X*, each K(F) is weakly* compact. For each F in I, let x'F be an element
of K(F). Then (x'F)PE[ is a net in the weakly* compact set K(0) and has
the property that Xc
E K(F) whenever F =5 G, and so has a subnet with
a weak* limit Xo in n{ K(F) : F E I}. It follows that I(xo - x*)xl < 1
whenever X E m- 1 B x , so Ilxo - x*11 :so m. This means that
but this last set lies inside U by (1) when n = m-l. This is a contradiction,
since xC; E K(0) :;;;; X* \ U. It follows that there is a finite subset Fm of X
such that (1) and (2) are satisfied when n = m, which mcans that the
construction of the sequence Fa, F 1 , F2 , ... can be accomplished. It may be
assumed that each Fn is nonempty, since each empty Fn may be replaced
by {O}.
By (2), the elements of U~=a Fn can be listed as a sequence (x n ) that
converges to O. It follows from (1) that B(x*, (x n )) :;;;; U, which is enough
to establish that U is b-weakly* open.
2.7.4 Corollary. (J. Dieudonne, 1950 [63]). Let X be a normed space and
let A be a subset of X*. Then the following are equivalent.
(a) The set A is b-weakly* open.
(b) The set A n B is relatively weakly* open m B whenever B is a
bounded subset of X*.
(c) The set An tBx- is relatively weakly* open in tBx- whenever t > O.
PROOF. Let 'I be the collection of all subsets U of X* such that U n B is
relatively weakly* open in B whenever B is a bounded subset of X*. It is
easy to check that 'I is a topology for X*, that 'I and the weak* topology
of X* induce the same relative topology on each bounded subset of X*, and
that 'I' ~ 'I whenever 'I' is a topology for X* such that 'I' and the weak*
topology of X* induce the same relative topology on each bounded subset
of X*. By the preceding theorem, the topology 'I is the bounded weak*
topology of X*, which shows that (a) and (b) are equivalent. It is clear
that (b) '* (c). Also, for each bounded subset B of X* there is a positive
tB such that B ~ tBBx-, from which it easily follows that (c) '* (b).
2.7.5 Corollary. (J. Dieudonne, 1950 [63]). Let X be a normed space and
let A be a subset of X*. Then the following are equivalent.
(a) The set A is b-weakly* closed.
(b) The set A n B is relatively weakly* closed in B whenever B is a
bounded subset of X*.
(c) The set An tBx- is weakly* closed in X* whenever t > O.
(d) The set A contains the weak* limits of all bounded nets in A that
are weakly* convergent in X*.
240 2. The Weak and Weak* Topologies
PROOF. The equivalence of (a), (b), and (c) follows from the preceding
corollary and the fact that tBx- is weakly* compact, hence weakly* closed,
whenever t > O. The equivalence of (c) and (d) is clear.
contained in the next theorem: The dual spaces with respect to the two
topologies are the same if X is a Banach space. When interpreting this
theorem and its corollaries, keep in mind that a linear functional on X* is
weakly* continuous if and only if it has the form x* f-+ x*xo for some Xo
in X.
Theorem 2.7.8 has the main result of this section as an easy corollary.
Exercises
2.79 Prove that a bounded net in the dual space of a normed space is b-weakly*
Cauchy if and only if it is weakly* Cauchy.
2.82 Let X be a Banach space. Prove that the bounded weak* topology of X*
is metrizable if and only if X is finite-dimensional. Exercise 2.80 might
be helpful.
2.83 The purpose of this exercise is to show that if X is a Banach space and
1: is a topology for X* such that the relative topologies inherited by each
bounded subset of X* from 1: and the weak* topology 1:w ' are the same,
then it does not necessarily follow that 1: w * c;;:; 1:, even if 1: is a completely
regular locally convex topology. To this end, let X = Co, let Q be the
natural map from Co into c~*, let Y be the subspace of Co consisting of
the finitely nonzero sequences, and let 1: be the Q(Y) topology of c~.
(a) Prove that 1: is a completely regular locally convex topology for c~
such that 1: ~ 1:w *.
(b) Prove that the relative topologies inherited by each bounded subset
of c~ from 1: and 1:w * are the same. Exercise 2.67 might be helpful.
2.84 Let C be a convex subset of a normed space X.
(a) Prove that C is closed if and only if en tBx is closed whenever
t > o.
(b) Prove that C is weakly closed if and only if en tB x is weakly closed
whenever t > o.
These are, of course, the analogs for the norm and weak topologies of the
Krein-Smulian theorem on weakly* closed convex sets. Notice that these
analogs do not require X to be complete.
2.85 The hypotheses of both Theorem 2.7.8 and the Krein-Smulian theorem
on weakly* closed convex sets include the requirement that the normed
space X in question be a Banach space. The purpose of this exercise is
to show that this requirement cannot in general be omitted. Let X be an
incomplete normed spacej let Q be the natural map from X into X* j let Y
be the closure of Q(X) in X'*j let 1:w ', 1:bw ' , and 1:y be, respectively, the
weak*, bounded weak*, and Y topologies of X*j and let (X,:,.)*, (X;w')*'
and (X;')* be the respective dual spaces of X' under these topologies.
(a) Show that the relative topologies inherited by each bounded subset
of X from 1:w ' and 1:y are the same.
(b) Show that (X,:,.)* ~ (X;')* c;;:; (X;w')*. Conclude that the require-
ment in the hypotheses of Theorem 2.7.8 that the normed space be
complete is necessary.
244 2. The Weak and Weak* Topologies
(c) Find a convex subset C of X* that is not weakly* closed even though
en tBx' is weakly* closed whenever t > O.
2.86 The bounded weak topology. Let X be a normed space and let 'Ibw be the
collection of all subsets U of X such that Un B is relatively weakly open
in B whenever B is a bounded subset of X. Then 'Ibw is the bounded
weak topology of X. For convenience, properties related to this topology
will be called "b-weak" properties. Let 'Iw and 'In be the weak and norm
topologies of X respectively. Prove each of the following statements.
(a) The collection 'Ibw is a topology for X.
(b) 'Iw <:;; 'Ibw <:;; 'In.
(c) The relative topologies inherited by each bounded subset of X from
'Iw and 'Ibw are the same. Furthermore, if 'I is a topology for X
such that the relative topologies inherited by each bounded subset
of X from 'I", and 'I are the same, then 'I <:;; 'Ib,,,.
(d) A bounded net (xe in X is b-weakly convergent to some x in X if
and only if (Xa) is weakly convergent to x.
(e) A subset A of X is b-weakly open if and only if AntBx is relatively
weakly open in tEx whenever t > O.
(f) A subset A of X is b-weakly closed if and only if An B is relatively
weakly closed in B whenever B is a bounded subset of X, which
happens if and only if An tBx is weakly closed in X whenever
t > o.
2.87 Let X be a normed space. The purpose of this exercise is to study another
topology for X related to the bounded weak topology developed in Exer-
cise 2.86. For each x in X and each sequence (x;') in X* that converges
to 0, let
has its first n - 1 terms equal to m- 1 , its nth term equal to m, and all
the rest of its terms equal to O. Let A = {x(m, n) : m, n EN}.
(a) Show that A is b-weakly closed.
(b) Show that 0 EA. -~ Conclude that A is not 'I'D-closed, and t herefiore
that 'I'D ~ 'I'bw for co
This example is due to R. F. Wheeler [243J.
The preceding result is one of the three key ingredients in the proof of
the following important characterization of reflexivity, with the Banach-
Alaoglu theorem and Goldstine's theorem being the other two.
2.8.2 Theorem. A normed space is reflexive if and only if its closed unit
ball is weakly compact.
PROOF. Let Q be the natural map from a normed space X into X**.
It follows easily from Proposition 2.8.1, the Banach-Alaoglu theorem, and
246 2. The Weak and Weak* Topologies
Suppose that Yo E By. Select j so that Ilyo - yjll < ~Ml. Then
= k
Therefore max{ IYoxl : x E F M } 2: k. It follows that
for every y. in Y.
The following result is central to the proof of the Eberlein-Smulian theo-
rem to be given below. It is called Day's lemma because the argument used
to prove it is essentially the same as an argument used by Mahlon Day to
prove that every weakly sequentially compact subset of a normed space is
weakly closed; see [54, Theorem III.2.4].
(1) IIx*11 :::; 2max{ Ixxl : x E Fn} whenever X* E ({xi, ... ,x~}); and
(2) max{ IX~+lXI : x E Fn} < n~l.
To start the construction, let xi be any element of A., let Xl be a member
of Ex such that 21xixil 2:: Ilxill, and let Fl = {xd. Now suppose that a
248 2. The Weak and Weak* Topologies
As has already been mentioned, the hard parts of the proof of the follow-
ing theorem are the arguments that (c) =} (d) and that (c) =} (a). These
were, essentially, the respective contributions of Smulian and Eberlein to
this result. See [67, p. 466] for more details.
The next two results were previously obtained in optional Section 1.13
from a sequential characterization of reflexivity by R. C. James. See The-
orems 1.13.4, 1.13.5, and 1.13.8.
2.8.13 Lemma. Suppose that ':r is the norm or weak topology of a separa-
ble Banach space X and that K is a ':r-compact subset of X. Then co( K)
is ':r-compact.
PROOF. It may be assumed that K =I 0. Notice that K is weakly compact,
whether ':r is the norm or the weak topology. For the rest of this proof, the
topology of K is assumed to be its relative ':r topology whenever K is
treated as a topological space.
This proof makes extensive use of the standard identification of C(K)*
with rca(K); sec Example 1.10.6. Let K, be the identity function on K,
viewed as a map from the topological space K into the topological space
consisting of X with the topology ':r. For each x* in X*, the map x* K, is
in C(K), so JKx*K,d/1 exists for each /1 in rca(K).
Let /10 be a member of rca(K). It will be shown that there is a unique xl"o
in X such that x* x 1"0 = J K x* K, d/10 for each x* in X*. To this end, suppose
that (x~) is a sequence in X* that is weakly* convergent to some xo. For
each x in K and each positive integer n,
The following theorem was obtained by Krein under the additional as-
sumption that X is separable. The general case is due to Krein and Smulian.
Banach spaces that are the closed linear hull of a compact set or have a
weakly compact closed unit ball are obviously weakly compactly generated,
which immediately leads to two large classes of weakly compactly generated
normed spaces.
There are weakly compactly generated normed spaces that are neither
separable nor reflexive; see Exercise 2.98. As is shown in Exercise 2.99, the
space 00 is not weakly compactly generated, so there are Banach spaces
that lack this property.
See Mahlon Day's book [56, pp. 72-77] for more on weakly compactly
generated normed spaces. Klaus Floret's book [79] is a good source of fur-
ther information about weak compactness in general.
Exercises
2.89 Suppose that K is a weakly compact subset of the dual space of a normed
space X.
(a) Prove that the relative topologies induced on K by the weak and
weak* topologies of X are the same.
(b) Part (a) might seem to suggest that if x" E X", then there must
be an x in X such that x" and Qx agree on K, where Q is, as usual,
the natural map from X into X**. Show that this does not have to
be so, even if K is norm compact and X is a Banach space.
2.90 Prove that a subset of 1 is compact if and only if it is weakly compact.
Conclude that no infinite-dimensional subspace of 1 is reflexive.
2.91 Use the results of this section to prove that Co is not reflexive.
2.92 Let X be a normed space and let Q be the natural map from X into X*.
(a) Find conditions on X necessary and sufficient for Q(X) to be dense
in X**.
(b) Find conditions on X necessary and sufficient for Q(X) to be weakly
dense in X".
2.93 Find a compact subset K of a Banach space such that co(K) is not closed.
(A peek at Exercise 2.94 would not hurt.)
2.94 The goal of this exercise is to show that the conclusions of Lemma 2.8.13,
the Krein-Smulian weak compactness theorem, and Mazur's compactness
theorem can all fail if the normed space in question is not required to be
complete. Let X be the separable incomplete normed space consisting of
the vector space of finitely nonzero sequences with the = norm, and let
{en : n EN} be the collection of standard unit vectors of this space.
Show that the subset {D} U { n -1 en : n EN} of this space is compact,
but that its closed convex hull is not even weakly compact.
256 2. The Weak and Weak* Topologies
result extends to every weakly closed subset of a Banach space; that is, that
a weakly closed subset A of a Banach space X must be weakly compact if
sup{ Ix*xl : x E A} is attained whenever x* E X*. (A moment's thought
about the half-open interval (0, IJ in the Banach space R shows why A must
be required to be closed in some sense. Also, requiring A to be closed in
the norm sense is not enough; see Exercise 2.101.)
James proved Klee's conjecture in a 1964 paper [115J using some of the
same ideas he had used in his proof of the reflexivity theorem. In his 1972
paper containing the simplified proof of the reflexivity theorem, James
showed how that argument could be modified to obtain a simplified proof
of the more general weak compactness result.
The purpose of this section is to give a proof of Klee's conjecture follow-
ing James's lead from his 1972 paper. The general plan of attack is to use
Lemma 1.13.10 to prove the weak compactness result for nonempty, sep-
arable, weakly closed subsets of the closed unit ball of a Banach space in
Proposition 2.9.1, then use this proposition and Lemma 1.13.13 to obtain
the corresponding result for nonempty, balanced, weakly closed subsets of
the closed unit ball of a real Banach space in Proposition 2.9.2, and fi-
nally use Proposition 2.9.2 to obtain the general result and several useful
variations of it in Theorem 2.9.3. This program parallels the one used in
Section 1.13 to prove James's theorem, with Propositions 2.9.1 and 2.9.2
taking the place of Theorems 1.13.11 and 1.13.14 respectively. This is a full
agenda, so it would be best to get started.
that 1Illw really is a norm on W, with the key observation being that a
member of W that is zero on A must be zero on V. Let I: A - W* be
defined by the formula (J(x))(v*) = vx; that is, let I be the "natural
map" from A into W*. Notice for this that the definition of 11llw assures
that f(x) really is a bounded linear functional on W whenever x E A, and
that f(A) ~ Bw . Since V* is a separating family of linear functionals
on V, the function I is one-to-one. It is clear that a net (x",) in A is weakly
convergent in X to an x in A if and only if l(xaJ ~ f(x) in W*, from
which it follows that I is a homeomorphism from A with its relative weak
topology as a subset of X onto f(A) with its relative weak* topology as
a subset of W*. Since A is not weakly compact in X, the set I(A) is not
weakly* compact in W*. As a subset of the weakly* compact subset Bw.
of W* , the set I (A) must not be weakly* closed in W*. Fix an element F of
--w
f(A) \ f(A). Notice that there cannot be a v in V such that Fv = v*v
for each v* in W, for it is easy to see that such a v would have to be in
A W \ A, contradicting the fact that A is weakly closed. In particular, it
follows that F f= 0, and therefore that 11F11w > O. Since A ~ B v ,
sup{ IFv*1 : v* E Bv.} :::; sup{jFv*1 : v* E W, sup{ Iv*xl : x E A} :::; 1 }
= 1IFIIw"
from which it follows that F E V** and IlFllv" :::; 1IFIIw. Let Qv be the
natural map from V into V**. Since V is complete, the set Qv (V) is closed
in V**, and since F . Qv(V), it follows that d(F,Qv(V)) > 0, where d is
determined by the norm of V**. Let b.. be such that
0< b.. < d(F,Qv(v))
and let {an : n EN} be a countable dense subset of A. If n E Nand
aI, ... ,an+l are scalars, then
it follows that 0 < () < 1. Therefore f) and (x~) satisfy (b), which finishes
the proof that (a) => (b).
Now assume the existence of a f) and a sequence (x~) as in (b), and
let ({3n) be a sequence of positive reals summing to 1. Then the 0 and
the sequence (Y~) guaranteed by Lemma 1.13.10 do all that is required
of them in (c). In particular, notice that if x E A, then the facts that
Y~ E co( {xj : j 2:: n}) for each n and limn x~x = 0 together imply that
limn Y~x = O. It follows that (b) (c). *
Suppose that (c) holds. Fix a sequence ({3n) of positive scalars summing
to 1. Let 0 and (Y~) be as in (c) and let z* = 2:;:1 {3jyj. It will be shown
that sup{ Iz*xl : x E A} is not attained. Let Xo be an element of A and
let n be a positive integer such that IYixol < 00 whenever j > n. Then
: ; It
J=l
{3j Yi XO I ~E~j'Y;xo'
< su p { It {3jY;XI :
J=l
x E A} + of) f=
J=n+1
{3j
< 0 ( 1- f=
f) (3j )
j=n+1
+ of) f=
j=n+l
{3j
=0
The only use that will be made of this notation in the rest of this book is
in the next proposition. It was shown in Lemma 1.13.12 that L(x~) i- 0.
Notice also that liminfnx~x <::: x*x whenever x E L(x~) and x E X, since
liminfnx~x = -limsuPnx~(-x).
260 2. The Weak and Weak* Topologies
Then (fJn) is a sequence of positive scalars that sums to 1. Let a and (y~) be
as in (c). Let w* be any member of L(y~) and let z* = 2::;:1 fJj(yj - w*).
It will be shown that sup{ Iz*xl : x E A} is not attained. To this end,
suppose that Xu E A. Since lim infj yj Xo :S w* Xo and 0 :S a, there is an n
such that
*2.9 James's Weak Compactness Theorem 261
Since w*x ::; limsuPj yjx ::; 1 whenever x E Ex, it follows that Ilw* II ::; l.
Therefore
= L (3j(yj - w*)(xo)
00
Z*xo
j=l
n
L (3j(yj - +L
00
::; sup { It
J=l
{3j(yj - w*)(x)1 : x E f (3j
A} + (08 - 2tl.){3n+1 + 2J=n+2
< o{ f 1- 8 (3j)
j=n+1
+ (08 - 2tl.){3n+l +2 f {3j.
j=n+2
L
00
L {3j
00
= 0 - (08 - 2tl.)
j=n+2
<0
= su p { 1~{3j(y; - w*)(x)1 : x E A}
PROOF. The weak continuity of the functions mentioned in (b), (c), and (d)
assures that (a) implies each of (b), (c), and (d).
For the rest of this proof, it may be assumed that A is bounded, for
the hypotheses of (b), (c), and (d) all assure that x*(A) is a bounded set
of scalars for each x* in X*, and therefore, by Corollary 2.5.6, that A
is bounded. Clearly, it may then be assumed that A ~ B x. Let Xr be
the Banach space formed from X by using real scalars (which does not
imply any assumption that IF = C; if IF = lR, then Xr and X are the same
Banach space). Even when IF = C, straightforward arguments based on
Proposition 1.9.3 and Corollary 2.1.22 show that the weak topologies of Xr
and X are the same topology for the set underlying these Banach spaces.
Suppose that (b) holds. Let
B = n {x:
x*EX
X E X, Ix*xl : : ; sup{ Ix*yl : YEA} }.
from which it follows that sup{ lu*xl : x E A} is attained. This shows that
(d) => (c) and finishes the proof of the theorem.
Exercises
2.100 Either prove the following weak* analog of James's weak compactness
theorem or find a counterexample: Suppose that A is a nonempty weakly*
closed subset of the dual space of a Banach space X. Then A is weakly*
compact if and only if sup{ Ix'xl : x' E A} is attained whenever x E X.
2.101 Find a nonempty closed subset A of a Banach space X such that A is
not weakly compact even though the supremum of Ix"1 on A is attained
whenever x* E X*.
2.102 Without citing the 1971 example by James mentioned in this section, find
a nonempty weakly closed subset A of an incomplete normed space X such
that A is not weakly compact even though the supremum of Ix*1 on A is
attained whenever x* E X*. Exercise 2.94 might help.
2.103 Suppose that X is a Banach space having a subset A with nonempty inte-
rior such that the supremum of Ix"1 on A is attained whenever x" E X".
Prove that X is reflexive.
2.10.3 Example. In real t'~, the extreme points of the closed unit ball are
the four points (1, 1), that is, the vertices of the polygonal boundary of
the closed unit ball. In real Euclidean 2-space, the set of extreme points of
the closed unit ball is the entire unit sphere.
2.10.4 Example. Suppose that (an) is an element of the closed unit ball
of Co. Let no be a positive integer such that lano I < ~, and let (;3n) and Cl'n)
2.10 Extreme Points 265
be the members of Co such that (3n = "'In = On if n : no, but (3no = ono + !
and "'Ino = Ono - !.Then ((3n) and bn) are different elements of Beo
such that (on) = !((3n) + !bn), so (On) is not an extreme point of B cQ .
Therefore Bco has no extreme points, so not even in Banach spaces are
bounded nonempty closed convex sets guaranteed to have extreme points.
is an extremal subset of C.
PROOF. Parts (a) and (b) follow easily from Definition 2.10.1. For (c), sup-
pose that C is compact and that x* E X*. Let s = max{Rex*x: x E C}
and let D s = {x : x E C, Re x* x = s }, a nonempty closed convex subset
of C. If x and yare members of C such that tx + (1- t)y E Ds for some t
in (0,1), then tRex*x + (1- t) Rex*y = s, which together with the defi-
nition of s implies that Re x* x = Re x* y = s and thus that x, y E D s' The
set D s is therefore an extremal subset of C.
FIGURE 2.2. A polygon P in the Euclidean plane, and its closed convex hull.
2.10.11 Example. As was shown in Example 2.lO.4, the closed unit ball
of Co has no extreme points, so Co is not isometrically isomorphic to the dual
space of any normed space. Similarly, the space L1 [0,1] is not isometrically
isomorphic to the dual space of any normed space, since it can be shown
that E LdO ,lj has no extreme points. See Exercise 2.110.
2.10.14 Lemma. (N. Bourbaki, 1953 [34, p. 80]). Suppose that K 1, ... , Kn
are compact convex subsets of a TVS. Then CO(K1 U U Kn) is compact.
PROOF. Each K j may be assumed to be nonempty. Let (2:7=1 t;a)x;"'))O:EI
be a net in co(K l U ... U Kn), where each term of the net is being writ-
ten as a sum of the type described in Lemma 2.10.13. Since each K j is
compact, as is the interval [0,1]' there is a subnet (2:7=1 t;(3)xt)){3EJ of
(2:j=l t;"')x;a))aEI such that, for each j, there are elements tj of [0,1]
and Xj of K j such that t;(3) ---> tj and xJ(3) ---> Xj. It then follows from the
continuity of the vector space operations that 2:7=1 tt)x;{3) ---> 2:7=1 tjXj.
Also, the continuity of the map (a1'.'.' an) f---+ 2:7=1 aj from Euclidean
n-space into IF assures that 2:7=1 tj = 1, so 2:j=1 tjXj E CO(K1 u .. U Kn)
Since every net in CO(K1 U ... U Kn) has a convergent subnet with a limit
in CO(K1 U ... U Kn), the set CO(K1 U ... U Kn) is compact.
Exercises
2.107 Let x be an element of a nonempty closed convex subset C of a Hausdorff
TVS. Show that x is an extreme point of C if and only if it has this
property: Whenever XI,X2 E C and x = ~(XI + X2), it follows that Xl =
X2 = X.
270 2. The Weak and Weak* Topologies
e
2.108 Identify all of the extreme points of the closed unit ball of 1 , then show
that Be, is the closed convex hull of its set of extreme points. Do not use
any results from this section in your arguments.
e
2.109 Identify all of the extreme points of the closed unit ball of ao , then show
that Be oo is the closed convex hull of its set of extreme points. Do not use
any results from this section in your arguments.
2.111 Prove that real 0[0,1] is not isometrically isomorphic to the dual space
of any normed space.
2.112 Let 0 be a nonempty closed convex subset of a Hausdorff TVS X. An
element Xe of 0 is an exposed point of 0 if there is an x in X' such that
Re x is bounded from above on 0 and attains its supremum on 0 at Xe
and only at Xe.
(a) Show that an exposed point of 0 must be an extreme point of O.
(b) Find a Banach space Xo such that some extreme point of Bxo is
not an exposed point of Bxo' (This can be done by constructing an
appropriate norm on ne. Exercise 1.37 might be helpful.)
Exposed points were introduced by S. Straszewicz [229] in 1935. Two good
sources for more on exposed points and related objects are Klee's 1958
paper [136] and Day's book [56, pp. 105~1O6].
2.113 Let P be a polygon in the Euclidean plane, not assumed to be convex,
whose sides intersect only at its vertices and whose vertices are each com-
mon to exactly two sides, and let K be the compact region in the plane
consisting of P together with the bounded component of its complement.
In the discussion of such polygonal regions that precedes Lemma 2.10.12,
it is claimed that "co(K) (which is the same as co(K) in this case) is a
closed convex polygonal region whose vertices are a subset of the vertices
of K." Prove this. Base your arguments on elementary principles, without
using the Krein-Milman theorem or Theorem 2.10.15 in any way.
2.114 Let A be a nonempty subset of a Hausdorff LeS X such that co(A) is
compact, let E be the set of extreme points of co(A), and let B be a
subset of A. Show that the following are equivalent.
(a) co(B) = co(A).
(b) E ~ B.
(c) sup{Rex'x: x E B} = sup{Rex'x: x E A} whenever x EX".
-ltOoll p d/1>1
= 11/1>11
= /1>+([0, 1]) + /1>_([0, 1]),
272 2. The Weak and Weak* Topologies
from which it follows that either /.L+ or /.L- is the zero measure, that is, that
/.L is either a nonnegative or a nonpositive measure.
Now suppose that P is a nonconstant member of Sx and that F is the
nonempty finite subset of [0, 1] on which Ipi is 1. Let /.L be a member of X*
such that I/.L(p) I = II/.LII If I/.LI([O, 1] \ F) > 0, then
II/.LII =! r J[D,l]
Pd/.L!
:S r
J10,1]\F
Ipi dl/.LI + rdl/.LI
JF
< r
J10,1]\F
dl/.LI + rdl/.LI
JF
= II/.LII,
a contradiction that shows that 1111([0,1] \ F) = 0, that is, that /.L is finitely
supported.
Thus, a member of X* that is norm-attaining must be nonnegative, non-
positive, or finitely supported. Let ..\ be Lebesgue measure on [0,1] and
let /.La be the member of X* given by the formula
/.Lo(A) = ..\(A n [0, 1/2]) - ..\(A n [1/2, 1]).
Then II/.Lo I = 1. If 11 is a member of X* that is finitely supported, then it
is easy to see that II/.L - /.La I = 11/111 + 11/10 II ~ 1, whereas if 11 is a member
of X* that is nonnegative or non positive , then it is equally easy to see
that II/.L - /.Loll ~ 1/2. It follows that the open ball of radius 1/2 centered
at /.La contains no norm-attaining members of X*, and therefore that the
collection of norm-attaining members of X* is not dense in X*.
See Exercise 2.117 for other equivalent definitions of wedges and cones
that are sometimes used. Notice that if K is a wedge, then OK s;:; K, so
wedges always contain O.
2.11.5 Definition. (E. Bishop and R. R. Phelps, 1963 [29]). Suppose that
X is a TVS such that X =J {O}. Suppose further that Xo is an element of
a subset A of X and that K is a cone in X with nonempty interior such
that An (xo + K) = {xo}. Then K is a support cone for A, the point Xo is
a conical support point of A, and Xo + K supports A at Xo.
The reason for defining conical support points only in nontrivial topo-
logical vector spaces is that without this restriction, the convex subset {O}
of the normed space {O} would have 0 as a conical support point but not
as a support point. This would be inconsistent with the behavior of conical
support points of convex sets in larger topological vector spaces, as the
following proposition shows.
that there is a nonzero Yo in X such that Yo and -Yo are both in K, which
contradicts the fact that K n (-K) = {O}. It follows that Xo :. (xo + K)O,
and therefore that en (xo + K)O = 0. By Eidelheit's separation theorem,
there is an x* in X* and a real number s such that Rex*x ::; s for each x
in C, Rex*x 2:: s for each x in xo+K, and Rex*x > s for each x in (xo+K)o.
It follows that x* =I- 0 and that Rex*xo = s = sup{ Rex*x : x E C}, and
therefore that x* supports C at Xo.
Some temporary notation is now needed that will not apply outside this
section. Suppose that X is a normed space, that x* E Sx', and that t > 1.
Then
K(x*, t) = {x: X E X, Ilxll ::; Re x* (tx) }.
Since Ilx*11 = 1, there is an Xo in X such that Rex*(txo) > Ilxoll, which
assures that {x : x E X, Ilxll < Rex*(tx)} is a nonempty open set inside
K(x*, t). It then follows easily that K(x*, t) is a closed cone with nonempty
interior.
2.11.7 Lemma. (E. Bishop and R. R. Phelps, 1963 [29]). Suppose that
x is an element of a complete subset A of a normed space X, that x* is
a member of Sx' such that Rex' is bounded from above on A, and that
t> 1. Then there is an Xo in A such that Xo E x+K(x*, t) and xo+K(x*, t)
supports A at xo.
PROOF. Let B = A n (x + K(x*, t)). Notice that an element z of A is in B
if and only if liz - xii::; Rex*(t(z - x)). The set B is a closed subset of
the complete set A and so is itself complete. Since K(x*, t) is a cone, it
follows easily that the relation on B defined by declaring that y ::::; z when
z - y E K(x*, t) is a partial order for B. It is also clear that x E B and that
x ::::; z whenever z E B.
Suppose that ~ is a nonempty chain in B. Then ~ can be used as the
index set for a net, so the restriction of Re x* to <!: is a net in R Since
Re x*y ::; Re x* z whenever y ::::; z, the boundedness from above of Re x*
on A implies that the net Re x* II!: converges and therefore is Cauchy. Since
liz - yll ::; Re x* (t(z - y)) whenever y, z E Band y ::::; z, the net It (or, more
properly, the identity function on <!:, viewed as a net) is a Cauchy net in B
and so converges to some u in B. Since K(x*, t) is closed, it follows that
u - y E K(x*, t) for each y in ~, that is, that u is an upper bound for ~
in B.
By Zorn's lemma, the set B contains a maximal element Xo. Then Xo E
x + K(x', t) since x ::::; Xo. Suppose that yEA n (xo + K(x*, t)). Then
y E Xo + K(x*, t) ~ x + K(x*, t) + K(x*, t) = x + K(x*, t),
so Y E Band Xo ::::; y. It follows from the maximality of Xo that y = Xo, and
therefore that An (xo + K(x*, t)) = {xu}. The set Xo + K(x*, t) therefore
supports A at Xo
*2.11 Support Points and Subreflexivity 275
The preceding theorem, with a little help from Proposition 2.11.6, im-
mediately yields the following result when the Banach space is not {O}. If
the space is {O}, then the result is trivially true since both subsets of the
space have empty boundaries.
and consider (1- s)x+sy. The Bishop-Phelps support point theorem there-
fore implies that the answer to Klee's question is yes.
Let e be a closed convex subset of a Banach space X. Then in addition
to the support points of e being dense in the boundary of e, it turns out
that the support functionals for e must be dense in X* provided that a few
restrictions are placed on e and X. To see what those restrictions might
be, first notice that e cannot be empty or all of X, for then e would have
no support functionals at all. It is not enough of a restriction on e just
to require that it be nonempty and not all of X; see Exercise 2.119. As
will be shown, it is enough to require e to be nonempty and bounded,
provided that X =f {O} so that e cannot be X. This will follow easily
from a separation theorem due to Bishop and Phelps. The proof of this
separation theorem is based on Lemma 2.11.11 below, which is in turn a
consequence of the following result, which says, roughly speaking, that if
276 2. The Weak and Weak* Topologies
By Lemma 2.11.10, either Ilxi - x 211 :S for Ilxi + x211 :S E. However, the
second inequality cannot hold, since Ilxi + xiii 2: (xi + xi)(xo) > E, which
completes the proof.
2.11.12 Theorem. (E. Bishop and R. R. Phelps, 1963 [29]). Suppose that
C and A are nonempty subsets of a Banach space X such that C is closed
and convex and A is bounded. Suppose further that E > 0, that xi E Sx.,
and that
PROOF. It may be assumed that 0 < E < 1 and that X is a real Banach
space. Let
Suppose that X and Yare normed spaces. As one would expect, a mem-
ber T of E(X, Y) is called norm-attaining if there is an x in Ex such that
IITxl1 = IITII In light of the Bishop-Phelps subrefiexivity theorem, it is
*2.11 Support Points and Subreflexivity 279
natural to ask the following question that Bishop and Phelps posed in [28]:
What conditions on two Banach spaces X and Y assure that the collection
of norm-attaining members of B(X, Y) is dense in B(X, Y)? There must
be some additional conditions imposed on at least one of the spaces; in
particular, in a 1963 paper Joram Lindenstrauss [153] showed that there
is a Banach space X such that the set of norm-attainers in B(X) is not
dense in B(X). Partial answers to the question have emerged from research
done on a property for Banach spaces called the Radon-Nikodym property,
in particular from work by Jean Bourgain [35) published in 1977. See the
books by Richard Bourgin [37] and by Joseph Diestel and J. Jerry Uhl [59]
for discussions of the Radon-Nikodym property and its relationship to the
question raised by Bishop and Phelps.
Since Bishop and Phelps showed that the norm-attainers are dense in
B(X, Y) for every Banach space X when Y = IF, it is particularly nat-
ural to ask what conditions on a Banach space Y would assure that the
norm-attainers are dense in B(X, Y) for every Banach space X. A good
starting point for anyone interested in that problem is Timothy Gowers's
1990 paper [90], in which he shows that each space 1!p such that 1 < p < 00
lacks this property.
Exercises
2.115 Let X be the vector space of finitely nonzero sequences, equipped with
the e2 norm. Show that X is an incomplete subreflexive normed space.
(c) Prove that K is a cone if and only if it is a wedge with the property
that x = whenever x, -x E K.
2.118 The preceding exercise is useful, though not at all essential, for this one.
Suppose that P is a wedge in a real vector space X. Define a relation ~
on X by declaring that x ~ y when y - x E P.
(a) Prove that ~ is a preorder on X.
(b) Prove that if x, y, z E X, t ~ 0, and x ~ y, then x +z ~ y +z and
tx ~ ty.
(c) Prove that ~ is a partial order if and only if P is a cone.
280 2. The Weak and Weak* Topologies
The pair (X,~) is called an ordered vector space, the members of P are
called the positive elements of X, and P itself is called the positive wedge
(or positive cone, if appropriate) of X.
2.119 Find a closed convex subset C of a Banach space X such that the collection
of support functionals for C is nonempty but not dense in X'. (It can be
done in Il~?)
2.120 Call a subset A of a normed space X orthodox iffor each x in X \ A there
is a norm-attaining x' in X such that Rex'x > sup{Rex'y: yEA}.
Prove that a normed space is su breflexive if and only if each of its bounded
closed convex subsets is orthodox.
2.121 Prove the following improvement of Theorem 2.2.28 for Banach spaces:
Let K and C be disjoint nonempty convex subsets of a Banach space X
such that K is compact and C is closed. Then there is a member x'
of X such that inf{Rex'x : x E C} is attained by Rex' on C and
max{Rex'x: x E K} < min{Rex'x: x E C}.
2.122 Let X be a normed space such that X f:. {O}. If t E lR and x is a nonzero
member of X, then the subset of X defined by the formula
Linear operators have already received quite a bit of attention in this book,
primarily as tools for probing the structure of normed spaces. The purpose
of this chapter is to reverse that emphasis temporarily by studying linear
operators between normed spaces as interesting objects in their own right,
with the properties of normed spaces obtained in the first two chapters used
as the tools for the study. Almost all of the attention will be on bounded
linear operators, though there is also an interesting theory of unbounded
ones; see, for example, Chapter 13 of [200] or Chapter VII, Section 9 of [67].
The adjoint of an isomorphism from one normed space onto another was
introduced in Section 1.10, primarily to be able to prove that normed spaces
that are isomorphic or isometrically isomorphic have dual spaces bearing
that same relationship to each other. The first order of business is to extend
the notion of adjoint operators and learn more about them.
3.1.1 Definition. Suppose that X and Yare vector spaces and that
T E L(X, Y). Then the algebraic adjoint of T is the linear operator T#
284 3. Linear Operators
from y# into X# given by the formula T#(y#) = y#T. That is, the al-
gebraic adjoint T# is defined by letting (x, T#y#) = (Tx, y#) whenever
x E X and y# E y#.
3.1.2 Proposition. Suppose that X and Yare normed spaces and that
T E L(X, Y). Then T is bounded if and only if T#(Y*) C;;;; X*. If T
is bounded, then the restriction T* of T# to Y*, viewed as a member
of L(Y*,X*), is itself bounded, and IIT*II = IITII.
PROOF. The fact that T is bounded if and only if T#(Y*) C;;;; X* is just
a restatement of Proposition 2.5.10. Now suppose that T is bounded, so
that T# (y*) C;;;; X*, and let the restriction T* of T# to Y be viewed as
a member of L(Y*, X*). An application of Proposition 1.10.11 (a) shows
that
By Proposition 1.10.11 (b), the fact that the second supremum in the above
equality is finite assures that T* is bounded and has its norm equal to that
supremum, and therefore to IITII.
3.1.3 Definition. Suppose that X and Yare normed spaces and that
T E B(X, Y). Then the (normed-space) adjoint of T is the restriction
of T# to Y*, that is, the linear operator T* from y* to X* given by the
formula T*(y*) = y*T.
in the obvious way, and B(lF P, lF q ) (p, q > 0) to be identified with the col-
lection of all p x q matrices of scalars, with the action of a member T of
B(lFP, lF q ) on an element x of lF P given by the formula Tx = Tt . x, where
Tt is the transpose of the matrix T and represents matrix multiplication.
In particular, when IF is identified with lF 1 in the obvious way, a member x*
of (lFP)* acts on a member x of lFP through the formula x*x = (x*)t . x.
Suppose that m and n are positive integers and that T E B(lFm,lFn). If
x E lF m and y* E (lFn)*, then
so T*y* = (Tt)t . y*. It follows that T* is, as a matrix, equal to Tt. The
notion of an adjoint operator is therefore, in a sense, a generalization of
the notion of the transpose of a matrix.
The verification of the following proposition is easy. The corollary then
follows after a glance at Proposition 3.1.2.
3.1.5 Corollary. If X and Yare normed spaces, then the map T ....... T*
is an isometric isomorphism from B(X, Y) into B(Y, X*).
The preceding two examples might give one the idea that the adjoint of
a one-to-one bounded linear operator between normed spaces must itself
be one-to-one. As the next example shows, this is not the case. The actual
relationships between such properties as being one-to-one and being onto
for bounded linear operators and their adjoints will be settled by Theorems
3.1.17,3.1.18, and 3.1.22.
which implies that Q-;'Q X' is the identity map on X* and therefore that
Q-;' maps X*** onto X*. If Q-;' were also one-to-one, then Qx' would have
to map X* onto X***, contradicting the fact that X* is not reflexive. The
isometric isomorphism Qx therefore does not have a one-to-one adjoint.
PROOF. Suppose first that T E B(X, Y). Let (y~) be a net in y* that is
weakly* convergent to some y'. For each x in X,
3.1.13 Proposition. Suppose that X and Yare normed spaces and that
T E B(X, Y). Let Qx and Qy be the natural maps from X and Y
into their respective second dual spaces. Then T**Qx(X) ~ Qy(Y) and
Q}/T**Qx = T.
PROOF. Fix an x in X. First notice that whenever y* E Y* ,
3.1.14 Corollary. Suppose that X and Y are normed spaces and that
T E R(X, Y). Let Qx and Qy be the natural maps from X and Y into
their respective second dual spaces.
(a) The operator T is one-to-one if and only if the restriction of T**
to Qx(X) is one-to-one.
(b) The operator T maps X onto Y if and only if T** maps Qx(X)
onto Qy(Y).
PROOF. The preceding proposition implies that T**Qx = QyT, from
which the corollary follows easily.
3.1.16 Lemma. Suppose that X and Yare normed spaces and that T E
B(X, Y). Then ker(T) =.L (T*(Y*)) and ker(T*) = (T(X)).L.
PROOF. Since y* is a separating family of functions for Y, an element x
of X is in ker(T) if and only if
for each y* in Y*, which is the same as saying that x E .L(T*(Y*)). Simi-
larly, an element y* of y* is in ker(T*) if and only if (3.1) holds for each x
in X, which is equivalent to requiring that y* be in (T(X)).L.
290 3. Linear Operators
3.1.17 Theorem. Suppose that X and Yare normed spaces and that
T E B(X,Y).
(a) The operator T is one-to-one if and only if T*(Y*) is weakly* dense
in X*.
(b) The operator T* is one-to-one if and only if T(X) is dense in Y.
PROOF. By Lemma 3.1.16 and Propositions 2.6.6 (c) and 1.10.15 (c),
(3.2)
and
(3.3)
Since T is one-to-one if and only if ker(T) = {O}, which happens if and only
if (ker(T)).L = X*, it follows from (3.2) that T is one-to-one if and only
if X* = T*(Y*) w', proving (a). Part (b) follows from a similar argument
based on (3.3).
Here is the promised extension of the fact that square matrices are in-
vertible exactly when their transposes are. Part of this result was previ-
ously obtained as Theorem 1.10.12. Unlike Theorem 1.10.12 and Propo-
sition 3.1.15, this theorem has a completeness hypothesis that cannot in
general be removed. See Exercise 3.8.
See Theorem 1. 7.13 for the basic properties of the map S in the following
lemma.
3.1.20 Lemma. Suppose that X and Y are normed spaces and that T E
B(X, Y). Let rr be the quotient map from X onto XI ker(T), let S be
the bounded linear operator from XI ker(T) into Y such that T = Srr,
let Z = T(X), and let R be T viewed as a member of B(X, Z).
(a) The set S(XI ker(T) is closed if and only if T(X) is closed.
(b) The set S* (Y*) is closed if and only if T* (Y*) is closed.
292 3. Linear Operators
(c) The set S*(Y*) is weakly* closed if and only if T*(Y*) is weakly*
closed.
(d) The set R(X) is closed if and only if T(X) is closed.
(e) The set R*(Z*) is closed if and only if T*(Y*) is closed.
(f) The set W(Z*) is weakly* closed if and only if T*(Y*) is weakly*
closed.
PROOF. Part (a) follows immediately from the fact that the range of S is the
same as that of T, while (b) and (c) hold because T* = 7f* S* and 7f* is both
an isometric isomorphism and a weak*-to-relative-weak* homeomorphism
from (XIM)* onto the weakly* closed subspace MJ. of X*. Part (d) is
obvious. For (e) and (f), notice that if y* E y* and z* is the member of Z*
formed by restricting y* to Z, then
3.1.21 Theorem. Suppose that X and Yare Banach spaces and that
T E B(X, Y). Then the following are equivalent.
(a) The set T(X) is closed.
(b) The set T*(Y*) is closed.
(c) The set T*(Y*) is weakly* closed.
PROOF. Let 7f be the quotient map from X onto XI ker(T) and let S be the
bounded linear operator from XI ker(T) into Y such that T = S7r. Then S
is one-to-one since its kernel contains only the zero element of XI ker(T).
It is a consequence of this and parts (a), (b), and (c) of Lemma 3.1.20 that
the theorem is true if it holds under the additional hypothesis that T is
one-to-one. It may therefore be assumed that T is one-to-one. Now suppose
that the theorem holds under the additional hypothesis that T(X) = Y.
The general case then follows from parts (d), (e), and (f) of Lemma 3.1. 20,
so it may also be assumed that T(X) is dense in Y. By Theorem 3.1.17,
the operator T* is one-to-one and T* (Y*) is weakly* dense in X*.
Suppose that T(X) is closed and therefore that T(X) = Y. Then Corol-
lary 1.6.6 implies that T is an isomorphism from X onto Y, so it follows
from Theorem 1.10.12 that T* is an isomorphism from Y* onto X* a.nd
therefore tha.t T*(Y*) is closed. This shows that (a) => (b).
Suppose next that T* (Y*) is weakly* closed. Then T* (Y*) = X*, so T*
is an isomorphism from Y* onto X*. It follows from Theorem 3.1.18 that
T is an isomorphism from X onto Y and therefore that T(X) is closed,
which shows that (c) => (a).
3.1 Adjoint Operators 293
Finally, suppose that T*(Y*) is closed. Since Y* and T*(Y*) are both
Banach spaces, the operator T* is an isomorphism from y* onto T* (Y*).
Suppose that (x:) is a net in T*(Y*) n Bx" that is weakly* convergent
to some x* in X*. Then x* E Bx" since B x " is weakly* closed. Let
y: = (T*)-l(X:) for each a. Then (y:) is a bounded net and so, by the
Banach-Alaoglu theorem, has a subnet (y~) that is weakly* convergent to
some y* in Y*. It follows from the weak*-to-weak* continuity of T* that
x~ = T*y~ }1!," T*y*, so T*y* = x*. Since x* E T*(Y*) n B x ", the set
T* (Y*) n B x" is weakly* closed, which by Corollary 2.7.12 of the Krein-
Smulian theorem on weakly* closed convex sets implies that T*(Y*) is
itself weakly* closed. This shows that (b) => (c) and finishes the proof of
the theorem.
3.1.22 Theorem. Suppose that X and Yare Banach spaces and that
T E B(X,Y).
(a) The operator T maps X onto Y if and only if T* is an isomorphism
from Y* onto a subspace of X*.
(b) The operator T* maps y* onto X* if and only if T is an isomorphism
from X onto a subspace of Y.
PROOF. The operator T maps X onto Y if and only if T(X) is both closed
and dense in Y, which by Theorems 3.1.17 and 3.1.21 is equivalent to
T* being one-to-one and having closed range, which by Corollary 1.6.6 is
equivalent to T* being an isomorphism from y* onto a subspace of X*. The
operator T* maps y* onto X* if and only if T* (Y*) is both weakly* closed
and weakly* dense in X*, which is equivalent to T being one-to-one and
having closed range, which is in turn equivalent to T being an isomorphism
from X onto a subspace of Y.
See [205] for a discussion of adjoints in settings more general than that
of bounded linear operators between normed spaces.
Exercises
3.1 Suppose that x' is a bounded linear functional on a normed space X.
Describe the adjoint of x*.
3.2 Suppose that X is Co or R.p such that 1 < p < 00 and that mEN. Let T m
and T -m be the right-shift and left-shift operators defined by the formulas
and
3.3 Show that there is no continuous map from B(JF, co) onto B(c~, JF').
3.4 Suppose that X and Yare normed spaces.
(a) Prove that if X t {O} and Y is not reflexive, then some member
of B(Y', X') is not weak*-to-weak* continuous.
(b) Prove that the isometric isomorphism T f-+ T' maps B(X, Y) onto
B(Y', X') if and only if either X = {O} or Y is reflexive.
3.5 Prove that every weakly* closed subspace of the dual space of a normed
space is the range of the adjoint of some bounded linear operator. That
is, prove that if X is a normed space and N is a weakly* closed subspace
of X', then there is a normed space Y and a T in B(X, Y) such that
T*(Y*) = N.
3.6 Prove that every weakly* closed subspace of the dual space of a normed
space is the kernel of the adjoint of some bounded linear operator. That
is, prove that if X is a normed space and N is a weakly* closed subspace
of X*, then there is a normed space Y and a T in B(Y, X) such that
ker(T*) = N. Conclude that a subspace of the dual space of a normed
space is weakly* closed if and only if it is the kernel of the adjoint of some
bounded linear operator.
3.7 (a) Suppose that X and Y are Banach spaces, that T E B(X, Y), and
that Y* is an isomorphism from Y* onto X*. Prove that T' is also
a weak*-to-weak* homeomorphism from Y* onto X*.
(b) Give an example to show that the conclusion of (a) can fail if X
and Yare only assumed to be normed spaces.
3.8 (a) If the normed space Y in Theorem 3.1.18 is incomplete, then the the-
orem really says nothing about T itself that was not already shown
in Chapter 1, but instead amounts only to a statement about T*.
What is that statement?
(b) Show that Theorem 3.1.18 would not be true if its statement were
amended to require Y instead of X to be complete.
3.9 Show that the conclusion of Theorem 3.1.21 does not have to hold if the
completeness hypothesis is dropped for either X or Y.
3.10 Suppose that P is a property defined for Banach spaces such that
(1) if a Banach space X has property P, then so does every Banach
space isomorphic to X;
(2) if a Banach space has property P, then so does every closed subspace
of the space; and
(3) a Banach space ha::; property P if and only if its dual space ha..<;
property P.
For example, reflexivity satisfies these conditions. Prove that if X and Y
are Banach spaces such that X has property P and there is a bounded
linear operator from X onto Y, then Y has property P. Conclude that if
a Banach space X has property P and M is a closed subspace of X, then
X/M has property P.
3.2 Projections and Complemented Subspaces 295
and
from which it readily follows that the range and kernel of P are both closed
subspaces of X and that X is the internal direct sum of these two subspaces;
see Exercise 3.14. It is easy to check that P is bounded if and only if ((3n)
is bounded, and that IIPII = 1 + 11((3n)lloo whenever ((3n) is bounded.
Notice that all of this is just as valid if X is all of C[XJ provided that ((3n)
is required to be bounded to assure that P takes its values in 00' in which
case P must be a bounded projection.
which implies that T(Tx) = Tx since the collection of all linear functionals
on a vector space is always a separating family for that vector space. Thus,
the operator T is a projection. The remaining claims in the theorem are
proved by similar arguments.
As usual, the symbol I in the next several results represents the identity
operator on the vector space in question.
algebraic internal direct sum of the ranges (or kernels) of PI and P2 Thus, a
decomposition of X as a sum of two algebraically complementary subspaces
corresponds to a decomposition of I as a sum of two projections; that is, to
a decomposition of I as a sum of two operators that are "smaller identities"
in the sense that each acts as the identity operator on its range.
Corollary 3.2.12 suggests the following question: If M and N are com-
plementary subspaces of a Hausdorff TVS X, must the projection with
range M and kernel N be continuous? The answer does not have to be yes,
even when X is a normed space.
The idea used to prove that (b) =? (a) in the preceding theorem can also
be used to prove results about the extensibility of more general continuous
functions. See Exercise 3.22.
The closed subs paces {O} and X of a Hausdorff TVS X are certainly
complemented, since X is the internal direct sum of the two. The follow-
ing theorem says that closed subspaces of X that differ by only a finite
number of dimensions from X or, if X is locally convex, from {O}, are also
complemented.
Recall that if M is a subspace of a vector space X, then the codimension
of M in X is the dimension of the quotient vector space XI M.
Therefore
and so q ::; pIIY*II. Consequently, there are only finitely many index ele-
ments 0: such that Iy*(xa: + co)1 2: p-l. Since p was an arbitrary positive
integer, it follows that there are only countably many index elements 0:
such that y*(xa: + eo) # O.
Now suppose that C is a countable subset of (foo/co)*. It follows that
there are only countably many members 0: of I such that z*(xa: + co) # 0
for some z* in C. Since I is uncountable, there must be two different mem-
bers 0:1,0:2 of I such that z*(xal +co) = Z*(X02 +co) = 0 for each z* in C,
which shows that C is not a separating family for eX) / Co. The space f= / Co
therefore lacks property P, which is the desired contradiction.
3.2.21 Corollary. No bounded linear operator from foo onto Co maps each
element of Co to itself
3.2.22 Corollary. Let Q be the natural map from Co into cQ*. Then Q(co)
is not complemented in co*.
PROOF. Suppose Q( co) were complemented in co'. Then there would be
a bounded projection P in coo with range Q(co). Let S: foo -+ fi' and
T: f 1 -+ Co be the standard identifying isometric isomorphisms, and let
Po = S-lT* P(T*)-lS. Routine verifications show that Po is a projection
in foo with range co---a contradiction.
3.2 Projections and Complemented Subspaces 303
The preceding corollary might lead one to ask if it is ever possible for
the natural image of a nonreflexive Banach space in it second dual to be
complemented. It is quite possible, as the next result shows.
Exercises
3.14 Provide the missing details in Example 3.2.5 by showing that the range
and kernel of P are both closed subspaces of X, that X is the internal
direct sum of these two subspaces, that P is bounded if and only if CBn}
is bounded, and that IIPII = 1 + II (,Bn) 1100 whenever (,Bn) is bounded.
304 3. Linear Operators
3.15 For each T in L(JF 2, JF2), let the determinant of T be the determinant of
the matrix meT) of T with respect to the standard basis for JF2, and let
the trace of T be the trace of meT), that is, the sum of the elements
along the main diagonal of meT). Prove that a member T of L(JF 2, ]F2) is
a projection if and only if it satisfies one of the following three conditions.
(a) T = f.
(b) T = O.
(c) The determinant of Tis 0 and the trace of Tis 1.
3.16 (a) Prove that if M is a complemented subspace of a Banach space, then
all subspaces complementary to M are isomorphic to one another.
(b) Give an example of a complemented subspace of a Banach space
that has infinitely many different subs paces complementary to it.
3.17 (a) Suppose that P is a bounded projection in a normed space X. Prove
that P(X) = .L(ker(P") and P(X) = (ker(P)l..
(b) Suppose that M and N are complementary subspaces of a Banach
space X. Prove that M.L and N.L are complementary in X.
(c) Give an example to show that the conclusion of (b) need not hold if
X is only assumed to be a normed space.
3.18 Suppose that P is a bounded projection in a normed space X. Prove that
P(X) has finite dimension n if and only if P*(X) has finite dimension n.
(One proof of this uses Exercise 3.17 (a) and either Exercise 1.81 or the
first isomorphism theorem for vector spaces; see, for example, [105, p. 397]
for this theorem.)
3.19 This exercise uses the material of optional Section 2.3. Show that Theo-
rem 3.2.14 and Corollary 3.2.15 have natural extensions to F-spaces.
3.20 Suppose that 0 ::; p < 1. Prove that no finite-dimensional subspace
of Lp[O, 1] other than {O} is complemented. Exercise 3.19 may be helpfuL
3.21 This exercise improves the conclusion of Exercise 1.141. Suppose that X
is a Banach space and that there is a bounded linear operator from X
onto 1. Prove that X has a complemented subspace isomorphic to 1.
3.22 Suppose that M is a subspace of a Banach space X and that M is com-
plemented in X.
(a) Prove that if M is closed, then the following holds: For every topo-
logical space Z and every continuous function I from Minto Z,
there is a continuous function Ix from X into Z that agrees with f
onM.
(b) Show by example that the conclusion in (a) does not necessarily
hold if M is not required to be closed, even when Z is assumed to
be a normed space and f is assumed to be linear.
3.23 Suppose that M is a closed subspace of a Banach space. It follows from
Exercise 3.17 (b) that if M is complemented, then so is M.L. The main
goal of this exercise is to prove that the converse is not in general true.
3.3 Banach Algebras and Spectra 305
(a) Prove that if X is a Banach space and Qx and Qx' are the natural
maps from X and X into X" and X" respectively, then Qx' (X')
and (Qx(X).L are complementary subspaces of X.
(b) Find a closed subspace M of a Banach space such that M.L is com-
plemented but M is not.
3.24 No result in this section says anything about complemented subspaces of
TVSs that are not Hausdorff. There is a good reason for this. Find it.
3.3.1 Definition. Suppose that X is a set, that + and e are binary op-
erations from X x X into X, and that . is a binary operation from F x X
into X such that
(1) (X, +, . ) is a vector space;
and for all x, y, z in X and every scalar a,
(2) x e (y e z) = (x e y) e z;
(3) x e (y + z) = (x e y) + (x e z) and (x + y). z = (x. z) + (y e z); and
(4) a(xey)=(ax)ey=xe(ay).
Then (X, +, e,. ) is an algebra. This algebra is an algebra with identity if
X '" {O} and there is a member e of X such that
(5) eex=xee=x
whenever x EX, in which case e is the (multiplicative) identity of X. If
1111 is a norm on the vector space (X, +, . ) such that
With all notation as in the preceding definition, the products ax and xey
are usually abbreviated to ax and xy respectively, and it is usually said
that X is an algebra (or normed algebra or Banach algebra if there is a
norm involved) rather than using the more formal notations (X, +, e, . )
and (X, +, e,, 11-11).
306 3. Linear Operators
The reason for requiring an algebra with identity to contain more than 0
is to make it impossible for 0 to be a multiplicative identity for the algebra.
See the proof of Proposition 3.3.10 (b).
3.3.2 Example. The set IF with its usual vector space operations, its
usual multiplication, and the norm given by the absolute value function is
a Banach algebra with identity 1.
3.3.5 Example. As a special case of the preceding example, the space Roc> is
a Banach algebra with identity (1,1,1, ... ). Notice that the multiplication
of elements of Roc; is done termwise.
The following example is the most important one for the purposes of this
book.
The multiplication operations defined in the above examples are the stan-
dard ones for their spaces and are the ones that will be assumed when these
spaces are treated as algebras in this book.
3.3 Banach Algebras and Spectra 307
(j) Ilell 2: l.
(k) If x is an invertible element of X, then Ilx-III ~ IIxll- i .
308 3. Linear Operators
Many of the familiar laws of exponents from basic arithmetic also hold
for algebras, and are easy consequences of Definition 3.3.9 and parts (g)
and (h) of Proposition 3.3.10 (but beware of assuming that (xY)" = xnyn
unless it is known that xy = yx).
It often happens that an identity for a normed algebra has norm 1, and
the examples given in this section might lead one to suspect that this must
always be the C3.'le. However, part (j) of Proposition 3.3.10 cannot in general
be sharpened, even for Banach algebras. See Exercises 3.27 and 3.28.
Addition of vectors and multiplication of vectors by scalars are contin-
uous operations for a nOfmed algebra because of the continuity of these
operations for normed spaces. It turns out that multiplication of vectors
by vectors is also continuous. The proof is essentially thc same as the fa-
miliar proof of the continuity of multiplication in the scalar field.
t
n=O
1I00-(n+l) xn ll = 10:1- 1 (Ilell + t
n=l
II(o:-lx)nll)
All the ingredients are now at hand for the examination of the continuity
of inversion promised earlier in this section.
II(x+Z)-I_X-111 = II(e+x-Iz)-lx-1_x-111
:-:; Ilx-lllll(e + X-IZ)-I - ell
= Ilx-IIIII~(_x-Iz)n - ell
= IIx-IIIII~(_x-Iz)nll
n=1
= (,
as required.
With all notation as in the preceding proposition, suppose that x EX.
Then p(x) = { Q : Q E IF, Qe - x E G(X) }. The next result is therefore an
immediate consequence of the continuity of the map Q f--> Qe - x from IF
into X.
It follows from the preceding proposition and Corollary 3.3.14 that the
spectrum of an element of a Banach algebra with identity is closed and
bounded. This can be summarized as follows.
oc
n=O
Since (3e - x has both a right inverse and a left inverse, it is invertible.
In the following result, the existence of the limit in the formula is part
of the conclusion. The existence of the limit does not actually require the
complex normed algebra in question to be complete or to have an identity;
see Exercise 3.31.
whenever x E X.
PROOF. Fix an x in X. Suppose first that (3 E a(x). For each positive
integer n, the preceding lemma assures that (3n E a(x n ) and therefore that
l(3nl:s Ilxnll and 1(31 :S Ilxnll l / n . It follows that 1(31 :S liminfnllxnlil/n and
thus that rCT(x) :S liminfnllxnlil/n.
Now suppose that 'I' is a scalar such that I'l'l > rCT(x). Let x* be a
member of X*. Then x* Rx is analytic on {a : a E te, lal > rCT(x)} by
Lemma 3.3.22. Applying x* to the series expansion for Rx from Theo-
rem 3.3.13 shows that x* Rx(a) = E~=o x*(xn)a~(n+l) when lal > IIxll.
This series expansion must actually hold for x* Rx(a) when lal > ra(x)
by the uniqueness of Laurent series expansions on annuli; see, for example,
[42, p. 103]. In particular, it holds for x* Rxb), so E~=o x*(xnh~(n+1)
converges, which in turn implies that limnx'b~(n+1)xn) = O. Since x is
an arbitrary member of X', it follows that 'Y~(n+l)xn ~ 0 and therefore
314 3. Linear Operators
PROOF. It may be assumed that n > 1. Suppose that j E {2, ... ,n} and
that Xl, ... ,Xj~l are linearly independent. Suppose moreover that there
are scalars f31, ... , f3j~l such that Xj = f31Xl + ... + f3j~lXj~l. Then
Since exj-exk f 0 when k = 1, ... ,j-1, it follows that f31 = ... = f3j~l = o.
This implies that Xj = 0, a contradiction. It must be that Xl, ... , Xj are
linearly independent, so it follows by induction that Xl, ... ,X n are linearly
independent.
(S~))
r)
dim(ker(S)) = dim
= dim ((S~)
= dim(S(X).L)
= dim(ker(S*)).
Exercises
3.25 The purpose of this exercise is to show that every normed algebra can
be completed to a Banach algebra. To this end, suppose that X is a
normed algebra. Show that there is a Banach algebra Y and a map T
from X onto a dense subspace of Y such that T is a normed algebra
isometric isomorphism, that is, a normed space isometric isomorphism
for which T(XIX2) = T(xI)T(X2) whenever Xl, X2 EX. Show that if X
ha.s identity e, then Y ha.s identity T(e).
3.26 Suppose that X is an algebra. Let X[e] be the vector space sum X x IF; see
the introduction to Section 1.8 for the definition of a vector space sum.
Define a multiplication of elements of X[e] by the formula (x, a)(y, (3) =
(xy + o.y + (3x, 0.(3).
(a) Let e be the element (0,1) of X[e]. Prove that with its multiplication
and its vector space operations a.s a vector space sum, the space X[e]
is an algebra with identity e.
(b) Suppose that 11llx is a norm for X. Define a norm for X[e] by the
formula lI(x, a) IIX[e] = Ilxll x + 10.1. Prove that this really does define
a norm for X[e]. Prove that if X with 11llx is a normed or Banach
algebra, then X[e] with 11IIX[e] is, respectively, a normed or Banach
algebra with identity.
(c) With all notation a.s in (b), suppose that X with the norm 11llx is
a normed algebra and therefore that X[e] with the norm 11IIX[e] is
a normed algebra with identity. Prove that the map T: X -+ X[e]
given by the formula T(x) = (x,O) is a normed algebra isometric
isomorphism from X into X[e]; see Exercise 3.25 for the definition.
Thus, every Banach algebra without an identity can be embedded into a
Banach algebra with identity as a subalgebra having codimension 1.
3.21 Suppose that t 2: 1. Give an example of a Banach algebra with identity e
such that lIell = t.
3.28 Suppose that X is a normed algebra with identity e. For each x in X, let
Tx(Y) = xy whenever y E X. Show that the map x f-+ Tx is a normed
algebra isomorphism (that is, a normed space isomorphism such that
TXl TX2 = T X1X2 whenever Xl, X2 E X) from X onto a sub algebra of B(X).
Conclude that X is isomorphic as a normed algebra to a normed algebra
with identity e' such that Ile'll = 1, whether or not lIell = 1.
3.29 Suppose that (o.n) is a member of the Banach algebra 00. Identify the
spectrum of (an).
3.30 Let X be the real Banach algebra with identity formed from the com-
plex Banach algebra 00 by restricting multiplication of vectors by scalars
to lR x 00' and let x be the member (i, 0, 0, ... ) of X.
(a) Show that O"(x) = 0.
(b) Find limnllxnli n . Comment on what this says about extending the
l /
(c) Let Sl be the left-shift operator from Y into Y given by the formula
for Tl from (b). Show that CT(Sz) = CTp(Sz) = II)).
3.4.1 Definition. (D. Hilbert, 1906 [103]; F. Riesz, 1918 [195]). Suppose
that X and Yare Banach spaces. A linear operator T from X into Y
is compact if T(B) is a relatively compact subset of Y whenever B is a
bounded subset of X. The collection of all compact linear operators from X
into Y is denoted by K(X, Y), or by just K(X) if X = Y.
As will be seen in Example 3.4.5, not every compact operator has finite
rank.
Recall that a subset S of a metric space X is totally bounded if, for every
positive , there is a finite subset FE of S (or, equivalently, of X) such that
every point of S is within distance of a member of FE; that is, such that
S is covered by the finite collection of open balls in X having radius and
centered at the points of FE' It is a basic fact from the theory of metric
spaces that a subset of a complete metric space is relatively compact if and
only if it is totally bounded; see, for example, [67, p. 22] or [129, p. 101J.
This, together with other obvious arguments, proves the following collection
of characterizations of compact operators.
3.4.5 Example. Define the linear operator T from 2 into itself by the
formula T(a n ) = (n- 1a n ). Let ((,8j,n));':l be a bounded sequence ofmem-
bers of 2' It will be shown that (T(,8j,n));':l has a convergent subse-
quence. Since 2 is reflexive, each of its bounded sequences has a weakly
convergent subsequence, so it may be assumed that there is a member (,8n)
of 2 such that w-limj(,8j,n) = (,8n). It follows that limj ,8j,n = ,8n for
each n. Also, for each positive there is a positive integer n, such that
L~=n,ln-1(,8j,n - ,8n)12 < for every j. These two facts together imply
that limj T(,8j,n) = T(,8n). It follows from the equivalence of (a) and (d)
in Proposition 3.4.4 that T is compact. Notice that T(X) contains the
sequence (en) of standard unit vectors of 2, so T is an example of a
compact operator that does not have finite rank. Notice also that the
range of T is not closed since it contains, for each positive integer n,
the element (l-1,2- 1, ... ,n- 1,O,O, ... ) of 2, but does not contain the
limit (1- 1 ,2- 1 , ... ) obtained by letting n tend to 00.
3.4 Compact Operators 321
It also turns out that the product of two compact operators must be
compact. In fact, the compactness of the product follows even if only one
of the operators is compact and the other is bounded.
When A is a subset of an algebra X and x EX, the sets {xy : yEA} and
{yx : yEA} are denoted, as one might expect, by xA and Ax respectively.
It follows that Ilf - fj 1100 < E. The set S is therefore totally bounded and
so is relatively compact.
It it; not true that every compact operator has a weak*-to-norm contin-
uous adjoint. See Exercise 3.44.
The next portion of this section is devoted to obtaining Theorem 3.4.23,
Frederic Riesz's analysis of the spectrum of a compact operator from an
3.4 Compact Operators 325
infinite-dimensional Banach space into itself. The path that will be taken
to this theorem is related to Riesz's original method. See [67, p. 609] for the
history of Riesz's result and its relationship to earlier work of Fredholm,
and for references to various methods of obtaining it.
The first lemma en route to Theorem 3.4.23 is so useful throughout
functional analysis that it has become a named result.
3.4.18 Riesz's Lemma. (F. Riesz, 1918 [195]). Suppose that M is a proper
closed subspace of a normed space X and that 0 < () < 1. Then there is
an x in Sx such that d(x, M) ?: ().
PROOF. By Corollary 1.9.7, there is an x* in Sx' such that M ~ ker(x*).
Let x be an element of Sx such that Ix*xl ?: (). If y E M, then Ilx - yll ?:
Ix*x - x*YI = Ix*xl ?: (), as required.
whenever n E N.
Suppose that (I -T)(X) = X but f -T is not one-to-one. If n is a positive
integer, then (f -T)n(x) = X and therefore (f _T)n+l maps some member
of X to 0 that (I - T)n does not, so ker((I - T)n) ~ ker((f - T)n+l). By
Riesz's lemma, there is a sequence (x n ) in Sx such that if n E Nand
y E ker((I - T)n), then Xn E ker((I - T)n+I) and Ilx n - yll ?: 1/2. If
n,m EN and n > m, then (I - T)(x n ) + TX m E ker((f - T)n), so
1
IITxn - TXml1 = Ilxn - ((I - T)(x n ) + Tx m) II?: 2
It follows that (Txn) has no convergent subsequence, which contradicts the
compactness of T.
Under the hypotheses of the preceding lemma, it actually turns out that
(af - T)(X) = X if and only if af - T is one-to-one. This is an immediate
consequence of the next lemma.
326 3. Linear Operators
(I - T)(z) = lim
n
(I - T)(Yn) = 0,
The compactness ofT* also implies that (1* -'I'*)(X*) is closed and thus, by
Theorem 3.1.21, weakly* closed. It follows from Lemma 3.1.16 and Propo-
sition 2.6.6 (c) that
Next, suppose that dim(ker(I - T)) > codim((I - T)(X)). It follows from
Theorem 3.2.18 (a) that there is a closed subspace N of X such that X
is the internal direct sum of (I - T)(X) and N. Applying the first iso-
morphism theorem for Banach spaces to the projection with range Nand
kernel (I - T)(X) shows that codim(I - T)(X)) = dim(N), so there is
a bounded linear operator S that is not one-to-one that maps the finite-
dimensional Banach space ker(I - T) onto the smaller-dimensional Banach
space N. Consider S to be a member of B(ker(I - T), X), and notice that
S is compact. As in the first part of this proof, let M be a closed subspace
of X such that X is the internal direct sum of ker(I - T) and M, and
let P be the projection in X with range ker(I - T) and kernel M. Let
R = T + SP. Then R is compact since T and S are so, and
PROOF. It may be assumed that each O:n is nonzero. For each positive
integer n, let Xn be an eigenvector associated with O:n and let Mn =
({ Xl, ... , Xn}), a closed subspace of X that has dimension n by Theo-
rem 3.3.31; then Mn ~ Mn+1 and (O:n+l1 - T)(Mn +1) ~ Mn, so by
Riesz's lemma there is a Yn+l in SMn + 1 such that d(Yn+1, Mn) 2': 1/2 and
(O:n+l1 - T)(Yn+r) E Mw It follows that if j, kEN and j > k> 1, then
(a) The spectrum a(T) of T is a countable compact set whose only pos-
sible limit point is O.
(b) a(T) = {O} U ap(T).
(c) If 0: is a nonzero eigenvalue of T, then the eigenspaces of T and T*
associated with a have the same finite dimension.
The preceding theorem does not say that the spectrum of a compact
operator from an infinite-dimensional Banach space into itself is the dis-
joint union of {O} and ap(T). For example, the compact operator (an) f-->
(a1,0,0, ... ) from 2 into itself has 0 as an eigenvalue (and the associ-
ated eigenspace is infinite-dimensional). On the other hand, notice that
3.4 Compact Operators 329
= lim{sup{
n
I/(T - Tn)(x)112 : x E Bx}}
= lim
n
liT - Tn 1/.
This proves that every compact linear operator from a Banach space into 2
is the limit of a sequence of bounded finite-rank linear operators from that
Banach space into 2' Since bounded finite-rank linear operators from a
Banach space into a Banach space are compact, another way of expressing
this is to say that 2 has the following property.
The argument that proves the following result is essentially the same as
that of Example 3.4.25.
3.4.27 Theorem. The spaces Co and p such that 1 ::; p < 00 have the
approximation property.
B = {Ltnunxn: tn 2':
n
0and Un E H for each n, Ltn::;
n
I}.
Notice that there is no problem with the definition of B, since X is a
Banach space and each of the formal series that is supposed to belong to B
is absolutely convergent. It will now be shown that B is closed. Suppose
that (Ln tj,nUj,nXn)'j=1 is a sequence in B convergent to some member
of X. By a straightforward diagonalization argument, it may be assumed
that, for each n, the sequence (tj,n)~1 converges to some nonnegative tn
and (Uj,n)~1 converges to some Un in H. It follows that Ln tn ::; 1 and,
by an argument involving the fact that Xn ---+ 0, that limj Ln tj,nUj,nXn =
Ln tnunxn Therefore B is closed. It is not difficult to show that B is
convex; when H is a closed ball in IF centered at 0, this uses the fact that
if Ln tnunxn, Ln t~u~xn E Band 0 < t < 1, then
Then A2 <;;; B(0,2- 2). Since A2 is nonempty and relatively compact, there
arc members Xn2 +1,' .. , xn3 of 2A2 such that 2A2 <;;; U;!n2+1 B(Xj, 2- 3).
Let
n3
L
m
Xo - Tnxjn E 2-m Am <;;; B(O, 4- m )
n=l
PROOF. The forward implication in (a) follows immediately from the pre-
ceding lemma, while the converse comes from Lemma 3.4.29 and the fact
that every subset of a compact subset of X is relatively compact.
For (b), suppose that A is a relatively compact subset of X and that (xn)
is a sequence in X as in (a). It may be assumed that each Xn is nonzero.
Let 11llx be the norm of X. For each n, let Yn = Ilxnll~1/2xn if IIxnllx < 1
and let Yn = Xn otherwise. Let S = co( {Yn : n EN}), a compact subset
of X since I Yn I x -+ O. Then co( {xn : n EN}) <;;; S. Let Y and I . II y be as
in Lemma 3.4.28 for this S. If Ilxnllx < 1, then Ilxnll~1/2xn E B(Y,II'b-),
so Ilxnlly ::; Ilxnll3/2. It follows that Ilxnlly -+ 0, which proves (b).
334 3. Linear Operators
The proof of the preceding theorem is essentially the one found in [156]
with minor modifications to allow for the possibility of complex scalars.
For a long time one of the major open problems in Banach space the-
ory, the approximation problem, was whether every Banach space has the
approximation property. This was finally settled in the negative in a 1973
paper by Per EnBo [76], who found a separable reflexive Banach space,
necessarily infinite-dimensional, that lacks the approximation property. 1
3.4.33 Definition. (D. Hilbert, 1906 [103]). Suppose that X and Yare
Banach spaces. A linear operator T from X into Y is completely continuous
or a Dunford-Pettis operator if T(K) is a compact subset of Y whenever
K is a weakly compact subset of X.
in the famous "Scottish book" of open problems kept at the Scottish Coffee House in
Lwow, Poland, by Banach, Mazur, Stanislaw Ulam, and other mathematicians in their
circle; see [160, problem #153]. Mazur offered a live goose as the prize for a solution.
About a year after solving the problem, Enflo traveled to Warsaw to give a lecture on
his solution, after which he was awarded the goose.
3.4 Compact Operators 337
Since sequential continuity does not in general imply continuity, the pre-
ceding proposition stops a bit short of claiming that every completely con-
tinuous operator is weak-to-norm continuous, and in fact not all completely
continuous operators are; see Exercise 3.51. It does of course follow from
the preceding proposition, and in fact directly from the definition of com-
plete continuity, that every weak-to-norm continuous linear operator from
a Banach space into a Banach space is completely continuous.
The reason that Hilbert is given joint credit with Riesz for founding the
field of compact operator theory is that Hilbert was interested in linear op-
erators with domain 2 when he gave the definition of complete continuity,
and for such operators compactness and complete continuity are equivalent.
Exercises
3.39 Characterize the compact projections in a Banach space among all pro-
jections in that Banach space.
3.40 (a) Suppose that X is a reflexive Banach space. Prove that every mem-
ber of B(X, e1 ) is compact.
(b) Suppose that Y is a reflexive Banach space. Prove that every mem-
ber of B(co, Y) is compact.
338 3. Linear Operators
3.50 Suppose that T is a linear operator from a normed space X into a normed
space Y. Prove that the following are equivalent.
(a) The operator T is continuous.
(b) The set T(K) is a compact subset of Y whenever K is a compact
subset of X.
(c) The set T(K) is a weakly compact subset of Y whenever K is a
weakly compact subset of X.
Notice the analogy between this result and the definition of complete
continuity.
3.51 (a) Suppose that Y is a Banach space. Prove that every member of
B(ll, Y) is completely continuous. (Compare Exercise 3.40 (a).)
(b) Give an example of a linear operator from a nonreflexive Banach
space into a reflexive Banach space that is completely continuous
but not compact.
(c) Give an example of a linear operator from a Banach space into a
Banach space that is bounded but not completely continuous.
(d) Give an example of a linear operator from a Banach space into a
Banach space that is completely continuous but not weak-to-norm
continuous.
3.5.1 Definition. (S. Kakutani, 1938 [125]; K. Yosida, 1938 [245]). Sup-
pose that X and Yare Banach spaces. A linear operator T from X into Y
is weakly compact if T(B) is a relatively weakly compact subset of Y when-
ever B is a bounded subset of X. The collection of all weakly compact
linear operators from X into Y is denoted by KW(X, Y), or by just KW(X)
if X = Y.
The fact that the range of a compact operator is closed if and only
if the operator has finite rank does have an analog for weakly compact
3.5 Weakly Compact Operators 341
operators, and again that analog is suggested by the similarity between the
role of compact sets in finite-dimensional Banach spaces and that of weakly
compact sets in reflexive Banach spaces.
The collections KW(X, Y) and KW(X) of Definition 3.5.1 have the same
algebraic and topological closure properties that were shown to hold for
K(X, Y) and K(X) in Section 3.4. Before showing this, it will be useful
to have the following result relating the weak compactness of an operator
to the location of the range of its second adjoint. This theorem is due to
Gantmacher when the spaces are separable, while the general case is due
to Nakamura.
As was done for compact operators in the preceding section, the weakly
compact linear operators from a Banach space X into a Banach space Y will
now be characterized among all members of B(X, Y) in two different ways
based on the behavior of their adjoints. The first characterization is just
Schauder's theorem with compactness replaced by weak compactness, while
the second, an analog of Theorem 3.4.16, involves a continuity property for
the adjoint.
y***)
T***
(xCI!**, = (T** Xc>**, y***) -+ ** y ***)
(T** x, = (x,
** T*** y ***) ,
as required.
Suppose conversely that T* is weakly compact. Then T** is weakly com-
pact by what has already been proved. Since T = Ql}T**Qx, it follows
that T is weakly compact.
3.5.16 Example. It can be shown that for every Banach space Y, each
member of B(l, Y) is completely continuous; see Exercise 3.51. Therefore
1 trivially has the Dunford-Pettis property. Another (perhaps not entirely
independent) proof that 1 has this property can be found in Exercise 3.59.
The definition of the Dunford-Pettis property given above has the draw-
back that it involves every Banach space Y rather than just the space X
in question. The following theorem gives characterizations of the Dunford-
Pettis property that lack this defect.
* Vn - V V
vn vn - V *)( Vn - V )
= (* + Vn* V + V * Vn - 2V * V
--+ 0+ vv + vv - 2v*v = 0,
which shows that (d) => (c). It is obvious that (a) => (b), so the theorem
is proved.
The spaces Ll (n, E, p,) such that (n, E, p,) is a finite measure space have
the Dunford-Pettis property, a fact that is essentially due to Dunford and
Pettis [66), and the spaces C(K) such that K is a compact Hausdorff space
346 3. Linear Operators
3.5.19 Example. Let X be complex 2 and let []) be the closed unit disc
in C. Let 11 and Tr be, respectively, the left-shift and right-shift oper-
ators on X, that is, the bounded linear operators from X into X de-
fined by the formulas Tl (al,a2, ... ) = (a2,a3, ... ) and Tr (al,a2, ... ) =
(0, aI, a2' ... ). It can be shown that 0"(11) = []) and O"p(Tl) = [])O, and that
O"(Tr) = []) and O"p(Tr) = 0; see Exercise 3.38. Also, the identity operator I
on X has 1 as an eigenvalue, and the associated eigenspace, which is all
of X, is certainly not finite-dimensional. By Proposition 3.5.4, the opera-
tors T l , Tn and I are all weakly compact. Together, they show that none
of the conclusions of Theorem 3.4.23, or even reasonable generalizations of
those conclusions such as the possibility that either the point spectrum or
its complement in the spectrum must be countable, need hold for weakly
compact operators from an infinite-dimensional Banach space into itself
(except for the statement about the compactness of the spectrum, and
that is true for the spectrum of every element of every Banach algebra
with identity).
Exercises
3.52 Give an example of a bounded linear operator T from some reflexive Ba-
nach space X onto a dense subspace of some non reflexive Banach space Y.
Conclude that T is weakly compact but T(X) is not reflexive.
3.53 Suppose that X and Yare Banach spaces, that T E L(X, Y), that M is
a closed subspace of X such that M <;;; ker(T), that 1f is the quotient map
from X onto X/M, and that S is the unique map, automatically linear,
from X/M into Y such that T = So 1fj see the commutative diagram in
Theorem 1.7.13.
3.5 Weakly Compact Operators 347
Much of the theory of finite-dimensional normed spaces that has been pre-
sented in this book is ultimately based on Theorem 1.4.12, which says
that every linear operator from a finite-dimensional normed space X into
any normed space Y is bounded. A careful examination of the proof of
that theorem shows that it essentially amounts to demonstrating that if
Xl, ... , Xn is a vector space basis for X, then each of the linear "coordi-
nate functionals" O'IXI + ... + O'nXn 1-4 O'm, m = 1, ... , n, is bounded; this
can be seen from the nature of the norm /. / used in the proof and the
argument near the end of the proof that there is no sequence (Zj) in Bx
such that /1 Zj / 2: j for each j. It should not be too surprising that many
topological results about finite-dimensional normed spaces are ultimately
based on the continuity of the members of the family ~# of coordinate
functionals for some basis ~ for the space, since it is an easy consequence
of Proposition 2.4.8, Theorem 2.4.11, and the uniqueness of Hausdorff vec-
tor topologies for finite-dimensional vector spaces that the norm topology
of the space is the ~# topology of the space.
Though it is possible for an incomplete infinite-dimensional normed space
to have a vector space basis ~ such that all coordinate functionals for ~ are
bounded, this can never happen for an infinite-dimensional Banach space;
see Exercises 4.1 and 4.2. Thus, the topology of an infinite-dimensional
Banach space is not nearly so closely tied to coordinate functionals with
respect to a vector space basis as is the topology of a finite-dimensional
Banach space. It is primarily for this reason that vector space bases have
not been used very much in this book's exploration of the general theory
of Banach spaces.
350 4. Schauder Bases
Schauder included in his definition the requirement that each of the co-
ordinate functionals 2::n anIn I-> am such that mEN be continuous. As
will be shown in Corollary 4.1.16, that actually follows from the rest of the
definition.
The following generalization of the notion of a Schauder basis also has
its uses.
vector space bases are often called Hamel bases after the German analyst
and applied mathematician Georg Karl Wilhelm Hamel (1877~1954).
The basis of the preceding example has the special property that each
of its terms has norm 1. As will now be shown, every Banach space having
any basis whatever has a basis with this property that can be obtained
from the original basis in the obvious way.
The use of the terms "bounded basic sequence" and "bounded basis"
in the sense of the preceding definition is common but unfortunate, since
the usual meaning of the word "bounded" when used to describe a se-
quence (xn) in a Banach space does not imply that infnllxnll > O. To avoid
confusion, the word will not be applied to basic sequences in this book.
following proposition shows how this is done, and is obvious enough that
it is really just an observation.
4.1.11 Example: The classical Schauder basis for e[O, 1]. This example
is from Schauder's 1927 paper [206] that introduced the notion of Schauder
4.1 First Properties of Schauder Bases 353
'G 1
'~ 1
FIGURE 4.1. The first few terms of the classical Schauder basis for e[a, 1].
if 2n-2
2
Tn
- 1 < t < 2n-l
- 2m
- 1,
otherwise.
See Figure 4.1 for the graphs of So through S8, which should demystify
these formulas a bit.
Suppose that 1 E e[o, 1]. Define a sequence (Pn):;'=o in e[o, 1] in the
following way. Let
Po = 1(0)so,
Pl = Po + (J(1) - Po(l))sl'
P2 = + (J(1/2) -
Pl Pl(1/2))s2,
P3 = P2 + (J(1/4) - P2(1/4))s3,
P4 = P3 + (J(3/4) - P3(3/4))S4'
P5 = P4 + (J(1/8) - P4(1/8))s5,
P6 = P5 + (J(3/8) - P5(3/8))S6,
P7 = P6 + (J(5/8) - P6(5/8))S7'
P8 = P7 + (J(7/8) - P7(7/8))S8,
and so forth. Then Po is the constant function that agrees with 1 at 0, while
PI agrees with 1 at 0 and 1 and interpolates linearly in between, and P2
agrees with 1 at 0, 1, and 1/2 and interpolates linearly in between, and so
forth. For each nonnegative integer n, let an be the coefficient of Sn in the
formula for Pn. Then Pm = 2::=0 ans n for each m. It follows easily from
354 4. Schauder Bases
4.1.12 Definition. Suppose that a Banach space X has a basis (xn). For
each positive integer m, the mth coordinate functional x;" for (xn) is the
map Ln anXn f---+ am from X into F, and the mth natural projection Pm
for (xn) is the map Ln anXn f---+ L~=l anXn from X into X.
\Vith all notation as in the preceding definition, it is clear that each x;"
is a linear functional on X and that each Pm is a projection from X onto
({ Xl, ... ,xm})' To show that these maps are bounded, it is convenient
to work not with the original norm of the underlying Banach space, but
instead with the following one.
It is important to notice that the norm 1111 (Xn) in the preceding definition
depends not only on (x n ), but also on the space's original norm.
~ 2112)on,J, - on,h)xnll
n (xn)
for each m. Letting j = j, and letting j' tend to infinity shows that
f
Il n=l nXn - f
n=l
on,j,xnll ~~ (4.1)
Now let mE be such that if m2 2: m1 > mEl then IIL:~m, on,],xnll < E/3.
It follows that
II ~
n-ml
Onxnll ~ II ~
n-ml
OnXn - ~
n-ml
on,j,xnll + II ~
n-ml
On,j,Xnll < E
The term monotone for a basis with basis constant 1 was introduced by
M. M. Day in his 1958 book [53], while the older term orthogonal for such
bases dates back to independent studies of them by V. Ya. Kozlov [142]
and R. C. James [1l0] published in 1950 and 1951 respectively.
Since each natural projection for a basis has norm at least 1, basis con-
stants are always greater than or equal to 1, and a basis is monotone if and
only if each of the natural projections for the basis has norm 1.
Of course, a basic sequence is a basis for the closed linear hull of the set
consisting of the terms of the sequence and therefore has a basis constant
associated with it. As one would expect, a monotone basic sequence is one
whose basis constant is 1. In general, terminology used for bases automat-
ically extends to basic sequences in this fashion as long as the extension
makes sense.
It must be emphasized that basis constants depend on the norm of the
space. In particular, it will follow from Exercise 4.11 and Corollary 4.1.22
that every Banach space with a basis has a nonmonotone basis (xn) that
becomes monotone when the space is renormed with 11II(xn ) .
4.1.19 Example. If X is Co or p such that 1 :::; p < 00, then the standard
unit vector basis for X is monotone.
K
= sup
{ IIL~:"l anxnll '"
I Ln anxn II : Dn anxn E X
\ { } }
0, mEN ,
The equivalence of (a) and (b) in the next result follows immediately
from the preceding proposition. It is clear that (b) and (c) are equivalent,
while the equivalence of (c) and (d) is an easy consequence of the definition
of the (xn) norm.
For example, the standard unit vector basis for each space p such that
1 ::; p < 00 is strictly monotone, while the corresponding basis for Co is
monotone but not strictly monotone.
So far, the results of this section have been about conclusions that can
be drawn when it is known that certain sequences are basic rather than
properties sequences can have that assure that they are basic. The next
theorem is probably the most important result of the latter kind.
whenever ml,m2 EN, ml::; m2, and Q:l, ... ,Q: m2 E IF; and
(3) [{xn:nEN}]=X.
4.1 First Properties of Schauder Bases 359
PROOF. By what has already been proved in this section, the sequence (xn)
has properties (1), (2), and (3) if it is a basis for X.
For the proof of the converse, assume that (xn) satisfies (1), (2), and (3).
Suppose first that (J3n) and (,n) are sequences of scalars such that Ln J3nxn
and Ln 'YnXn both converge and are equal. By (2),
4.1.27 Example: The Haar basis for Lp[O, 1], 1 ::::: p < =. Suppose that
1 ::::: p <=.
Define a sequence (h n ) in Lp[O,lJ as follows. Let hI be 1
360 4. Schauder Bases
I I
I t I t I t I I
I I II I I I I II
I I I I I I
-I -I ~~ -1 ~ -I
'~
-I
I
I I
I I
~
I
t
-1
I I
II
I I
I I
-I
I I
II
I I
I I
1
-I
I I
I I
111
I I
FIGURE 4.2. The first few terms of the Haar basis for Lp[D, 1], 1 ::; p < 00.
otherwise.
See Figure 4.2 for the graphs of the first few of these functions, which will
make it clear how they are being constructed. The vertical dashed lines
in the graphs are included for clarity. Notice that each h n is a positive
multiple of the derivative of the corresponding member Sn of the classical
Schauder basis for e[O, 1].
Suppose that (an) is a sequence of scalars, that no and mo are positive
integers such that no 2:: 2 and 2 mo - 1 < no :s: 2ffi o, and that hand /2 are the
respective intervals [2n2 rno-2
o
2no-1 - 1) and [2n2 o-1
- 1 '2"T1l.0 - 1 ~ - 1) . Let a
0 ' 2171.0 171
2:: O.
Therefore 11L:~,,=~1 anhnll p ::; 11L:~"=1 anhnll p , and so 11L::~1 anhnll p :s:
11L::~1 anhnll p whenever ml,m2 EN and ml :s: m2
It is easy to check that ({ h n : n EN}) contains the indicator function of
every dyadic interval [~-;;,l, 2';" ) such that m is a nonnegative integer and n is
4.1 First Properties of Schauder Bases 361
1 "Remarquons toutefois que tout espace du type (8) a une infinite de dimensions
renferme un ensemble lineaire ferme a une infinite de dimensions qui admet une base."
362 4. Schauder Bases
whenever ml, m2 E N, ml < m2, and aI, ... ,am2 E IF. The sequence (xn) is
therefore basic by Corollary 4.1.25 and has basis constant no more than M
by Proposition 4.1.20.
see the comments preceding Lemma 4.1.28. In his paper, Pelczynski points
out that the result was first proved by Czeslaw Bessaga in his thesis, also
using Mazur's method.
Notice the close similarity between the proof of this theorem and that of
Theorem 4.1.30.
whenever ml, m2 E N, ml < m2, and aI, ... ,am2 E IF. The sequence (XnJ
is therefore basic by Corollary 4.1.25.
The final topic for this section is the brief exploration of the connection
between bases and the approximation property promised earlier. The main
result in this direction is the following one.
4.1.33 Theorem. Every Banach space with a basis has the approximation
property.
PROOF. Suppose that X is a Banach space with a basis (x n ), that (Pn )
is the sequence of natural projections for (x n ), and that M is the basis
constant for (xn). Let K be a nonempty compact subset of X and let I'
be a positive number. By Theorem 3.4.32, it is enough to find a positive
integer no such that IIPnox - xii < c whenever x E X.
Let Yl, ... ,Ym be members of K such that every member of K is within
distance c/(M + 2) of some Yj, and let no be a positive integer such that
IlPnoYj-Yjll < c/(M+2) whenj = 1, ... ,m. Suppose that x is any member
of K. Let jo be such that Ilx - Yjo I < c/(M + 2). Then
Exercises
bounded, and that, in fact, only finitely many can be so. Let '13t be the
collection of members of '13# that are bounded.
(a) Show by example that '13t might be nonempty.
(b) Show that sup{ 11/11 : I E '13t }-j. +00.
(c) Show that '13t is finite.
4.2 Give an example of an infinite-dimensional normed space with a vector
space basis '13 such that every coordinate functional for '13 is bounded.
(Compare Exercise 4.1.)
4.3 Prove the claims made in Example 4.1.3.
4.4 Prove that the sequence (1,1,1, ... ), el, e2, e3, ... is a basis for the Banach
space c of Exercise 1.25.
4.5 Prove that if Banach spaces X and Y have bases, then so does X EEl Y.
4.6 Suppose that (Xn) is a basis for a Banach space X and that (nj) is a
sequence of positive integers such that N \ {nj : j EN} is infinite. Let
M = [{ Xnj : j EN}]. Prove that both M and X / M have bases.
4.1 Suppose that X is a complex Banach space and Xr is the real Banach
space obtained from X by restricting multiplication of vectors by scalars
to R x X. Show that a sequence (Xn) in X is a basis for X if and only if
the sequence (Xl,ixl,X2,ix2, ... ) is a basis for X r .
4.8 Suppose that (xnl is a basis for a Banach space and that (An) is a sequence
of nonzero scalars.
(a) Must each coordinate functional for (AnXn) be the same as the cor-
responding coordinate functional for (Xn)?
(b) Must each natural projection for (AnXn) be the same as the corre-
sponding natural projection for (Xn)?
(c) Must the basis constant for (AnXn) be the same as the basis constant
for (Xn)?
4.9 Prove that the classical Schauder basis for e[G, 1] is monotone. Is it strictly
monotone?
4.10 Suppose that 1 S p < 00. Is the Haar basis for Lp[G, 1] strictly monotone
for any or all such p?
4.11 Prove that every Banach space with a basis has a nonmonotone basis.
4.12 (a) Suppose that (xnl is a basis for a Banach space X. What is wrong
with the following "proof" that there is only one norm for X equiv-
alent to the original norm of X with respect to which (Xnl is mono-
tone? "By the equivalence of (a) and (d) in Proposition 4.1.21, every
norm for X equivalent to the original norm of X with respect to
which (Xn) is monotone is the same tl.'l the (Xn) norm of X, so there
is only one such norm."
(b) Display a basis (Xn) for a Banach space X and two different norms
for X equivalent to the original norm of X with respect to which
(xn) is monotone.
4.1 First Properties of Schauder Bases 367
4.20 Give a proof of Theorem 4.1.33 that uses nothing deeper about the ap-
proximation property than its definition, and in particular avoids the use
of Theorem 3.4.32.
4.21 Prove that a separable Banach space X has the bounded approximation
property if and only if there is a sequence (Tn) of finite-rank members
of B(X) such that limn Tnx = X for each X in X.
4.22 Prove that every Banach space with a monotone basis has the metric
approximation property.
368 4. Schauder Bases
4.23 The major content of Theorem 4.1.24-namely, that conditions (1), (2),
and (3) in the statement of that theorem together imply that the se-
quence (Xn) is a basis-does appear in Banach's book [13] in an equiva-
lent form, but it is not all that easy to spot. Banach's result is actually
about biorthogonal sequences rather than bases, where a biorthogonal "se-
quence" is actually a pair of sequences (x n), (x~) such that (Xn) lies in
some Banach space, the sequence (x~) lies in the dual of that space, and
X;"Xn is 1 if m = nand 0 otherwise. This is Banach's result, which appears
as Theorem 4 of Chapter VII on page 108 of his book:
Suppose that X is a Banach space; that (Xn) is a sequence in X and
(x~) a sequence in X that together form a biorthogonal sequence;
that [{ Xn : n EN}] = X; and that, for each x in X, the partial
sums ofthe formal series Ln(X~X)Xn are bounded. Then Ln(x~X)Xn
converges for each x in X.
Here is another way to look at this result. Suppose that X, (x n ), and (x~)
satisfy the initial parts of the hypotheses of Banach's result up to and
including the requirement that [{ Xn : n EN}] = X. For each positive
integer m, let Pm be the member of B(X) given by the formula Plnx =
L;;'=l (x~x )Xn. Banach's result says that if sUPIn IIPlnxl1 is finite for each x
in X, then limm Pmx exists for each x in X. The purpose of this exercise
is to demonstrate that Banach's result is essentially the same as the major
content of Theorem 4.1.24 by showing that each can be readily derived
from the other.
(a) Suppose that X, (Xn), and (x~) satisfy the hypotheses of Banach's
result. Use Theorem 4.1.24 to show that the sequence (Xn) is a basis
for X having (x~) as its sequence of coordinate functionals, and that
the conclusion of Banach's result follows.
For the rest of this exercise, assume that Banach's result is known to hold
but that neither Theorem 4.1.24 nor the consequences of it demonstrated
in part (a) have been proved.
(b) Show that under the hypotheses of Banach's result and using the
notation in the explanation of Banach's result given above, it follows
that lim ln Plnx = x whenever x E X, that is, that x = Ln(X~X)Xn
for each x in X.
(c) Suppose that (Xn) is a sequence in a Banach space X satisfying
conditions (1), (2), and (3) of Theorem 4.1.24. Use Banach's result
to prove that (xn) is a basis for X. You may duplicate the relevant
portion of the proof of Theorem 4.1.24 up to and including the
construction of the maps Pin, but you should then concentrate on
obtaining the sequence (x;.) as in the hypotheses of Banach's result
with the goal of applying his result as quickly thereafter as possible.
and could have been given in Section 1.3. The first of these propositions
has a much shorter, though less elementary, proof; see Exercise 4.25.
and
and
and
Continue the construction of the sequences (Pn) and (qn) in this fashion.
Now let 1T' be the permutation of N obtained by listing N in the following
order. First list 1T(1) through 1T(Pl), then follow this by the members of
1, ... , ql not already listed. Follow this by the members of 1T(1), ... , 1T(P2)
370 4. Schauder Bases
not already listed, and in turn follow that by the members of 1, ... ,Q2
not already listed, and so forth. Since the partial sums of Ln X7l"(n) swing
back and forth between being within c/3 of Ln X7l'(n) and being within c/3
of Ln X n , the series Ln X7l"(n) does not converge.
4.2.2 Definition. In any setting in which the notion of series makes sense,
a subseries of a formal series '"'"
L-n Xn is a formal series '""'.
~J X n J obtained
from a subsequence (Xnj) of (xn).
and IIL3~Pk Xnj II ~ E for each k. It may be assumed that infinitely many
positive integers are omitted from Uk {nj : Pk S; j S; qk}. Let (Tn) be the
sequence consisting of those omitted integers in ascending order, and let 7f
be the permutation of N into the order
for each k. Let (nj) be formed by applying 7f to the terms of the se-
quence Sl, ... , t 1, S2, ... , t2, . .. to get 7f( 81), ... , 7f( h), 7f( S2), ... , 7f( t2), ...
and sorting the resulting sequence into ascending order. Then the sub-
series L j x nJ of Ln Xn does not converge.
The next lemma is proved in a bit more generality than is really needed
here, but the more general version will have a further use in Section 4.3 to
prove Proposition 4.3.9.
Considering the title of this section and the emphasis already placed
on unconditionally convergent series, the reader may have anticipated the
following definition.
4.2.11 Example. Suppose that X is Co or lp such that 1 :::; P < 00. Then
the standard unit vector basis for X is clearly unconditional.
II f
n=n,+1
,6n a n x nll2: Ilf(3n anxnll-
n=l
f: Ilanxnll
n=l
2: l.
374 4. Schauder Bases
Letting j' tend to infinity shows that if (,Bn) E Si= and m is a positive
integer, then
IILanXn
n
- Lan,jXnll
n bmu-(x n )
::; f
Xn\\
bmu-(xn) < \\l.:13n
\\l.:anX n \\
n n bmu-(xnl
whenever Ln anxn, Ln 13n x n E X and lanl ::; l13nl for each n.
PROOF. If Ln anxn , Ln 13nxn E X and lanl ::; l13nl for each n,
then there
is a member (rn) of Be oc such that an = In13n for each n, which by Propo-
sition 4.2.17 implies that
as required.
Bounded multiplier unconditional norms provide a way to turn any Ba-
nach space having an unconditional basis into a Banach algebra. Suppose
that a Banach space (X, +,,1111) has an unconditional basis. By Corol-
lary 4.2.13, the space has a normalized unconditional basis (xn). Define a
multiplication of the elements of X by the formula
It follows from the bounded multiplier test that the series on the right side
of this formula does converge, since the convergence of Ln anXn and the
fact that Ilxnll = 1 for each n together imply that (an) E co. It is easy to
check that for all x, y, z in X and every scalar a,
(1) xe(yez)=(xey)ez;
(2) x e (y + z) = (x e y) + (x e z) and (x + y) e z = (x e z) + (y e z); and
(3) a(xey)=(ax)ey=xe(a.y).
376 4. Schauder Bases
1 2::anf3nxnll
n bmu-(xn)
<:; II(an)lIooI12::f3nXnll
n bmu-(xn)
algebra. This algebra cannot have a multiplicative identity; the only candi-
date is '2::n x,,, and that series does not converge. The algebra does, how-
ever, have the following property.
whenever '2::n anxn , '2::n f3n x n EX. Then (X, +,.,', 11llbmu-(x n )) is a com-
mutative Banach algebra without identity.
4.2.22 Theorem. Suppose that (X, 1111) is a real Banach space having an
unconditional basis (x n ). Define a partial order ::S on X by declaring that
Ln anXn ::S Ln f3n x n when an ::; f3n for each n. Then (X,::s, 11llbmu-(x n)
is a Banach lattice. Furthermore, if x and y are members of X, where
x = Ln anXn and y = Ln f3n x n, then x Vy = Ln max{ an, f3n}x n , X 1\ Y =
Ln min {an ,f3n}xn, and Ixl = Lnlanlxn.
PROOF. It is easy to verify that (X, ::S) is an ordered vector space. Now sup--
pose that x, y E X, where x = Ln anXn and y = Ln f3nxn. It follows from
the bounded multiplier test that Lnlanl Xn and Lnlf3nl Xn both converge,
and therefore that Ln(lanl + lf3nl)x n converges. The bounded multiplier
test then implies that both Ln max { an, f3n} Xn and Ln min { an, f3n} Xn
converge. It is clear that Lnmax{an,f3n}xn and Lnmin{an,f3n}xn are,
respectively, the least upper bound and greatest lower bound for x and y,
so (X,::s) is a vector lattice. Notice that Ixl = Ln ma..x:{a n , -an}x n =
Lnlanl x n
Finally, suppose that Ixl ::S IYI, that is, that lanl ::; lf3nl for each n. Then
Ilxllbmu-(xn) ::; Ilyllbmu-(x n) by Proposition 4.2.18, so (X,~, 11llbmu.(x n) is
a Banach lattice.
4.2.23 Example. Suppose that X is Co or p such that 1 ::; p < 00 and that
(en) is the standard unit vector basis for X. Then the original norm of X is
the same as its bmu-(e n ) norm, so the multiplication of elements of X given
by the formula (an) (f3n) = (a n f3n) turns X into a commutative Banach
algebra. Furthermore, if IF = 1R, then X is a Banach lattice with respect to
the partial order given by declaring that (an) ::S (!3n ) when an ::; f3n for
each n.
378 4. Schauder Bases
The reader interested in Banach lattices can find out more about them
from [157], while [205] is a good source of information on lattices in more
general topological vector spaces.
When (xn) is a basis for a Banach space, the (xn) norm is a useful tool
for demonstrating the continuity of the natural projections {Pn : n EN}
for (xn); see the proof of Theorem 4.1.15. If the basis (xn) is unconditional,
then the bmu-(xn) norm can be used to prove the boundedness of a much
larger class of linear maps.
Notice that the maps Pn of Definition 4.1.12 are just special cases of
those of Definition 4.2.24.
1 T h nl(Lanxn)11
n bmu-(x n
=
1
IILinanxnl1 bmu-(xnl <- IILanxnl1 bmu-(xnl
n n
whenever 2:: n OnXn E X and A <::: N, which is in turn the smallest real
number M such that
for each pair A and B of finite subsets of N such that A <::: B and each
collection {on : nEB} of scalars.
PROOF. It is a straightforward consequence of the definition of Kub that
112: anXn - 2:
nEA nEN\A
anxnll ~ MII2: anxn l
n
for each pair A and B of disjoint finite subsets of N and each collection
n E A u B} of scalars.
{ O:n :
so IILnEA O::nxnll ::; ~(l+Ku)IILn O::nxnll It follows that Kub ::; ~(1 + Ku).
Now ~(1 + Ku) ::; Ku because Ku ~ 1, while Ku ::; 2Kub since
::; 2Kub \I Ln
O::nXn \I
The test for being a basis given in Theorem 4.1.24 has its analog for
unconditional bases.
for each pair A and B of finite subsets of N such that A c:;;: Band
each collection {O::n : nEB} of scalars; and
(3) [{xn:nEN}]=X.
PROOF. By results earlier in this chapter, the sequence (xn) has properties
(1), (2), and (3) if it is an unconditional basis for X.
Suppose conversely that (xn) satisfies (1), (2), and (3). Then (xn) is a
basis for X by Theorem 4.1.24, so the only issue is whether it is uncon-
ditional. Suppose that L n O::nXn E X and that L j O::nj x nj is a subseries
of this series. It is enough to show that L j o::njx nj converges. This follows
immediately from the convergence of L n O::nXn and the fact that
II~
J-Tnl
O::nJxn J I ::; Mil ~ o::nxnll
n-nTTll
4.2.33 Corollary. A sequence (xn) in a Banach space is an unconditional
basic sequence if and only if each Xn is nonzero and there is a real number M
382 4. Schauder Bases
such that
for each pair A and B of finite subsets of N such that A ~ B and each
collection { an : nEB} of scalars.
when n 2: 3. Since
when n 2: 3, an easy induction argument shows that 1/2 :=:: an - an -1 :=:: 3/4
when n 2: 2. It follows from this that (an) is strictly increasing and un-
bounded and that 112:,7=1 tj 1100 = an for each n.
Now let bn = (2:,7=1(-1)J+ 1tj)(vn ) for each n; that is, let (v1,b 1 ) =
(1/2,1) and, when n ;::: 2, let (vn' bn ) be the coordinates of the vertex added
4.2 Unconditional Bases 383
for each n.
If (sn) were unconditional and Ku were its unconditional constant, then
it would have to be true that
for each n, which contradicts the fact that (an) is an unbounded sequence
of positive numbers. The basis (sn) is therefore conditional.
Notice that, in the notation of the proof of the preceding theorem, the
formal series Ln tn and L n (-1)n+lt n do not converge; in fact, the se-
ries Ln ant n cannot converge for any sequence (an) of signs since the terms
of the series do not tend to 0. This illustrates the power of the methods
that have been developed in this section, since the preceding proof shows
that the classical Schauder basis for e[O, 1] is conditional without actually
producing any members of e[O, 1] whose expansions with respect to the
basis are only conditionally convergent.
It actually turns out that no basis for e[O, 1] is unconditional. Proofs of
this can be found in [156] and [216).
By induction, there is a subsequence (gn]) of (gn) such that I 2::~=1 gnj III ~
k/2 for each k.
Suppose that (h~) were an unconditional basis. Let Kub be its uncondi-
tional basis constant. Then for each positive integer k,
a contradiction.
As with the proof that the classical Schauder basis for e[O, 1] is condi-
tional, the above proof does not actually produce any elements of the space
whose basis expansions are only conditionally convergent.
It can be shown that L1 [0, 1] has no unconditional bases at all; proofs
of this can be found in [156] and [216]. It does turn out that the Haar
basis for Lp[O, 1] is unconditional if 1 < p < 00, but that is not as easy to
show as the conditional nature of the basis when p = 1; see [157] and [216]
for proofs. Exercise 4.30 gives an example of a Banach space that has two
natural bases, one of which is unconditional and the other conditional.
Much has been said in this section about unconditional bases, but little
about unconditional basic sequences. This does not mean that they are
unimportant. In fact, the settling of an old problem about unconditional
basic sequences sparked a burst of activity in Banach space theory in the
early 1990s. Bessaga and Pelczyiiski, having just provided in their 1958
paper [26] one of the first published proofs that every infinite-dimensional
Banach space has a basic sequence in it, did not wait long before asking
whether all infinite-dimensional Banach spaces have unconditional basic
sequences lurking in them; their paper [25] containing the question is the
very next paper in the same volume of the same journal. 2 The problem
2Gowers and Maurey mention in [94J that although Bessaga and Pe!czynski's paper
apparently marked the first appearance of this question in print, Mazur was aware of it
at least ten years earlier.
4.2 Unconditional Bases 385
remained open until the summer of 1991, when Timothy Gowers found
a counterexample based on earlier work of Thomas Schlumprecht [210].
Shortly thereafter, Bernard Maurey independently found essentially the
same counterexample by essentially the same argument, so the two decided
to publish jointly. Their paper [94] appeared in 1993. Upon seeing Gowers's
and Maurey's original preprints, William B. Johnson pointed out that the
proofs could be modified to show that their space was the first example of a
hereditarily indecomposable Banach space; see [94] for the definition of this
property. Acting on Johnson's observation, Gowers [92] was able to adapt
the construction to produce the first example of an infinite-dimensional Ba-
nach space X having a closed subspace Y of codimension 1 not isomorphic
to X, thus settling a question of Banach known as the hyperplane problem.
Soon afterward, Gowers [91, 93] was able to produce counterexamples that
settled several other major open problems in Banach space theory.
A good, brief, and accessible summary of the Schlumprecht-Gowers-
Maurey-Johnson accomplishments, as well as several related problems still
open at the time, can be found in Peter Casazza's 1994 book review [39].
Exer-cises
4.24 Let (Xn) be a sequence in a normed space X and let I be the collection
of all finite subsets of N directed by declaring that A :j B when A ~ B.
Define a net (SA) with index set I by letting SA = LnEA Xu; as usual,
the empty sum is defined to be the zero element of X. Prove that Ln Xn
is unconditionally convergent if and only if (SA) converges.
4.25 Prove the following generalization of Proposition 4.2.1: Suppose that X is
a topological vector space for which X is a separating family. If Ln Xn is
a series in X such that Ln x (n) converges for each permutation 7l" of 1'\1,
7f
or the formula
bases (xn) and (Yn) for Banach spaces are enough alike that the Banach
spaces have to be isomorphic.
4.3.1 Definition. Two bases (xn) and (Yn) for Banach spaces are equiva-
lent if, for every sequence (an) of scalars, the series En anXn converges if
and only if En anYn converges.
4.3.2 Proposition. Suppose that (xn) and (Yn) are bases for the respective
Banach spaces X and Y. Then (xn) and (Yn) are equivalent if and only if
there is an isomorphism T from X onto Y such that TX n = Yn for each n.
PROOF. It is clear that (xn) and (Yn) are equivalent if there is an isomor-
phism T from X onto Y such that TX n = Yn for each n. Suppose conversely
that (xn) and (Yn) are equivalent. If (En !3n,jXn)fr= 1 is a sequence in X that
converges to some member En !3nxn of X and (En !3n,jYn)fr=l converges
to some Y in Y, then it follows from the continuity of the coordinate func-
tionals for (xn) and (Yn) that Y = En !3nYn. It is an easy consequence of
the closed graph theorem and the equivalence of (xn) and (Yn) that the
map En anXn f--> En anYn is an isomorphism from X onto Y that maps
each Xn to the corresponding Yn.
4.3.5 Corollary. Suppose that a Banach space X has a basis (xn) and
that D is a dense subset of X. Then X has a basis equivalent to (xn)
whose terms all come from D.
PROOF. Let (en) be the standard unit vector basis for 1. Suppose first
that (xn) is a basic sequence equivalent to (en). Let T be an isomorphism
from [{ Xn : n EN} 1 onto 1 that maps each Xn to the corresponding en
Then sUPnllxnl1 = suPnllT-lenll :s; liT-III < +00. Furthermore, if mEN
and 0:1, ... ,O:m E IF, then
suppose that ml, m2 EN, that ml ::; m2, and that al,, a'm2 E IF. Then
L
m2
It follows that 2::n ane n converges if and only if 2::n anXn converges, so
(xn) and (en) are equivalent.
Notice that the hypotheses of the following theorem are stronger than
those of the preceding one, since the statement that (Xn) is basic is now
in the hypotheses rather than in one of the equivalent statements in the
conclusion. See Exercise 4.34 for the reason for this.
whenever mEN and al, ... , am ElF', so the claimed inequality is obtained
by letting M = IITII.
Now suppose instead that infnllxnll > 0 and there is a positive con-
stant M such that 112:::=1 anxnll ::; Mmax{ lanl : n = 1, ... ,m} whenever
390 4. Schauder Bases
mEN and 01, ... , am E F. Let (an) be a sequence of scalars. If the series
'Ln OnXn converges, then so does 'Ln Onen since (an) E Co. On the other
hand, if 'Ln Onen converges, then
whenever m1,m2 E Nand m1 :s; m2, so 'Ln 0nXn must also converge. The
basic sequences (xn) and (en) are therefore equivalent.
integer N there are positive integers PN, qN such that qN ~ PN > Nand
IILJ~PN x nj II ~ f.. After thinning (xnJ if necessary, it may be assumed
that there is an increasing sequence (mk) of positive integers such that
IIL7:;.~-1 xnjll ~ f for each k. Let Yk = L~;.~-l x nj for each k. Then
the formal series Lk Yk is weakly unconditionally Cauchy since Ln Xn is
so. Since Yk ~ 0, it follows from Theorem 4.1.32 that (Yk) can be thinned to
a basic sequence that will also be denoted by (Yk). By Theorem 4.3.10, the
basic sequence (Yk) is equivalent to the standard unit vector basis for Co .
4.3.16 Proposition. Every block basic sequence taken from a basic se-
quence in a Banach space is itself a basic sequence and has basis constant
no more than that of the original basic sequence.
PROOF. Suppose that (Yn) is a block basic sequence taken from a basic
sequence (xn). Then each Yn is nonzero by the uniqueness of expansions of
members of [{ Xn : n EN}] in terms of (xn). Let K be the basis constant
for (xn). Then IIE:'~lanxnll ::; KIIE:'~lanxnll whenever ml,m2 EN,
m1 ::; m2, and a1, ... ,a m2 E IF. It follows that the same is true if Xn is re-
placed by Yn, so (Yn) is basic by Corollary 4.1.25 and, by Proposition 4.1.20,
has basis constant no more than K.
= IITIIII~ anIlYnll-lT-lYnlll
~ IITIIIIT-ll1ll~ f3nllznll;;,lznlloo
~ MIITIIIIT- 1 11 max{ lf3nl : n = 1, ... ,m},
The final result of this section is the following variation on the fact that
every sequence in a Banach space that converges to zero weakly but not in
norm has a basic subsequence; see Theorem 4.1.32.
396 4. Schauder Bases
and
1= IIYm] II
= IIL(x~YmJ)xnll
11.
< Ilz}ll + 2- 1 ,
so Ilzj II > 1/2 > O. Therefore (zn) is a block basic sequence taken from (x n ).
The basis constant K(znl for (zn) can be no more than K(xnJ by Proposi-
4.3 Equivalent Bases 397
tion 4.3.16, which together with what has been proved above shows that
j j
j
1
2K(xnl
1
<---
- 2K(znl'
and therefore that limrn x~ (1IYrnll- 1 Ym) = 0 for each n. By the special case
proved above, some subsequence (1IYrn" 11- 1 Ym,,) of (1IYnll-1Yn) is a basic
sequence equivalent to some block basic sequence (zn) taken from (xn), so
(Yrn,,) is a basic sequence equivalent to the block basic sequence (1IYm" Ilzn)
taken from (xn).
In this book, the only need for the Bessaga-Pelczynski selection principle
will be for the proof in the next section that a Banach space has 1 em-
bedded in it if Co is embedded in its dual. However, the principle has many
other applications to Banach space theory. See [58] for several important
ones.
Exercises
4.32 Suppose that X is C[O, 1] or Lp[O, 1] such that 1 ::; p < 00. Show that X
has a basis whose terms are all polynomials. (However, the sequence of
polynomials that is the most obvious candidate to be a basis for C[O, 1]
is not even a basic sequence. See Exercise 4.13.)
4.33 Suppose that 1 < p < 00. Prove the best analog of Theorems 4.3.6
and 4.3.7 for p that you can devise.
398 4. Schauder Bases
4.34 (a) Give an example of a sequence (Xn) in a Banach space such that
(Xn) is not basic even though infnllxnll > 0 and there is a positive
constant M such that
I
II~ anX~11 :S (~anx~) (~f3nxn) 1+ 6
= I( ~ anx~) ( ~ f3n X n) I + 0
Consider again the example given in the second paragraph of this section.
It is true that (ei n) is not a basis for fi in the usual sense, but it is
a basis for ti: in a' certain weak* sense. Suppose that x* E fi and that
(an) is the member of 1'00 that is identified with x* in the usual way.
It is easy to check that limk L~=l anei,nx = x*x whenever x E 1'1, so
X* = w* -limk L~=l anei,n Since x*e1,n = an for every term e1,n of the
standard unit vector basis for 1'1, no sequence of scalars ((3n) besides (an)
has the property that x* = w*-limk L~=l (3nei,w Notice that (e1,n) (or,
more properly, the sequence (Qe1,n), where Q is the natural map from 1'1
into fi*) is the sequence of "coordinate functionals" for the "weak* basis"
(ei.n) for fi
As it turns out, the sequence of coordinate functionals for any Banach
space basis (xn) has this property of being a weak* basis whose sequence
of coordinate functionals is (xn) in the sense of the preceding paragraph.
for each positive integer m and each sequence (an) of scalars such that
w * - l'Imk ",k
Ln=l anX * . ts.
n eX1S
PROOF. Fix an x* in X*. For each element Ln (3nxn of X and each positive
integer k,
4.4 Bases and Duality 401
k
x*xm = lif 2: anx~xm = am
n=l
for each positive integer m, which proves the uniqueness assertion of the
theorem and also shows that (Qxm)(W*-limk 2::~=1 anx~) = am whenever
mEN.
With all notation as in Theorem 4.4.1, the last conclusion in that theorem
is, roughly speaking, that each Xm can be found in [{ x~ : n EN}]* as the
mth coordinate functional for the basic sequence (x~). Since (xn) is a basis
for X, this leads naturally to the question of whether [{ x~ : n E N }]*
might have embedded in it an entire copy of X. This turns out to be so.
ly*(~Ctnxn)1 = I~Ctnx*xnl
= Ix* (~Ctnxn) I
: ; II; anxnll
Theorem 4.4.1 and Proposition 4.4.4 suggest several further lines of in-
quiry. In light of Theorem 4.4.1, it is natural to ask for a simple condition C 1
on a basis (x n ) for a Banach space X that is necessary and sufficient for
the sequence (x~) of coordinate functionals for (xn) to be a basis for X*,
not merely a basic sequence. Similarly, Proposition 4.4.4 and its derivation
from Lemma 4.4.3 suggest the search for a simple condition C 2 on (x n )
that is necessary and sufficient for the isomorphism W of Lemma 4.4.3 to
map X onto [{ x~ : n E N }]*, which would guarantee that X is, at least
up to isomorphism, a dual space. It would be particularly interesting if
(x n ) were to satisfy both C 1 and C 2 In that case, the fact that (x n ) sat-
isfies C 1 guarantees that the map W of Lemma 4.4.3 is just the natural
map Q from X into X**, and the fact that (x n ) satisfies C 2 then assures
that Q maps X onto X**, that is, that X is reflexive.
The purpose of this section is to pursue these issues and to see what can
be learned about the structure of Banach spaces with bases along the way.
Most of the ideas and results of this section are due to R. C. James and
come from his 1950 article [109].
To bcgin thc search for a condition C 1 as described abovc, suppose that
(x~) is the sequence of coordinate functionals for a basis (x n ) for a Banach
space X. A few moments' thought about the role of (x n ) as the sequence of
coordinate functionals for (x~) as described in Theorem 4.4.1 shows that
whatever condition C 1 might be, the basis (xn) will have it if and only if
Ilx* - 2::=1(X*xn)x~lI-+ 0 as m -+ 00 whenever X* E X*. Now fix an x*
in X* and a positive integer m. For each member 2: n anXn of X,
I L,Bnx~11
n (m)
= f: ,Bnx~11 ::; I
I n=m+1 (m)
f:
n=m+l
,BnX~11
for each positive integer m, so limmll~n ,Bnx~ll(m) = O. The basis (xn) is
therefore shrinking.
Suppose conversely that (xn) is shrinking. Let K be its basis constant
and let x be a member of X. Then for each positive integer m and each
member ~n QnXn of X,
With all notation as in Lemma 4.4.3, the next item on the agenda is to
find a simple condition C2 on (x n ) that is necessary and sufficient for W to
map X onto [{ x~ : n EN}]*. The search will begin with a condition that
is in a sense dual to the shrinking condition.
4.4.9 Example. Let (en) be the standard unit vector basis for i p , where
1 S; p < 00. If (an) is a sequence of scalars such that sUPmll2::'=l anenll p
is finite, then
so 2:n ane n converges to the element (an) of i p . The basis (en) is therefore
boundedly complete. However, the standard unit vector basis (eo,n) for Co is
not, since 2:n eO,n does not converge even though sUPmll2::'=l eo,nlioo = 1.
The shrinking property for a basis implies bounded completeness for the
"dual basis."
n n
As one might have suspected from the way the plot has developed, the
search for the condition C z that began in the discussion following Propo-
sition 4.4.4 is about to end. Bounded completeness is that condition.
4.4.12 Lemma. Let X, (x n ), (x~), and I}i be as in Lemma 4.4.3. Then I}i
maps X onto [{ x~ : n E N }]* if and only if (xn) is boundedly complete.
PROOF. Suppose first that (Xn) is boundedly complete. Let y* be a member
of [{ x~ : n E N }]*. The goal is to find a member of X that I}i maps to y*.
Let K and K' be the basis constants for (xn) and (x~) respectively. Then
for each positive integer m and each member 2:n (Jnx~ of [{ x~ : n EN}],
= 1~(y*x~)(Jnl
= Iy* ( ~ (Jnx~ ) I
~ K' Ily* I IlL f3nx~
n
II
Notice that the first of the following two results is a partial converse
of Proposition 4.4.10, while the second is an analog of Theorem 4.4.11 in
which the roles of bounded completeness and the shrinking property have
been exchanged.
4.4.14 Theorem. Suppose that (xn) is a basis for a Banach space X and
that (x~) is the sequence of coordinate functionals for (xn). Then (xn) is
boundedly complete if and only if the basic sequence (x~) is shrinking.
PROOF. It was shown in the proof of Theorem 4.4.13 that the basic se-
quence (x~) is shrinking if (Xn) is boundedly complete. Suppose conversely
that (x~) is shrinking. Let W be the isomorphism of Lemma 4.4.12. Then
(Wxn) is the sequence of coordinate functionals for the shrinking basis (x~)
for [{ x~ : n EN}], and so is a boundedly complete basic sequence by
Theorem 4.4.11. It follows that (xn) is also boundedly complete.
The following result has already been suggested in the comments follow-
ing Proposition 4.4.4.
As has already been observed, the standard unit vector basis for Co is
not boundedly complete. It turns out that no basis for Co is boundedly
complete, and in fact much more than that can be said.
n=l n=l
= II~ 21Iz~llbanel,nlll
= II~ 21Iz~llbanS(llz~II-1T*QXkn)lll
= 2bIIST*Q(~QnXkn) III
S 2bIISIIIIT*QIIII~anXknll
Since sUPn Ilxkn I s
t, it follows from Theorem 4.3.6 that (Xk,.) is a basic
sequence equivalent to (el,n), so f1 is embedded in X.
With only a bit more work, it can even be shown that if a Banach space X
has Co embedded in its dual, then there is a complemented copy of fl em-
bedded in X, and X* has embedded in it a subspace isomorphic to foo
and complemented by a weak*-to-weak* continuous projection. An expo-
sition of this result of Bessaga and Pelczyriski [26] can be found in Joseph
Diestel's book [58, pp. 48-49].
The preceding result has an analog for f1 and the shrinking property.
4.4 Bases and Duality 409
The rest of this section is devoted to the shrinking and bounded com-
pleteness properties for unconditional bases. The main result for this study
is the following one.
112:::::'=l,Bnxnll ::; M for each m but 2:::n ,Bnxn does not converge, so there
are sequences (Pn) and (qn) of positive integers and a positive 6 such that
P1 ::; q1 < P2 ::; q2 < P3 ::; q3 < ... and I12:::J:Pn {3jxjll 2': 6 whenever
n E N. Let Zn = ~J:Pn,BjXj for each n. Then (zn) is a block basic se-
quence taken from (xn). It follows from Proposition 4.2.18 that if mEN
and a1, ... ,am Elf, then
Exercises
4.40 Prove that C[O,l] is not isomorphic to the dual space of any normed
space.
4.41 Suppose that X is a Banach space with an unconditional basis. Prove that
if X is separable, then X is reflexive if and only if Co is not embedded
in X.
4.42 Suppose that X is a Banach space with an unconditional basis. Prove
that X is separable if and only if X is reflexive. Exercise 4.41 may help.
4.5.1 Lemma. Suppose that (xn) is a basis for a Banach space. Let
while
It follows that for each positive integer m there is a scalar am such that
limj am,j = am.
Suppose that f > O. Let j. be a positive integer such that if j, j' 2:: jf'
then II(an,j) - (an,],) lis < Eo If mEN and j, j' 2:: j., then
whenever j 2:: j,. It follows that (an) E S and that II(a n ,)) - (an)lls -> 0
as j -> 00, so Cauchy sequences in S do converge.
IX'(~(x'*x~)xn)1 = IX**(~(X'xn)x~)1
::; IIX*'llll~(x*xn)x~11
::; II x " 1111x' I
since x* = L:n(x*xn)x;,. Therefore IIL::=l(x'*x~)xnll ::; Ilx"1I when-
ever x" E X** and mEN, and so (x") = (x*'x~) E S whenever
*4.5 James's Space J 413
I f:
n=m,+l
anX*Xnl s: Ilx*lI(m,) I! f:
n=m,+l
anXnl! s: 21I x *lI(m,) II(an)lls,
which implies that l:n anx'xn converges since limmllx*ll(mJ = O. Define
xo': X' ---> IF by the formula xo*(x*) = Enanxxn. Then x;)* is clearly
linear. If x* E X* and mEN, then
so Ixo*x*1 s:
II(an)lIs Ilx'll whenever x' E X*. It follows that xo* E X'*
and x;)* = (xo*x~) = (an), so maps X** onto S.
Notice that the argument just given, along with the fact that is one-to-
s:
one, shows that Ix*'x* I II <px *' 11 s Ilx' II whenever x* E X'* and x* E X*,
and therefore that II x** II :s 11 <px** II s whenever X** E X *'. The reverse
inequality was obtained earlier, so is an isometric isomorphism from X**
onto S. The fact that (xn) is monotone then assures that
and its first and second dual spaces can be treated as if they were sequence
spaces as in Table 4.1, where the equalities of limits and suprema in the
norm column are due to the fact that (xn) and (x~) are both monotone,
the first by hypothesis and the second by Theorem 4.4.1.
Suppose that the sequences of scalars (an), (3n) , and bn) have been
identified with the members x** , X* , and x of X** , X* ) and X respectively.
Then the actions of (an) on (3n) and of ((3n) on bn) are given by the
formulas
and
since
n n
*4.5 James's Space J 415
and
Furthermore, if Q is the natural map from X into X** and the identifica-
tions of Table 4.1 are made, then it is clear from the formulas for (a n )(f3n)
and (f3n)(,n) given above that Qbn) = bn) for each member bn) of X.
Now suppose that X is Co and that (xn) is its standard unit vector
basis, which is indeed monotone and shrinking. Let W, Y, and Z be the
sequence spaces that the correspondences of Table 4.1 identify with Co, co,
and co* respectively. It is easy to check that Wand Z are just Co and =,
respectively, with their usual norms, and an easy argument based on the
standard way that Co is identified with 1 shows that Y is just 1 with its
usual norm. Furthermore, the identification of Co with itself given in the
table is just the identity map, while the identifications of cO and co* with 1
and 00 respectively that follow from the table are precisely the standard
identifications of these spaces that have been discussed previously in this
book. 3 Thus, the identifications of Table 4.1 are generalizations of the usual
characterizations of the first and second duals of Co to other Banach spaces
having monotone shrinking bases.
The identifications described in Table 4.1 will prove useful for analyzing
the space about to be defined.
4.5.3 Definition. For each member (an) of the vector space V underlying
real Co, let
and
Then James's space J is the real Banach space of all members (an) of V
such that II(an)lla is finite, with the norm 1IlIa.
Though the function II lib is not actually used to define J, this function
will have important uses in the study of the space to be done below. Of
3See Example 1.11.2 for the standard identification of co' with i=. It is easy to check
directly that if that identification has been made and i1 has been identified with Co in
the natural way, then (a n )(l'n) = En anl'n whenever (an) E co'
and (I'n) E co.
416 4. Schauder Bases
< E.
*4.5 James's Space J 417
whenever j 2: j" m 2: 2, and PI < ... < Pm. It follows that ((3n) E J and
that
whenever) 2: )E' Since ((3n,j) -+ ((3n) as) -+ 00, the space J is complete .
Let (en) be the sequence of standard unit vectors in real Co. Then each
of these vectors is in J, and Ilenll a = 1 for each n. The sequence (en) turns
out to be a monotone shrinking basis for J just as it is for Co.
m(j)-l
L ((3Pn,j - (3Pn+l,j)2 + ((3P=U),j - (3Pl,j)2 2: Eo
n=1
when) 2: )0' But a moment's thought about the definition of 11lla then
shows that 11((3n)lIa = +00, a contradiction. This finishes the proof that
(en) is a basis for J and verifies the formula for basis expansions.
All that remains to be shown is that the monotone basis (en) is shrinking.
Suppose it is not. Then there is an x* in J*, a positive 6, and a normalized
block basic sequence (un) taken from (en) such that x*u n 2: 6 for each n.
418 4. Schauder Bases
Therefore En x' (n- l u n) does not converge, so En n-1u n does not con-
verge. To obtain a contradiction to this and thereby finish the proof of the
proposition, it is enough to show that there is a positive constant c such
that IIE:~m, n- l u n llb ::; c(E:~ml n- 2)1/2 whenever m1 ::; m2
Let ((n) be a sequence of real numbers and (qn) a sequence of positive
integers such that 1 = ql < q2 < q3 < ... and Uj = E~~~J-1 (nen for
eachj. Notice that I(nl ::; 1 for each n since Ilujlla = 1 for eachj. A positive
integer P will be said to belong to a term Uj of (un) if qj ::; P < Qj+l
Let positive integers ml and m2 be such that m1 ::; m2, and let bn)
be the element E:~Tn' n-Iu n of J. Fix a positive integer m and positive
integers PI, ... ,Pm such that m :2: 2 and PI < ... < Pm, then let s =
E::l\"(Pn - I'Pn+I)2. Let It 1 be the collection of summands bPk - "YPk+ I?
of s such that Pk and PHI both belong to the same term of (un), that
is, such that there is some j for which qj ::; Pk < PHI < Qj+I' For each
member bPk - 'IPk+l? of Itl and the Uj to which it belongs, either
(I'Pk - 'IPk+l )
2
= ( .-I(
J Pk - J
.-I(
Pk+1
)2
= J'-2(r'>Pk -
r
'>Pk+1
)2
L n-21Iunll~ L n-
m2 ffi2
81 ::; 2 = 2 2.
n=Tnl
Now let ([.2 be the collection of summands bPk - 'IPk+!)2 of s such that Pk
and PHI do not belong to the same term of (un), and let 82 be the sum of
all members of ([.2. Fix a member bPk - "YPk+I)2 of ([.2 and let uj, and uj,
be the terms of (un) to which Pk and PHI, respectively, belong. Then "YPk
is either jl I (Pk or 0, and similarly "YPk+l is either j:;I(Pk+l or O. It follows
that
("YPk - 'IPk+l )
2<
-
('-I()2
11 Pk + (.-1(
12 Pk+l )2 + 211-1 12.-11(Pk '>Pk+1
r I
::; j1 2 + j:;2 + 2jl1j:;1
::;j1 2 +j:;2+ j1 2 +j:;2
= 2j1 2 + 2j:;2.
L
Tn2
82 ::; 4 n- 2 .
n=ml
*4.5 James's Space J 419
Therefore
As has already been shown, this is a contradiction that implies that (en)
is shrinking.
In fact, the space J is already precisely that space; in particular, the norm
given by (4.2) is precisely 11lIa' Furthermore, the space J** can be viewed
as the space of all sequences (an) of scalars such that (2:.::'=1 anen)~=1 is
bounded in J, with the norm also given by (4.2). For consistency, this norm
will be denoted by 1IlIa as for J. As an aid to understanding, in most of the
rest of this section the space J** will be treated as if it actually were this
sequence space rather than just isometrically isomorphic to it, and J will
often be viewed as the subspace of J** with which the natural map from J
into J** identifies it; again, see the comments following Proposition 4.5.2.
The reader wishing to rewrite the following arguments to include all of the
appropriate isometric isomorphisms will have no trouble doing so.
Suppose that (f3n) is a sequence of reals such that limn f3n does not exist.
Then there must be a positive f; and sequences (Pn) and (qn) of positive
integers such that 1 ~ PI < ql < P2 < q2 < ... and such that If3Pn -f3qn I 2: f;
for each n. For every positive integer m,
Though 11lla is being used to denote the norm of J**, that does not
mean that the formula of Definition 4.5.3 can be used to compute it. For
example, it is easy to see that every constant sequence (e, e, e, ... ) of real
numbers is in J** and has norm lei by (4.2), but the formula from Def-
inition 4.5.3 would incorrectly give 0 for the value of the norm. For the
moment, let the "norm" of a member (an) of J** computed using the for-
mula in Definition 4.5.3 be denoted by l(an)la' and let II(an)lla continue
to represent the actual value of the norm of (an). For each (an) in J**,
whenever m,Pl, ... ,Pm E N, m 2': 2, and PI < ... < Pm, from which it
follows that l(an)la ~ II(an)lla and in particular that l(an)la is finite. One
immediate consequence of this is that
since the only way that a member of J** can avoid being in J is to have a
nonzero limit.
Incidentally, the formula for 11lla of Definition 4.5.3 can be modified a
bit so that the resulting slightly more complicated formula does also work
to compute the norms of members of J**. See Lemma 4.5.8.
Let eo = (1,1,1, ... ). Then eo E J** and lIeolla = 1. It will now be
shown that (eO,el,e2, ... ) is a basis for J**. Fix an element hn) of J**
and let 'Y = limn 'Yn. Then limnhn - 'Y) = 0, so hn) - 'Yeo E J and
hn) - 'Yeo = L:~=1 hn - 'Y)en . Therefore
CXl
4.5.7 Theorem. Let Q be the natural map from J into J**. Then the
co dimension of Q(J) in J*' is 1. Thus, the space J is not reflexive.
(4.3)
or
m-2 ) 1/2
2- 1/ 2 ( "(a
~ Pn
- a Pn+l )2 + a2
PTn-l
+ a2
PI
( 4.4)
n=l
for some positive integers m, PI, ... ,Pm such that m 2: 2 and PI < ... < Pm.
(By convention, the sum from 1 to m - 2 is considered to be 0 if m = 2.)
422 4. Schauder Bases
PROOF. For each positive integer k, let I{l,oo.,k} be the indicator function
of the set {I, ... ,k}. Then
1/2
+ (apmI{l, ... ,k} (Pm) - a p, )2 )
A moment's thought shows that this last supremum is precisely the one
described in the statement of this lemma.
For the duration of this paragraph, let (an) be a fixed member of J**. It
is easy to see that II(O,al,oo.,am,O,O,oo.)lla = lI(al,.oo,a m ,O,O,oo.)lla
for each m, from which it follows that
and therefore that the "shifted" sequence (0, a1, a2, ... ) is in J**. Sub-
tracting (limn an)eo from this shifted sequence yields the member
of J** , which is in fact in J since the limit of this sequence is 0. If the norm
of this member of J is computed using the formula of Definition 4.5.3 (with
the 2- 1/ 2 brought inside the supremum), then the set whose supremum is
being taken is precisely the set whose supremum is used to compute II (an) Iia
in Lemma 4.5.8, so the above member of J has the same norm as (an).
Define T: J** - t J by the formula
The ideas and results of this section are essentially from R. C. James's
papers [109J and [110J that appeared in 1950 and 1951 respectively. In the
earlier paper, James constructed a Banach space that is almost J and is
isomorphic to it; however, the norm is not quite right for the space to be
isometrically isomorphic to its second dual, so James had to settle for an
isomorphism. By tweaking the norm a bit in the later paper, he was able
to obtain the stronger result.
The space J proved useful as a counterexample for several long-standing
conjectures. As was discussed at the end of Section 1.11, a comment in
Banach's book [13J can be interpreted to be the question of whether a sep-
arable Banach space must be reflexive if there is any isometric isomorphism
whatever from the space onto its second dual. The space J shows that the
answer is no. Also, the separability of J** shows that J is a counterexample
for the conjecture that a Banach space must be reflexive if its second dual
is separable. 4 See Exercises 4.44 and 4.45 for two other conjectures settled
in the negative by J. A list of references for other applications of J can be
found in [118, p. 633J.
Exercises
4.43 Prove that J has no unconditional basis.
4.44 Prove that J is an example of an infinite-dimensional Banach space X
such that X is not isomorphic to X ED X. Exercise 3.12 could help.
4.45 Prove that J is an infinite-dimensional real Banach space for which there
is no complex Banach space X such that J is isomorphic to the real
Banach space Xr obtained from X by restricting multiplication of vectors
by scalars to lR x X. It might help to look at Exercise 3.12 as well as
Proposition 1.13.1 and its proof.
4 However , it was not quite the first such counterexample. That honor went to J's
isomorphic cousin constructed in [109], as James observed in that paper.
5
Rotundity and Smoothness
When thinking of the closed unit ball of a normed space, it is tempting to vi-
sualize some round, smooth object like the closed unit ball of real Euclidean
2- or 3-space. However, closed unit balls are sometimes not so nicely shaped.
Consider, for example, the closed unit balls of real ei and e~. Neither is
round by any of the usual meanings of that word, since their boundaries,
which is to say the unit spheres of the spaces, are each composed of four
straight line segments. Also, neither is smooth along its entire boundary,
since each has four corners. These features of the closed unit balls have
a number of interesting consequences that cause the norms of these two
426 5. Rotundity and Smoothness
spaces to behave a bit unlike that of real Euclidean 2-space. For example,
if Zl and Z2 are different points on anyone of the four sides of one of these
balls, then 1 = Ilzlll = IIz211 = ~llzIil + ~llz211 = II~Zl + ~z211, so equality
is attained in the inequality IlzI + z211 ~ IIzIl1 + IIz211 despite the fact that
neither Zl nor Z2 is a nonnegative real multiple of the other. Furthermore,
the presence of the corners leads to the existence of multiple norming f1LnC-
tionals for some points Z of the unit sphere of each of these spaces, that is,
norm-one members z* of the dual space such that z*z = Ilzll. To see why
this would be, let Z be either of these two spaces and let Zo be one of the
four corners of B z. Then there are infinitely many different straight lines
that pass through Zo without intersecting the interior of the closed unit
ball; let hand l2 be two of them. By Mazur's separation theorem, there
are members zi and Zz of Z*, necessarily different, such that if j E {1,2},
then zj z = 1 when Z E lj and zjz ~ 1 when z E B z It follows readily
that zi and Zz are both norming functionals for Zo. As will be seen, it is
precisely the presence of the corners or sharp bends in the unit sphere that
caused this multiplicity of norming functionals for elements at the locations
of the bends.
The purpose of this chapter is to study the special properties of normed
spaces whose closed unit balls are round, in the sense that the unit spheres
include no nontrivial line segments, and of those whose closed unit balls
are smooth, in the sense that the unit spheres have no corners or sharp
bends. Roundness will be taken up first.
5.1 Rotundity
The property about to be defined was formulated independently by James
Clarkson and Mark Krein. Clarkson was particularly interested in the uni-
form version of this property that will be studied in the next section, while
Krein used it in joint work with Naum Akhiezer on the moment problem.
There is an entire host of equivalent ways to define this property, some of
which are given in this section.
[47]-[52J on rotundity and uniform rotundity between 1941 and 1957. 1 The
term strictly normed is Krein's.
Definition 5.1.1 actually says that a normed space is rotund when its
unit sphere includes no nontrivial straight line segments, but some arguing
needs to be done to show that. Suppose that X is a normed space. Just
for the remainder of this paragraph, let (XljX2) denote the "open line
segment" {tXl +(I-t)x2 : 0 < t < I} whenever Xl,X2 EX. If X is rotund
and Xl and X2 are different points of its unit sphere, then the definition
of rotundity assures that (XljX2) lies entirely in the interior of Ex, so no
nontrivial line segments lie in S x. Suppose conversely that no nontrivial line
segments lie in Sx and that Xl and X2 are different points of Sx. Then some
of the points of (XljX2) lie in B x, so it follows easily from Lemma 2.2.18
that all of the points of (Xl j X2) lie in Ex. The space X is therefore rotund.
Incidentally, the last portion of this argument also establishes the following
characterization of rotundity that allows the verification of the property
by examining only the midpoints of straight line segments rather than the
entire segments.
5.1.3 Example. The scalar field IF, viewed as a normed space over IF, is
obviously rotund. More generally, it is easy to see that every normed space
that is zero- or one-dimensional is rotund.
1 However, Day referred to the property as strict convexity in those papers, as most
others did before and, admittedly, have since. The term rotund first appeared in his 1958
book [53J.
428 5. Rotundity and Smoothness
from which it follows that neither Ll(n, 2:, J.t) nor Loo(n, 2:, J.t) is rotund.
In particular, the spaces 1\ and 00 are not rotund, and f and ~ are not
rotund when n ~ 2. Notice that f and ~ are rotund when n < 2 by
Example 5.1.3.
5.1.6 Example. Let el and e2 be the first two standard unit vectors of Co.
Let Xl = el + e2 and X2 = el - e2. Then
5.1.8 Example. The evidence presented so far might lead one to conjec-
ture that rotund Banach spaces are always reflexive. The purpose of this
example is to construct a nonreflexive rotund Banach space by finding a
rotund norm for 1 that is equivalent to its original norm. This construc-
tion might seem to have some unnecessary complications in it, but the new
norm is being designed to have some special properties that will be useful
in Example 5.1.22.
The first step is to define some functions. For each positive integer m
and each nonnegative real number t, let
f rn () _
t -
t2 + mt .
m+1
Then for each Tn,
(1) the function 1m maps [0,1] continuously onto [0,1]' and 1m(0) =
and 1rn(1) = 1;
(2) fm, 1:n , and 1:'n are all positive on (0, (0), so in particular the function
1m is strictly increasing on [0, (0);
(3) m~l t :S 1m(t) :S t whenever t E [0,1], with strict inequality when
t E (0, 1);
5.1 Rotundity 429
(4) fm{st l + (1- s)b) < sfm(h) + (1- S)fm(t2) whenever hand t2 are
different members of [0, (0) and 0 < s < 1; and
(5) Ifm(td - fm(t 2 )1 ~ max{ f:n(t) : 0 ~ t ~ I} It 1 - t21 ~ ~ It 1 - t21
whenever t l , t2 E [0,1].
Let {Am: mEN} be a collection of pairwise disjoint infinite subsets of N
whose union is N, and for each positive integer n let m(n) be the index m
of the set Am containing n. Let
Notice that if (an) E C, then lanl ~ 1 for each n. Suppose that (((n,j));:l
is a sequence in C that converges to some member ((n) of 1. For each j,
12: n
fm(n) (I(n,jl) - 2:
n
fm(n) (I(nl) I~ E!fm(n)(I(n,jl) - fm(n)(I(nl)!
n
:s ~ L II(n,jl- I(nll
n
n
= ~1I((n,j) - ((n)lIl,
n n
~llxI11:::; Ilxll r :::; IIxl11 whenever x E 1, so the norms 11111 and 11llr are
equivalent. Now
by Proposition 1.9.14, which together with the fact that C is closed in the
common topology of 1 and i,r shows that C is the closed unit ball of l,r'
The equivalence of 11111 and 11llr shows that l,r is a nonrefiexive Ba-
nach space, so all that remains to be proved is that the norm 11llr is
rotund. Suppose that ((3n) and ("'fn) are different elements of Se"r' Let
(I-n) = ~ ( ((3n) + ("'fn))' It is enough to find a real number a greater than 1
such that a(l-n) E C, for this will imply that Ila(l-n)llr :::; 1 and there-
fore that II (I-n) liT < 1. To this end, let no be a positive integer such that
(3no =I "'fno' Then either \~((3no + "'fno) \ < ~1(3nol + ~hnol or l(3nol =I hnol
It follows that one of the first two inequality symbols in the inequality
n n
= ~(a - 1)11(l-n)111,
One consequence of Examples 5.1.4 and 5.1.5 (which, for the real case at
least, is also apparent from diagrams of the corresponding unit spheres) is
that Euclirlean 2-space is rotund while i is not, even though the two spaces
are isomorphic. Thus, in contrast to most properties of normed spaces that
have been studied so far, rotundity is not isomorphism-invariant. The fol-
lowing is the most that can be said along these lines. The proof is obvious.
5.1 Rotundity 431
2~ II Xl + yll
= IIXI +X2 - (1-llx211-I)X211
~ II Xl +X211- (1-llx211- 1 )ll x211
= IIXIII + IIx211-llx211 + 1
= 2,
from which it follows that II ~(Xl + y) II = 1. Since Xl, Y E Sx, the rotundity
of X requires that Xl = Y = Ilx211-IX2. This shows that (a) '* (b).
Suppose conversely that (b) holds. Let Zl and Z2 be different members
of Sx. Then neither of the two vectors is a nonnegative multiple of the
other, which implies that Ilzl + z211 < Ilzlll + IIz211 = 2 and therefore that
11~(zl + z2)11 < 1. The space X is therefore rotund, so (b) '* (a).
It should be noted that for any normed space whatever and any two
members Xl and X2 ofthe space such that one of the vectors is a nonnegative
multiple of the other, it is true that IIxl + x211 = IlxIiI + Ilx211. Thus, the
preceding proposition yields the following strengthened form of the triangle
inequality for rotund normed spaces.
Since a member x' of the unit sphere of the dual space of a normed
space X supports B x at some point x of S x if and only if Re x x = x x = 1,
the following result is essentially just a restatement of the preceding one.
5.1.19 Corollary. (M. M. Day, 1941 [47]). If a normed space is rotund and
reflexive, then each of its nonempty closed convex subsets is a Chebyshev
set.
436 5. Rotundity and Smoothness
d(xo, C) :::; Ilyo - xoll :::; liminf llYn; - xoll = d(xo, C).
J
Therefore there is at least one point of C closest to Xo. Notice that the
rotundity of X has not yet been used.
Finally, the rotundity of X implies that C is a set of uniqueness, so
there is no other point of C besides Yo closest to Xo. Thus, the set C is
Chebyshev.
As a trivial consequence of this and the fact that every normed space
is a subspace of itself, a normed space is rotund if and only if each of its
subspaces is rotund. The following is a far better result along these lines.
.
of absolute value 1 such that ~ (1 + a) also has absolute value 1, which
would contradict the rotundity of IF. It follows that Xl and X2 are linearly
independent. Therefore ({Xl, X2}) is a two-dimensional subspace of X that
~~~~.
IIT(an)1I = IlL n
ang(n)11
~ L( L
m nEA~
lanl m: 1)
~ L(
m
L
nEA~
fTT1(la n l))
= L fm(n) (Ianl)
n
~ 1,
as claimed. Notice that this also shows that T is bounded, since C = B i,r .
Let Ui,,, and Uy be the open unit balls of l,r and Y respectively. It will
now be shown that T(Uil r) 2 Uy. Suppose that y E Uy. Since 112yll < 2,
the open balls of radius l' centered at 2y and the origin intersect (at y, for
438 5. Rotundity and Smoothness
114y - 2d l - d2 11 < 1.
Since 118y - 4d l - 2d2 11 < 2, there is a d3 in D such that
118y - 4d l - 2d2 - d3 11 < 1.
Continuing in the obvious fashion produces a sequence (d n ) in D such that
Now suppose conversely that both Xl and X 2 are rotund. Let (Xl,X2)
and (YI,Y2) be different members of SX, fBX 2 The proof will be finished
once it is shown that 11~(xl + YI,X2 + Y2)11 < 1. Notice that
If either Ilxlll =l-IIYlll or IIx211 =I- IIY211, then it follows from the rotundity of
real ~ that
Exercises
5.1 Suppose that (Xn) is a sequence of normed spaces and that 1 ~ P < =.
Then the Cp sum lp((Xn)) of the sequence (Xn) is the collection of all
sequences (xn) such that Xn E Xn for each nand LnllxnllP is finite,
along with the obvious vector space operations and the norm defined by
the formula
Prove that all of this really does define a normed space, and that lp (( X n))
is a Banach space if and only if each Xn is a Banach space.
440 5. Rotundity and Smoothness
5.2 With all notation as in the preceding exercise, suppose that 1 < p < 00.
Prove that fp((Xn)) is rotund if and only if each Xn is rotund. You
may use the fact about the rotundity of certain Lebesgue spaces that is
mentioned in Example 5.1.4.
5.3 Prove that if K is a compact Hausdorff space having more than one
element, then rca(K) is not rotund.
5.4 This exercise assumes some knowledge of inner product spaces. See, for
example, [24J (in which such spaces are called pre-Hilbert spaces) or [202J.
Prove that every inner product space is rotund.
5.5 Suppose that X is a normed space. Prove that the following are equivalent.
(a) The space X is rotund.
(b) Whenever (Xn) is a sequence in Sx and there is a member x of Sx'
such that limn x'X n = 1, all weakly convergent subsequences of (Xn)
have the same limit.
(c) Whenever (Xn) is a sequence in Sx and there is a member x of Sx'
such that limn x'xn = 1, all convergent subsequences of (Xn) have
the same limit.
5.6 (P. R. Beesack, E. Hughes, and M. Ortel, 1979 [22]). Suppose that X is
a complex normed space. Prove that the following are equivalent.
(a) The space X is rotund.
(b) For each pair Xl and X2 of different members of Sx, there is a
scalar a X" X2 such that Ilaxl,x2xl + (1 - ax" x2)x211 <
1.
5.7 Prove that a normed space is rotund if and only if no two closed balls in
the space having disjoint interiors intersect at more than one point.
5.8 If C is a nonempty closed convex subset of a Hausdorff TVS X, then an
element Xe of C is an exposed point of C if there is an x in X' such that
Re x is bounded from above on C and attains its supremum on C at Xe
and only at Xe. (This definition was previously given in Exercise 2.112, in
which it was shown that the property of being an exposed point is properly
stronger than that of being an extreme point.) Prove that a normed space
is rotund if and only if every point of its unit sphere is an exposed point
of its closed unit ball.
5.9 Prove that if X is a Banach space that is not zero- or one-dimensional,
then there is a nonrotund norm on X equivalent to the original norm.
5.10 Suppose that X is a normed space and that there is a one-to-one bounded
linear operator T from X into a rotund normed space Y. Prove that the
formula IIXllr = Ilxll + IITxl1 defines a rotund norm for X equivalent to its
original norm 1111.
5.11 This exercise requires James's theorem from either of the optional Sections
1.13 and 2.9.
(a) Prove that if every nonempty closed convex subset of a Banach space
is a set of existence, then the space is reflexive.
5.2 Uniform Rotundity 441
(b) Prove that a Banach space is rotund and reflexive if and only if each
of its nonempty closed convex subsets is a Chebyshev set.
(c) It is a 1976 result of Jorg Blatter [30J that a normed space is complete
if every nonempty closed convex subset of the space has an element
of minimum norm. Use this result to improve parts (a) and (b) of
this exercise by replacing "Banach space" with "normed space."
5.12 (a) Show that every reflexive subspace of a normed space is a set of
existence.
(b) Show that every nonempty weakly compact subset of a normed space
is a set of existence.
(c) Show that every nonempty weakly* compact subset of the dual space
of a normed space is a set of existence.
5.13 (a) Prove that if M is a subspace of a rotund normed space X and M
is a set of existence in X, then XjM is rotund. (Notice that the fact
that M is a set of existence implies that it is closed. Why?)
(b) From parts (a) of this exercise and Exercise 5.12, conclude that if
M is a reflexive subspace of a rotund normed space X, then XjM
is rotund. (This is a 1959 result of Victor Klee [137J.)
if E = 0;
ifO<E:S2
442 5. Rotundity and Smoothness
It is easy to see that for each normed space X, the modulus of rotundity
is a nondecreasing function of f such that {j x (0) = O. It is also clear that if
M is a subspace of X, then {j M (f) ~ 8x (f) when 0 ::; f ::; 2. Furthermore,
if X is a real one-dimensional normed space, then
if f = 0;
if 0 < f ::; 2,
which means that the definition of b{O} given above is exactly what is needed
to avoid having exceptions for the zero-dimensional case to the properties
given at the beginning of this paragraph.
The first portion of this section is devoted to obtaining Theorem 5.2.5,
which gives some alternative formulas for the modulus of rotundity that
are often used as its definition. These formulas are not actually used in any
crucial way later in this book. The main reason for developing them is to
give some examples of a certain type of argument involving the geometry
of two-dimensional real normed spaces that is often encountered in the
theory of normed spaces but that has not yet been used in this book.
A generous amount of detail will be provided to illustrate how much care
must be taken to avoid making unwarranted assumptions about the relative
positions of objects such as lines and points in the plane when carrying
out such an argument, as well as other unjustified assumptions based on
possibly unsupported intuition about the shape of closed unit balls.
Xo 0
(a) (b)
they are in Figure 5.1 (a), the point Yo lies on or above xo, -Xo. If Yo were
strictly to the left of Xo, Xl, then it would follow that either Yo E Wo or
YI E WI' However, either case would place one of the two points Xo and Xl
in the "interior" of a straight line segment with one endpoint in Bx and
the other in B1:, which would imply that either Xo or Xl is in B1:. This
contradiction shows that Yo lies on or to the right of xo, Xl. If Yo were to
lie on this line, then an easy argument involving the location of YI would
show that IlxI11 < 1, another contradiction. Therefore Yo lies strictly to the
right of XO,XI'
Since -Xo and -Xl are different elements of Sx, it follows that the points
of -xo, -Xl lying on or above Xo, -Xo all have norm at least 1, so no point
on Xo, Yo strictly between Xo and Yo can intersect -Xo, -Xl, nor can Yo
lie on that line; see Figure 5.1 (b). Thus, the point Yo lies strictly to the
left of -xo, -Xl' Since xo, Xl, 0, Xl - Xo, and -xo, -Xl are parallel, it also
follows that ~ (xo+yo), denoted by mo in Figure 5.1 (b), lies strictly between
xo, Xl and 0, Xl - Xo. Therefore 0, ~ (xo + Yo) intersects xo, Xl somewhere
on or above Xo, -Xo and at a point of 0, ~ (xo + Yo) that is reached by
traveling from 0 to ~ (xo + Yo) and then beyond.
The point to the preceding argument is that there are real numbers S
and t such that s > 1 and t 2': 0 that satisfy the equation
S
-(xo
2
+ Yo) = Xo + t(XI - xo).
If t ::; 1, then
x y
~(x-y) ~(y - x)
o
-y -x
Therefore Ilxn - Ynll --+ 0, which establishes that (b) :::;. (d).
Now suppose that (d) holds and that (xn) and (Yn) are sequences in Bx
such that 11~(Xn + Yn)11 --+ 1. Since 11~(Xn + Yn)11 ~ Hllxnll + IIYnl!) for
each n, it follows that Ilxnll --+ 1 and llYn II --+ 1, so Ilx n - Ynll --+ 0 by (d).
Therefore (d) =} (c), from which it follows that (b) {:} (c) {:} (d).
Suppose that X is uniformly rotund and that (xn) and (Yn) are se-
quences in Sx such that II ~(Xn + Yn)11 --+ 1 but Ilxn - Ynll does not tend
to O. Let Ox be the modulus of rotundity of X. It follows that there is a
subsequence (x nj ) of (xn) such that Ilxnj - Ynj II ;::: f for some positive f
and each j, which implies that 11~(Xnj +Ynj)1I ~ 1- OX(f) for each j, a
contradiction. Therefore (a) =} (b).
Finally, suppose that X is not uniformly rotund. Then there is an f such
that 0 < f ~ 2 but OX(f) = O. Therefore there are sequences (xn) and (Yn)
in Sx such that Ilxn -Ynll;::: f for each n but 11~(Xn+Yn)II--+ 1, so (b)
does not hold. This shows that (b) =} (a).
One very important class of uniformly rotund Banach spaces is that of the
spaces Lp(n, I:, J1.), where J1. is a positive measure on a a-algebra I: of subsets
of a set nand 1 < p < 00; notice that this includes the spaces lp and i;
such that 1 < p < 00 and n is a nonnegative integer. The uniform rotundity
of these spaces was first established in 1936 by James Clarkson [41] in the
same paper in which he introduced the notions of rotundity and uniform
rotundity. Proofs of this usually begin by establishing some inequality or
inequalities in the scalar field or for the norms of elements of Lp(n, I:, J1.),
and often require the separate consideration of the cases in which 1 < p :S 2
and 2 ~ P < 00; see, for example, Clarkson's paper or [21]. The proof
given here is from [141] and is based on a method from a 1950 paper by
E. J. McShane [166] that does not require the consideration of the two
different cases.
The following lemma is proved in quite a bit more generality than is
needed for the proof of Clarkson's result, but the extra generality will be
required later for the proof of Theorem 5.2.25.
5.2.9 Lemma. For each p such that 1 < p < 00 and each function
A: (0,2] --+ (0,1], there is a function 'Yp,>.; (0,2] --+ (0,1] such that if X
is a uniformly rotund normed space whose modulus of rotundity Ox has
the property that A(E) :S OX(E) when 0 < f ~ 2, then
whenever 0 < t ~ 2 and x and yare members of X such that Ilx - yll ;:::
tmax{llxlI,IIYII}
PROOF. Suppose to the contrary that there were a p such that 1 < p < 00
and a function A: (0,2] --+ (0,1] for which no such function 'Yp,A exists.
5.2 Uniform Rotundity 449
Let f(t) = (~(1 + t))V /(~(1 + tv)) when 0:::; t :::; 1. It is an easy calculus
exercise to show that f strictly increases on [0,1) to its maximum value
of 1, from which it follows that if X is a normed space and x and yare
members of X such that Ilxll = 1 and lIyll :::; 1, then
Notice that
. 11~(Xn+yn)llv
(5) h,;n Hllxnllp + IIYnllv) = 1.
It is an easy consequence of what has been proved about f that IIYnl1 -+ 1,
so in particular it may be assumed that no Yn is zero. Let Zn = IIYnll-1Yn
for each n. Then Ilzn - Ynll -+ 0, so it may be assumed that Ilxn - znll ~ ~t
for each n. It follows that
limll~(xn+zn)11
n
=limll~(xn+Yn)11
n
= 1,
a contradiction.
With all notation as in the statement of the preceding lemma, letting >.
be the restriction of the modulus of rotundity of the scalar field to (0,2]
produces the following special case of the lemma, which is what is actually
needed for the proof of Theorem 5.2.11.
5.2.10 Lemma. Suppose that 1 < p < 00. Then there is a function
Ip: (0,2] -+ (0,1] such that
P
la; f3 I :::; (l- rp (t) Ca IP ;lf3I P )
when 0 < t :::; 2 and a and 13 are scalars such that la - 131 ~ t max{lal, If3l}.
450 5. Rotundity and Smoothness
whenever w E 0, so
l-ii~(h+h)ii== 1nChI P
;lhI
P
_Ih;h/P)d{t
2: i ChiP; -I h; 12 I d{tIhl P
P
)
>
- "fp
(_E)
41/p
r Ihl P+ Ihl P
JA 2
d
{to
1
JO\A
2: fP - E
P
(lhl P + IhI P ) d{t
4 O\A
EP
2: EP - 4(llhll~ + Ilhll~)
EP
2'
from which it follows that max{llhlAllp, IlhlAllp} 2: E/(2 21 / P ). Therefore
1
l-ii'2(h + h)ii p 2:
P
"fp
(f)
4
IlhlAII~ + IlhlAII~
1/ p 2 2: "fp
([)
41/p
fP
2 P+ 2 '
and so
5.2.12 Corollary. Suppose that 1 < p < 00. Then fp is uniformly rotund,
as is f; whenever n is a nonnegative integer.
It is not difficult to check that this defines a Banach space; see Exercise 5.1.
Suppose that (Xn ,l), (Xn ,2), H(xn,d + (X n ,2)) E Sx. Then the triangle in-
equalities for f2 and each f;" imply that
= 1,
Notice that if a property P defined for normed spaces is such that when-
ever a Hormed space X has it then so do all normed spaces isometrically
isomorphic to X, then a normed space (X, 11llx) is isomorphic to some
normed space (Y, 11lly) with property P if and only if there is a norm 1111 p
for X equivalent to 11llx such that (X, II lip ) has property P. The forward
implication comes about by letting IIxlip = IITxlly for each x in X, where T
is the isomorphism, while the converse is obtained by noting that the iden-
tity map on X is an isomorphism from (X, II I x) onto (X, 1111 p). Thus,
5.2 Uniform Rotundity 453
the preceding corollary can be viewed as the statement that every normed
space with an equivalent uniformly rotund Banach norm is reflexive.
A normed space is called superrefiexive if it can be given a uniformly
rotund norm equivalent to its original norm (though that is not the original
definition of superreflexivity; see [56, pp. 168~ 173] for the original definition
and a proof that it is equivalent to the one given here). It follows from
Corollary 5.2.16 that every superreflexive Banach space is reflexive. Mahlon
Day showed in a 1941 paper [47] that there are reflexive spaces that are
not superreflexive.
In a way, the Milman-Pettis theorem explains why it should not be sur-
prising that the classical nonreflexive Banach spaces examined in the intro-
ductory examples of Section 5.1 are all nonrotund, and why it took so much
work to construct a nonreflexive rotund Banach space in Example 5.1.8.
A nonreflexive rotund Banach space X cannot be uniformly rotund, which
means that while the midpoint of every straight line segment connecting
distinct points on the unit sphere of X must dip into the interior of the
closed unit ball, there is some positive f such that one can always find two
points on the unit sphere at least distance E apart with the midpoint of the
line segment connecting the points taking as shallow a dip into Bx as one
might wish. Such a unit sphere is somewhat oddly shaped.
Incidentally, it follows from the Milman-Pettis theorem that the rotund
nonreflexive Banach space f l ,,. of Example 5.1.8 is another example of a
rotund Banach space that is not uniformly rotund.
By Corollary 5.1.19, every nonempty closed convex subset of a rotund
reflexive normed space is a Chebyshev set, so the following result is an
immediate consequence of the Milman-Pettis theorem.
Recall that a normed space has the Radon-Riesz property if, whenever
(x n ) is a sequence in the space and x an element of the space such that
Xn ~ x and Ilx n II ---> Ilxll, it follows that Xn ---> x. As was discussed in Sec-
tion 2.5, J. Radon and F. Riesz proved that the Lebesgue spaces Lp(O, L:, /1)
such that 1 < p < 00 have this property. It turns out that this is a direct
consequence of the uniform rotundity of those spaces.
5.2.18 Theorem. Every uniformly rotund normed space has the Radon-
Riesz property.
PROOF. Suppose that X is a uniformly rotund normed space and that
(x n ) is a sequence in X and x an element of X such that Xn ~ x and
Ilxnll ---> Ilxll. It is to be shown that Xn ---> x, so it may be assumed that
x -I- 0 and therefore, after discarding some initial terms of (xn) if necessary,
that each Xn is nonzero. Now IIxnll~lxn ~ Ilxll~lx, and it is enough to show
454 5. Rotundity and Smoothness
5.2.19 Corollary. (J. Radon, 1913 [192]; F. Riesz, 1928-1929 [196, 197]).
Suppose that J-L is a positive measure on a a-algebra E of subsets of a set n
and that 1 < p < 00. Then Lp(n, E, J-L) has the Radon-Riesz property.
5.2.23 Example. Let il,r be the rotund nonreflexive Banach space of Ex-
amples 5.1.8 and 5.1.22 and let Oll,r be its modulus of rotundity. It follows
readily from the rotundity of il,r that oll,r(2) = 1; see Exercise 5.14. It was
shown in Example 5.1.22 that if Y is a separable nonrotund Banach space,
then there is a closed subspace M of il,r such that il,r/.M is isometrically
isomorphic to Y, so in particular there is a closed subspace Mo of il,r such
that il,r/Mo is isometrically isomorphic to i~. Then the modulus of rotun-
dity 0ll,rIMo for this quotient space is the same as that for R~, so it follows
easily that Oll,rIMo(2) = O. Therefore Oll,r(2) > OI.I,rIMo(2), which shows
why the preceding lemma was stated and proved only proved for values of E
strictly less than 2.
Incidentally, this example relies on the fact that RI,r is not reflexive. See
Exercise 5.19 for the reason.
for each n, from which it follows that Ilxnll -> 1. Similarly, it may be
assumed that llYn II -> 1. Since
it also follows that II ~ (xn + Yn) II -> 1, which implies that IIxn - Yn I ---> 0
and therefore that lI(xn - Yn) + Mil -> O.
456 5. Rotundity and Smoothness
5.2.25 Theorem. Suppose that Xl' ... ' Xn are normed spaces. Then
Xl EEl ... EEl Xn is uniformly rotund if and only if each Xj is uniformly ro-
tund.
PROOF. Suppose first that Xl EEl EElXn is uniformly rotund. Since each Xj
is isometrically isomorphic to a subspace of Xl EEl ... EEl X n , it follows im-
mediately that each Xj is uniformly rotund.
Now suppose instead that each Xj is uniformly rotund. Let (Xl, ... , Xn)
and CYI, ... , Yn) be elements of SX,EB ... EBXn such that
Lct
2
?: E2 - ~ (11(Xl, .. ' ,xn)112 + II(Yl, ... ,Yn)11 2)
E2
2'
so O:=jEAllxj - YjI12)1/2 ?: t/2 1/ 2, from which it follows that
and so
Exercises
5.14 Prove that if X is a normed space and 6x is its modulus of rotundity,
then X is rotund if and only if 6x(2} = 1.
5.15 Find an explicit formula, in terms only of functions one might encounter
in a precalculus course, for the modulus of rotundity D2 of Euclidean 2-
space. (This formula is important because it provides a simple expression
for the least upper bound of the moduli of rotundity of all uniformly
rotund normed spaces of dimension at least two. It is a 1960 result of
G. Nordlander [176] that if X is a uniformly rotund normed space that
is not zero-dimensional if IF = IC and is neither zero- nor one-dimensional
if IF = 1ft, and if Dx is the modulus of rotundity of X, then DX(E) ::; b2(E}
when 0::; E ::; 2.)
458 5. Rotundity and Smoothness
5.16 This exercise assumes some knowledge of inner product spaces and Hilbert
spaces. See, for example, the references cited in Exercise 5.4.
(a) Suppose that X is an inner product space that is not zero-dimension-
al if IF = C and is neither zero- nor one-dimensional if IF = R Find
an explicit formula for the modulus of rotundity of X in terms only
of functions one might encounter in a precalculus course. (Compare
Exercise 5.15.)
(b) Use the formula found in (a) to show that every inner product space
is uniformly rotund.
(c) Conclude from (b) that every Hilbert space is reflexive.
5.17 The purpose of this exercise is to provide another proof of Theorem 5.2.11
when p 2: 2 that is perhaps a bit more in the spirit of Clarkson's. Let p
be such that 2 :.:; p < 00.
(a) Prove that if 0:, {3 E IF, then
and therefore
5.20 (M. M. Day, 1944 [50]). Prove that a normed space X is uniformly rotund
if and only if
5.23 Prove that the Banach space X of Example 5.2.13 is superreftexive. (Ex-
ercise 5.21 might he helpful.) Conclude that X is a reflexive Banach space
that is rotund but not uniformly rotund.
For example, the i and ~ norms are equivalent on JF2, so i is not (R) but
is (UR). As further examples, the statement that every uniformly rotund
Banach space is a rotund reflexive Radon-Riesz space can be abbreviated
to (UR) & (B) "*(R) & (Rf) & (H), while the fact that every superreflexive
Banach space is reflexive can be stated symbolically as (UR) & (B) (Rf). "*
The first generalization of uniform rotundity to be examined here is a
localization of that property obtained by requiring that for each fixed x
in the unit sphere of a normed space X and each positive E, there is a
positive 0 depending on E and x such that 11~(x + y)11 S 1 - 0 whenever
y E Sx and Ilx - yll 2': E.
for each n, it follows that IIYnl1 -> 1, so Ilx - Ynll -> 0 by (d). Therefore
(d) =? (c), from which it immediately follows that (b) <=;. (c) B (d).
Suppose that X is locally uniformly rotund and that x is an element
ofSx and (Yn) a sequence in Sx such that 11~(X+Yn)ll-> 1 but Ilx-Ynll
docs not tend to O. Let Ox be the LUR modulus of X. It follows that there
is a subsequence (Ynj) of (Yn) such that Ilx - Yn J I :::: E for some positive E
and each j, which implies that II!(x + YnJl1 ~ 1 - OX(E,X) for each j, a
contradiction. Therefore (a) =? (b).
462 5. Rotundity and Smoothness
5.3.6 Example. (M. A. Smith, 1978 [219]). This example uses the paral-
lelogram law for 2, namely, that if (an), (f3n) E 2, then
See Exercise 5.24. Also needed is the fact that if (an) E 1, then (an) E 2
and II(an)112 :s; II(an)111. The proof of this is easy: If (an) E Sfp then
lanl :s; 1 for each n, so 2:nlanl2 :s; 2:nlanl = 1 and (an) E B e2 .
For each member (an) of 1, let
o :s; 2(ll xlli + IIYklli) - Ilx + Yklli + 2(llxll~ + IIYkllD - Ilx + Ykll~
= 2(llxll~ + IIYkll~) - Ilx + Ykll~
= 4- Ilx + Ykll~
for each k, and since 4 - Ilx + Ykll~ ----+ 0, it follows that
5.3 Generalizations of Uniform Rotundity 463
for each k by the parallelogram law, it also follows that Ilx - Ykl12 --> 0, so
f3n,k --> an for each n.
It will now be shown that Ilx - Yk 111 --> O. If x = 0, then this fol-
lows immediately from the fact that IIYklll --> Ilxlll. Suppose instead that
x i= 0. Fix a positive f less than Ilxlh and let m be a positive integer
such that L:~=m+1Ianl < E. Notice that L:::'=llanl > Ilxlh - E. It follows
that L:::'=llf3n,kl > Ilxlll - E for large k, which together with the fact that
L: nlf3n,kl = IiYklil --> Ilxlh implies that L:~=Tn+11f3n,kl < 2E for large k.
Therefore for large k,
m 00 00
5.3.7 Theorem. (R. Vyborny, 1956 [240]). Every locally uniformly rotund
normed space has the Radon-Ricsz property. In symbols, (LUR) :::} (H).
464 5. Rotundity and Smoothness
One reason for the importance of local uniform rotundity is that it im-
parts some of the same benefits, such as the presence of the Radon-Riesz
property, as does uniform rotundity, while far more spaces can be equiv-
alently renormed to be (LUR) than to be (UR). It was shown in a 1971
paper by S. L. Troyanski [232] that every weakly compactly generated Ba-
nach space is (LUR), while no nonreflexive Banach space can be (UR).
The next generalization of uniform rotundity is obtained by letting the
weak topology play the rolc of the norm topology in the definition of uni-
form rotundity, in a sense that will be made clearer by Proposition 5.3.9.
5.3.8 Definition. (V. L. Smulian, 1939, 1940 [223, 224]). Suppose that
X is a normed space. Define a function Ox: [0,2] x Sx' ---f [0,1] by the
formula
The reason for specifically including 1 in the set whose infimum defines
the wUR modulus is to keep the modulus finite-valued in one particular
situation. With all notation as in Definition 5.3.8, suppose that some x*
in SX' is not norm-attaining. Then there are no members x and y of Sx
such that Ix*(x - y)1 ~ 2, so bx(2,x*) would be +00 were it not for the l.
As with uniform rotundity and local uniform rotundity, weak uniform
rotundity has sequential characterizations. The next result gives one along
the lines of the equivalence of (a) and (b) in Propositions 5.2.8 and 5.3.5.
The derivation of other characterizations analogous to parts (c) and (d) of
those two propositions is left as an exercise for the interested reader.
PROOF. Suppose that (b) fails. Then there are sequences (xn) and (Yn)
I
in Sx and an x* in Sx' such that ~ (xn Yn) + I -+ 1 but X* (Xn - Yn) does
not tend to O. Let Ox be the wUR modulus of X. It follows that there is a
subsequence (xnj ) of (xn) such that Ix' (xn) - Yn)) I 2: E for some positive E
and each j, implying that OX(E,X*)::; 1-11~(Xnj +YnJII for each j and
therefore that () x (E, x*) = O. Therefore X is not weakly uniformly rotund.
Suppose conversely that X is not weakly uniformly rotund. Then there is
an E such that 0 < E ::; 2 and an x' in Sx' for which OX(E) = O. Therefore
there are sequences (xn) and (Yn) in Sx such that Ix*(xn - Yn)1 2: E for
each n but 11~(Xn + Yn)ll-+ 1, so (b) does not hold.
It follows easily from the preceding proposition that weak uniform ro-
tundity lies between uniform rotundity and rotundity.
Mark Smith's paper [219] contains examples of Banach spaces (l2' II lid
and (l2' 11llw), each formed by putting a norm on l2 equivalent to its orig-
inal norm, such that (l2' II IlL) is (LUR) but not (wUR) while (l2' 11llw)
is (wUR) without being (LUR). This shows that neither of the conditions
(LUR) and (wUR) implies the other, and also shows that weak uniform ro-
tundity does not imply uniform rotundity. Smith also shows that (l2' 1Illw)
lacks the Radon-Riesz property, so (wUR) does not imply (H).
Incidentally, it turns out that the Banach space (i1' II liE) of Exam-
ple 5.3.6 is another Banach space that is (LUR) but not (wUR), since
V. E. Zizler showed in a 1971 paper [248J that II cannot be equivalently
renormed to be (wUR).
There is an obvious analog of weak uniform rotundity for dual spaces
that is obtained by exchanging the roles of the normed space and its dual
in the definition of the wUR modulus. This analog will be examined only
briefly here, but will have an important application in Section 5.6.
466 5. Rotundity and Smoothness
5.3.11 Definition. (V. L. Smulian, 1939, 1940 [223, 224]). Suppose that
X is a normed space. Define a function OX, : [0,2] x Sx ----> [0,1] by the
formula
Notice that unlike what was done for the wUR modulus, there is no 1
explicitly included in the set whose infimum is being taken to obtain the
w' UR modulus. This is not needed to assure that 8x- (2, x) is finite when
xESx, since for each x in S X there is an x* in S X' such that x* x = 1
and therefore such that I(x' - (-x')) (x) I = 2.
It should be noted that some sources say that if X is a normed space
such that X* satisfies the above definition of weak* uniform rotundity,
then it is X instead of X' that is called weakly* uniformly rotund. See
[107, p. 71] for an instance of the use of the term in this alternative sense,
and [239, p. 48J for an example of the use of the term the way it has been
defined here.
OX- (I', x) = inf {1 -II ~(x' + y*)11 : x', y' E Sx, I(Qx)(x* - Y*)I ~ f}
whenever 0 :S I' :S 2 and xESx. Comparing this to the formula for the
wUR modulus of X* shows that X* is weakly* uniformly rotund if it is
weakly uniformly rotund.
Now suppose that X* is not rotund. Let xi and x~ be different elements
of Sx- such that ~(xi + x 2) E Sx' and let Xo be an element of Sx such
that xixo =I- xzxo. Let 1'0 = I(xi -xz)(xo)l Then OX*(EO,XO) = 0, so X* is
not weakly* uniformly rotund.
formula
A glance at the definitions of local, weak, and weak local uniform rotun-
dity shows that each of the first two properties implies the third. Also, if X
is a normed space with wLUR modulus bx and there are distinct elements
x and y of X such that Ilxll = liyll = 11~(x+Y)11 = 1, then there is an x
in Sx' such that x*(x-y) ~ 0, which implies that bx (lx*(x-y)l, x,x*) = O.
It follows that weak local uniform rotundity implies rotundity. These ob-
servations are summarized in the following proposition.
Mark Smith's paper [219] has examples showing that none of the impli-
cations in the preceding proposition is reversible.
The next generalization of uniform rotundity is defined in terms of ge-
ometric properties of convex sets instead of the behavior of a modulus.
Recall that the diameter of a nonempty subset A of a metric space is given
by the formula diam(A) = sup{ d(x, y) : x, yEA }.
5.3.17 Theorem. (K. Fan and 1. Glicksberg, 1958 [77]). Suppose that X
is a normed space. Then the following are equivalent.
(a) The space X is strongly rotund.
(b) Whenever x* E Sx+, the diameter of {x: x E Ex, Rex*x::::> 1- 15}
tends to 0 as 6 decreases to O.
(c) Whenever (x n ) is a sequence in Sx for which there is an element x'
of Sx' such that Rex'x n --t 1, the sequence (xn) is Cauchy.
PROOF. Suppose first that (b) does not hold. This implies the existence of
an x* in Sx' and a positive E such that for each positive integer n there
are elements Xn and Yn of Ex for which Re x*xn 2: (1 + n- 1)-1, Re X*Yn ::::>
(1+n- 1)-1, and Ilxn-Ynll 2: E. Let A = {x: x E X,Rex*x::::> 1},
a nonempty convex subset of X. It is clear that d(O, A) ::::> 1. Since there
is a sequence (zn) of members of Ex such that Re x' Zn > 0 for each n
and Rex'zn --t 1, and since ((Rex*zn)-1Zn) is a sequence in A such that
II(Rex'zn)-lznll -> 1, it follows that d(O,A) = 1. Let Un = (1 + n- 1)xn
and Vn = (1 + n- 1 )Yn for each n. Then Un, Vn E An (1 + n- 1 )Ex and
Ilu n - vnll > E for each positive integer n, so diam(A n tEx) does not tend
to 0 as t decreases to d(O, A). Therefore X is not strongly rotund, which
shows that (a) =} (b).
It is clear that (b) =} (c). To see that (c) =} (a), suppose that X is not
strongly rotund. Then there is a nonempty convex subset C of X such that
d(O, C) = 1 and diam( C n tEx) does not tend to 0 as t decreases to 1.
Therefore there must be a positive I and a sequence (w n ) in C such that
Ilwnll -> 1 and Ilw2n-1 -w2nll 2: I for each positive integer n. By Eidelheit's
separation theorem, there is an x* in X* and a positive real number s such
that Rex*x ::::> s whenever x E C and Rex'x ::; s whenever x E Ex. It may
be assumed that s = 1. Then Ilx"11 = IIRex'll::; 1, and in faet Ilx'll = 1
since Rex*(lIwnll-1w n ) -> 1; for this, notice that
1::::> Rex*(llwnll-1wn) 2: Ilwnll- 1
5.3 Generalizations of Uniform Rotundity 469
5.3.20 Theorem. (K. Fan and 1. Glicksberg, 1958 [77]). Suppose that X
is a normed space. Then the following are equivalent.
(a) The space X is strongly rotund.
(b) Whenever C is a nonempty convex subset of X and (Yn) is a mini-
mizing sequence in C with respect to some x in X, the sequence (Yn)
is Cauchy.
PROOF. It is very easily seen that (b) is equivalent to the following state-
ment.
(b o) Whenever C is a non empty convex subset of X and (Yn) is a mini-
mizing sequence in C with respect to 0, the sequence (Yn) is Cauchy.
The equivalence of (a) and (b o) follows almost immediately from the defi-
nition of strong rotundity.
5.3.21 Theorem. (K. Fan and 1. Glicksberg, 1958 [77]). Suppose that X
is a normed space. Then the following are equivalent.
(a) The space X is a strongly rotund Banach space.
(b) Whenever C is a nonempty convex subset of X and (Yn) is a mini-
mizing sequence in C with respect to some x in X, the sequence (Yn)
converges.
(c) Whenever C is a nonempty closed convex subset of X and (Yn) is
a minimizing sequence in C with respect to some x in X, the se-
quence (Yn) converges to an element Y of C.
(d) Every nonempty closed convex subset of X is an approximativcly
compact Chebyshev set.
If (c) holds and C, x, (Yn), and yare as in the statement of (c), then Y is
the unique point of C closest to x.
PROOF. It follows immediately from Theorem 5.3.20 that (a) =? (b), and it
is obvious that (b) =? (c). Suppose next that (c) holds and that C, x, (Yn),
and yare as in the statement of (c). Then Ilx - yll = d(x, C). Furthermore,
if y' is any member of C such that Ilx - y'll = d(x, C), then (c) implies
that the sequence (y, y', y, y', . .. ) eonverges, which shows that Y = y'. This
establishes the last statement in the theorem. It is now clear that (c) =? (d)
5.3 Generalizations of Uniform Rotundity 471
show that (zn) converges, so, after perhaps discarding some terms from the
beginning of the sequence, it may be assumed that Re Z Zn > for each n.
Let A = {x : x EX, Re z* x = 1 }, a nonempty closed convex subset of X,
and let Wn = (Re z* zn)-l Zn for each n. It is clear that d(O, A) ~ 1, and
thus that d(O, A) = 1 since the sequence (w n ) lies in A and IlwnII - 1. Since
every subsequence of (w n ) is a minimizing sequence in A with respect to 0,
it follows that every subsequence of (w n ) has a subsequence that converges
to the unique point W of A closest to 0, so Wn - w. Therefore Zn - W,
which by Corollary 5.3.18 proves that X is a strongly rotund Banach space.
This shows that (d) '* (a).
The property given in part (c) of Theorem 5.3.17 might seem to have
some of the flavor of the Radon-Riesz property. In fact, it implies the
Radon-Riesz property.
5.3.22 Theorem. (K. Fan and I. Glicksberg, 1958 [77]). Every strongly ro-
tund normed space has the Radon-Riesz property. In symbols, (K) (H).'*
PROOF. Suppose that X is a strongly rotund normed space and that (x n ) is
a sequence in X and x an element of X such that Xn ~ x and IIxnll -llxll.
Since the goal is to prove that Xn - x, it may be assumed that x =I- o. It is
easy to see that it may then be assumed that x and each Xn lie in Sx. Let x*
be a member of Sx. such that x*x = Rexx = 1. Then Rexx n -> 1, so by
Theorem 5.3.17 the sequence (x n ) is Cauchy. Therefore (x n ) converges in
the completion of X to some y, which together with the fact that Xn ~ x
implies that Xn -> x.
5.3.23 Theorem. (K. Fan and I. Glicksberg, 1958 [77]). Every reflexive
rotund normed space having the Radon-Riesz property is strongly rotund.
In symbols, (Rf) & (R) & (H) '*
(K).
PROOF. Suppose that X is a normed space that is (Rf) & (R) & (H). Let C
be a nonempty closed convex subset of X. Then C is a Chebyshev set
by Corollary 5.1.19. By Theorem 5.3.21, it is enough to show that C is
approximatively compact. Let (Yn) be a minimizing sequence in C with
respect to some x in X and let y be the unique element of C closest to x.
It is enough to show that (Yn) has a subsequence converging to y. By the
reflexivity of X, there is a subsequence (YnJ of (Yn) that c:onverges weakly
to some element Y' of the weakly closed set C, and Ilx - Y'II = d(x, C) since
472 5. Rotundity and Smoothness
d(x,C) :::; Ilx - y/ll :::; liminfjllx - YnJ = d(x,C). Therefore y' = y. Since
x - Ynj ~ X - Y and Ilx - Ynj II -4 Ilx - YII, it follows that x - Yn] -4 X - Y
and therefore that Yn J -4 y.
It follows from what has already been proved that (Rf) & (R) & (H) =}
(K) & (B) =} (R) & (H). If it could be shown that every strongly rotund
Banach space is reflexive, then a nice characterization of the strongly rotund
Banach spaces would follow: They would be precisely the normed spaces
that are reflexive and rotund and have the Radon-Riesz property.
This characterization does in fact hold since every strongly rotund Ba-
nach space is reflexive, as was first shown by Fan and Glicksberg in their
1958 paper [77J. However, this does not seem to be all that easy to prove
from elementary principles. The known proofs (or at least the ones known
to this author at the time of this writing) all use some form of James's theo-
rem or, in one case, the Bishop-Phelps 8ubreflexivity theorem; see [168J and
Exercises 5.33 and 5.65. Both of these results appear in optional sections
of this book, and neither i8 particularly trivial. Though Fan and Glicks-
berg published their result several years before the general case of James's
theorem appeared in James's 1964 paper [112], they did use a weak form
of James's theorem from his 1957 paper [111J.
Though reflexive locally uniformly rotund normed spaces are strongly
rotund, it follows from the remarks of the preceding paragraph that nonre-
flexive locally uniformly rotund Banach spaces, such as the one of Exam-
ple 5.3.6, cannot be strongly rotund, so (LUR) does not imply (K). Mark
Smith's paper [219J contains additional examples to show that (K) neither
implies nor is implied by any of the properties (LUR), (wUR), and (wLUR).
For more on strong rotundity, see [224J and [226], and in particular Fan
and Glicksberg'8 paper [77J in which a number of reformulations of the
property are given in addition to the ones obtained above.
The last generalization of uniform rotundity to be studied in the body of
this section is obtained by starting with the first sequential characterization
of uniform rotundity in Proposition 5.2.8 and then stiffening the hypotheses
on the sequences (x n ) and (Yn) by requiring not just that 11~(Xn + Yn)11
tend to 1, but that there actually be an element of Sx to which ~(xn +Yn)
converges.
Notice that if (Xn) and (Yn) are sequences in the unit sphere of a normed
space X that is (MLUR), and if ~(xn + Yn) converges to some z in Sx,
then both Xn and Yn also converge to z. In fact, a moment's thought shows
that an equivalent definition of midpoint local uniform rotundity would
have been obtained if in Definition 5.3.25 it were required that Xn and Yn,
or for that matter just Xn, tend to the same limit as ~(xn + Yn), rather
than requiring that II Xn - Yn I --> O. There is a large number of such ways
to make small modifications to the conclusion about (xn) and (Yn) in Def-
inition 5.3.25 without changing the property being defined. The following
result gives one simple one that will be useful in what is to follow.
the construction in the obvious way of trivial sequences (xn) and (Yn) in Sx
such that ~ (xn + Yn) converges to the element ~ (x + y) of Sx even though
IIx n ~Ynll does not tend to O. Therefore X is not midpoint locally uniformly
rotund.
Now suppose instead that X is strongly rotund and that (xn) and (Yn)
are sequences in Sx such that ~(xn + Yn) converges to some member Z
of Sx. Then there is some z* in Sx* such that z* Z = Re z* Z = 1. Since
Rez*x n ::; 1 and Rez*Yn ::; 1 for each n and ~(Rez*xn + Rez*Yn) -> 1,
it follows that Re z*x n -> 1 and Re z*Yn -> 1. Let (zn) be the sequence
(Xl, YI, X2, Y2' .. ')' Then Re z* Zn -> 1, so (z,,) is Cauchy by Theorem 5.3.17.
Therefore Ilx n - Ynll -> 0, which shows that X is (MLUR).
Finally, suppose that X is locally uniformly rotund instead of strongly
rotund, and again suppose that (xn) and (Yn) are sequences in Sx such
that ~ (xn + Yn) converges to some Z in S x. Since
as n --+ 00, it follows that 11~(Xn + z)ll--+ 1. Therefore Ilx" - zil --+ 0 by
Proposition 5.3.5. Similarly, it follows that llYn - zll --+ 0, so Ilx n- Yn I -> O.
This establishes that X is (MLUR).
Since
for each n and since Ilu n + Vn - 2wll ---> 0, each subsequence of (un) is a
minimizing sequence in Bx with respect to 2w and therefore must have
a convergent subsequence whose limit is necessarily the unique point w
of Bx closest to 2w. It follows that Un -+ w, and similarly that Vn -+ w,
and therefore that Ilu n - vnll -+ 0. This proves that X is (MLUR).
476 5. Rotundity and Smoothness
(K)
l~
(H) (MLUR)
( U R ) - - - - - . (LUR)
i/ ~
(R)
~
(wLUR)
(wUR)
/
FIGURE 5.3. An implication diagram for generalizations of uniform rotundity.
Figure 5.3 shows all of the implications between single properties proved
in this section, except for those involving weak* uniform rotundity that
apply only to dual spaces; however, see also Theorem 5.3.23 and Corol-
lary 5.3.24. Notice that (H) does not imply any of the other properties in
the diagram, since i
is not rotund but, like all finite-dimensional normed
space8, has the Radon-Riesz property. It follows from this and Smith's
examples already mentioned above that no arrows can be added to this
diagram showing relationships not already implied by the diagram.
Extensive lists of other generalizations of uniform rotundity as well as
some of their characteristics and the relationships between them can be
found in [56, pp. 145-147] and [107, pp. 71-93]. A few of these general-
izations can be found in the exercises for this scction. Two of the general-
izations from the exercises, namely, uniform rotundity in weakly compact
sets of directions and uniform rotundity in every direction, as well as weak
uniform rotundity and uniform rotundity itself, are directionalizations of
uniform rotundity in the sense that they can be defined in terms of the
following modulus. See Exercises 5.29-5.32.
Exercises
5.24 Prove the parallelogram law for 2 mentioned in Example 5.3.6.
5.25 Give at least one sequential characterization of weak local uniform rotun-
dity.
5.26 Find a sequential characterization of weak* uniform rotundity.
5.27 (a) Prove that every approximatively compact set is closed.
(b) Give an example of an approximatively compact set that is not com-
pact, or even weakly compact.
5.28 (K. Fan and 1. Glicksberg, 1958 [77)). Suppose that k is a positive integer
such that k :2: 2. A normed space X is k-rotund if each sequence (Xn) in X
such that limnl, .. ,nk~(X)lIk-1 2:;=1 Xnj II = 1 is Cauchy. (The notation
limnl, ... ,nk~(X)llk-l 2:;=1 Xnj II = 1 means that for every positive t there
is a positive integer N< such that Illk- ' 2:;=1 Xnj II - 11 < E whenever
n" ... , nk :2: N,.) The abbreviation kR is used for the property of k-
rotundity.
(a) Prove that every uniformly rotund normed space is k-rotund, and
that every k-rotund normed space is (k + I)-rotund. That is, show
that (UR) =} (kR) =} ((k + I)R).
(b) Prove that every k-rotund normed space is strongly rotund. That
is, show that (kR) =} (K).
5.29 Prove that a normed space X is (UR) if and only if 6X(E, ~A) > 0
whenever 0 < E ~ 2 and A is a nonempty bounded closed subset of X\ {OJ.
5.30 Prove that a normed space X is (wUR) if and only if 6X(E, --+A) > 0
whenever 0 < E ~ 2 and A is a nonempty bounded weakly closed subset
of X \ {OJ.
5.31 (M. A. Smith, 1975, 1977 [217, 218]). A normed space X is uniformly
rotund in weakly compact sets of directions if 6x (E, --+A) > 0 whenever
o < E ~ 2 and A is a nonempty weakly compact subset of X \ {OJ. The
abbreviation URWC is used for this property.
(a) Prove that a normed space X is (URWC) if and only if it has this
property: Whenever (Xn) and (Yn) are sequences in Sx such that
II ~(Xn + Yn)11 ~ 1 and Xn - Yn ~ v for some v in X, it follows that
v =0.
(b) Show that (wUR) =} (URWC) =} (R).
478 5. Rotundity and Smoothness
(c) Show that (Rf) & (wUR) {=> (Rf) & (URWC); that is, that weak
uniform rotundity and uniform rotundity in weakly compact sets of
directions are equivalent properties for reflexive normed spaces.
Smith has shown in [219J that the nonreflexive Banach space (1, II liE)
of Example 5.3.6 is (URWC) but not (wUR).
5.32 (A. L. Garkavi, 1962 [84]). A normed space X is uniformly rotund in
every direction or uniformly convex in every direction or directionally
uniformly rotund if OX(E, ->z) > 0 whenever 0 < E ::; 2 and z E Sx. The
abbreviation URED is used for this property.
(a) Prove that a normed space X is (URED) if and only if it has this
property: Whenever (Xn) and (Yn) are sequences in Sx such that
11~(Xn + Yn)11 -> 1 and such that Xn - Yn E ({v}) for some v in X
and each n, it follows that Xn - Yn -> O.
(b) Show that (wUR) =} (URED) =} (R).
(c) (This uses material from Exercise 5.31). Show that (URWC) =}
(URED).
Smith gave an example in [219] of a Banach space that is (URED) but not
(URWC) , and another Banach space that is (R) but not (URED); both
examples are formed by equivalently renorming 2. It can be shown that
a normed space X is (URED) if and only if OX(E, ->A) > 0 whenever 0 <
E ::; 2 and A is a nonempty compact subset of X \ {O}; see [218]. Another
good source of information on uniform rotundity in every direction is the
paper of Day, James, and Swaminathan [57] devoted to the property.
5.33 This exercise requires James's theorem from either of the optional Sections
1.13 and 2.9. Prove that a normed space is a strongly rotund Banach space
if and only if it is a reflexive rotund normed space with the Radon-Riesz
property. That is, show that (K) & (B) {=> (R) & (Rf) & (H).
5.34 (Ivan Singer, 1964 [215]). A normed space X has the Efimov-Stechkin
property if, whenever (xn) is a sequence in Sx for which there is an
element x' of Sx* such that Rex'x n -> 1, the sequence (Xn) has a
convergent subsequence. The abbreviation CD is used for this property.
(a) Show that a normed space X has the Efimov-Stechkin property if
and only if every nonempty closed convex subset of X is approxi-
matively compact.
(b) Prove that every strongly rotund Banach space has the Efimov-
Stechkin property.
(c) Show by example that not every Banach space with the Efimov-
Stechkin property is rotund.
(d) Prove that every normed space with the Efimov-Stechkin property
is a Banach space.
(e) Prove that every normed space with the Efimov-Stechkin property
has the Radon-Riesz property.
(f) The rest of this exercise requires James's theorem from either of the
optional Sections 1.13 and 2.9. Prove that every normed space with
the Efimov-Stechkin property is reflexive.
5.4 Smoothness 479
(g) Prove that a normed space has the Efimov-Stechkin property if and
only if it is reflexive and has the Radon-Riesz property. That is,
show that (CD) :} (Rf) & (H).
(h) Conclude from the results of this exercise and Exercise 5.33 that a
normed space is a strongly rotund Banach space if and only if it is
rotund and has the Efimov-Stechkin property. That is, show that
(K) & (B) :} (R) & (CD).
5.35 The purpose of this exercise is to present a technique that can be used
to design a real normed space having ]R2 as its underlying vector space
such that the closed unit ball of the space has certain specified geometric
properties.
(a) Prove that if a subset C of ]R2 is balanced (which in this case is
equivalent to being symmetric about the origin), convex, and closed,
and has nonempty interior with respect to the Euclidean topology
for IR2, then C is the closed unit ball for some norm on ]R2 that is
equivalent to the Euclidean norm of ]R2.
(b) Use (a) to construct a Banach space X with the property that for
some clement x of S x there can be found a different element y of S x
such that ~ (x + y) E Sx, but for some other element x' of Sx there
is no y' in Sx such that x' oj y' and ~(x' + y') E ."ix .
5.36 (R. C. James, 1964 [113]). Suppose that X is a normed space and that
fix is the modulus of rotundity of X. Then X is in quadrate or uniformly
nonsquare if there existll an E such that 0 < E < 2 and fiX(E) > O. The
abbreviation NQ is used for this property. It is obvious that this property
is implied by uniform rotundity. Show that it does not imply rotundity.
The method of Exercise 5.35 might be helpful.
James has shown in [113] that every inquadrate Banach space is reflexive.
Notice that it follows from this and Exercise 5.14 that if fiy is the modulus
of rotundity for a rotund nonreflexive Banach space Y, then
if 0 < E < 2;
fiY(E)={~ if E = 2.
It even turns out that a normed space is (NQ) if and only if it is (UR),
that is, if and only if it is superreflexive. See [56, pp. 169-173] for a proof
of this and for references concerning the history of the result.
5.4 Smoothness
In the introduction to this chapter, it was said that a normed space's closed
unit ball is smooth if the unit sphere has no "corners" or "sharp bends." It
is time to make this "definition" a bit more rigorous. One clue as to how
this could be done is given by the illustration of the unit sphere of real Pi
in Figure 5.4. Through each of the four corners of this unit sphere, it is
480 5. Rotundity and Smoothness
possible to pass more than one line that does not penetrate the interior of
the closed unit ball, while this is not possible at any of the other points
of S,i' Since the support hyperplanes for B~ are precisely the straight lines
in fi that intersect S21 but not B~2'
<1
this suggests the following definition.
5.4.8 Example. The scalar field IF, viewed a..') a normed space over IF,
is obviously smooth. The same holds for every zero- or one-dimensional
normed space.
if k j E A;
if kj 1:. A.
5.4 Smoothness 483
[fd81 = [fd82 = 1,
3The reason for not substituting 1 for IIxo II in the following quotient or in similar
quotients in the rest of this chapter is to make the point that these expressions represent
difference quotients for the norm function.
484 5, Rotundity and Smoothness
'
I lIIl Ilxo + tY11 -llxull
"'-----"--'-"--'-'---'"
t->O~ t
and
Ilxo + tY11 - Ilxoll
' ---'------'---'-'-
I1m
t->o+ t
exist, and
'
I1m Ilxo + tY11 - Ilxo I < I'1m Ilxo + tY11 - Ilxo I ,
t->U- t - t--->O+ t
Furthermore, the function
'
Y f-t l 1m
Ilxo + tYII- Ilxoll
t->O+ t
is a sublinear functional on X,
PROOF, For the moment, consider the element Y of X to be fixed, and let
and
Ilxo - t2 yll - Ilxoll Ilxo + t2( -Y)II - Ilxull
-tz t2
<
Ilxo + tJ (-Y)II - Ilxoll
tl
Ilxo - tlyll - Ilxoll
-tl
Since Ilxoll ::; Hllxo - tlyll + Ilxo + tlyll), it is also true that
Ilxo - t1yll- Ilxoll Ilxo + t1YII-llxoll ,
~--~~~~<
-tl - tl
5.4 Smoothness 485
. IIxo
I1m + tSYlll - Ilxoll = S I'1m Ilxo + tSYl11 -- Ilxoll = sg ()
9 ( SYI ) = Yl .
t-+o+ t t->o+ ts
The function 9 is therefore positive-homogeneous. Also, if t > 0 then
Ilxo + t(Yl + Y2)11 -lixoll Ilxo + 2tyIiI- Ilxoll Ilxo + 2tY211 - Ilxo ll
t :S 2t + 2t '
so letting t tend to 0 through positive values shows that g(Yl + Y2) <
g(Yd + g(Y2)' Therefore 9 is finitely subadditive and so is sublinear.
Then G_(XO,Yo) and G+(xo, Yo) are, respectively, the left-hand and right-
hand Gateaux derivative oj the norm at Xo in the direction Yo. The norm is
Gateaux differentiable at Xo in the direction Yo if G - (xo, Yo) = G + (xo, Yo),
in which case the common value of G-(xo,Yo) and G+(xo, Yo) is denoted
by G(xo, Yo) and is called the Gateaux derivative oj the norm at Xo in the
direction Yo. If the norm is Gateaux differentiable at :Z:o in every direc-
tion y, then the norm is Gateaux differentiable at Xo. Finally, if the norm
is Gateaux differentiable at every point of the unit sphere, then it is simply
said that the norm is Gateaux differentiable.
t->O t
exists, in which case G(xo, Yo) is the limit. Therefore the norm of the space
is Gateaux differentiable if and only if the above limit exists for each Xo
486 5. Rotundity and Smoothness
in the unit sphere and each Yo in the space. Notice also that if Xo E S x,
then G~(xo, y) = -G+(xo, -y) whenever y E X, which together with the
sub linearity of G+(xo,) implies that G~ (xo,) is positive-homogeneous.
whenever y EX.
PROOF. Suppose first that Xo supports Ex at Xo and that y EX. If t > 0,
then
and so
fact that G+(xo,ryo) = -G_(xo,-ryo) for each real r and in particular for
each negative r. By the vector space version of the Hahn- Banach extension
theorem, there is a linear functional x;on X such that the restriction
of Re x; to V is Is and Rex;y S; G+(xo, y) whenever y E X. Notice that
this implies that
1= xixl + X~X2
:S Ilxillllxlll + Ilx~llllx211
:S (11xi1l2 + Ilx~1I2)l/2(llxlI12 + Ilx2112)1/2
= Ilx*IIII(Xl,X2)11
= 1,
and, since there must actually be equality throughout, that (1lxi II, IIX21i)
is a real multiple of (1Ixrli, Ilx211) Therefore Ilxill = Ilxlll and IIx211 =
Ilx211. It also follows that xixl = IIxillllxtll = IIxll12 and x2x2 = Ilx2112.
The same argument applied to y* shows that Ilyi II = IlxrIl, Ily::i11 = Ilx211,
yiXl = IIxIil 2 , and yiX2 = Ilx2112. However, it is an easy consequence of
the smoothness of Xl that there is a unique member z* of Xi such that
Ilz*11 = Ilxlll and Z'XI = Ilxll12. Therefore xi = yi, and similarly xi = yi,
so x* = y*. By Corollary 5.4.3, the normed space Xl EEl X z is smooth.
The convention will be adopted that with all notation as in the preceding
definition, when the map 1/ is everywhere singleton-valued it will be treated
490 5. Rotundity and Smoothness
in this text as if its values were members of Sx' rather than singleton sub-
sets of Sx" The following result, which is essentially just a restatement of
Corollary 5.4.3, says that this convention will be invoked precisely when X
is smooth.
5.4.26 Corollary. A normed space is rotund and smooth if and only if the
spherical image map for its unit sphere is singleton-valued and one-ta-one.
Spherical image maps for unit spheres of normed spaces are always
weakly* -compact-valued.
When a normed space is smooth and the spherical image map for its unit
sphere is treated as point-valued rather than set-valued, it is natural to ask
if the map has any continuity properties. More generally, one might ask
if the set-valued spherical image map for the unit sphere of an arbitrary
normed space has any properties analogous to continuity. In fact, it does.
5.4.28 Theorem. Suppose that v is the spherical image map for the unit
sphere of a normed space X and that G is a weakly* open subset of X*.
Then {x: x E Sx, v(x) ~ G} is an open subset of Sx.
PROOF. Suppose to the contrary that {x : X E Sx, v(x) ~ G} is not open
in Sx. Then there is an Xo in Sx and a sequence (xn) in Sx converging
to Xu such that v(xu) ~ G but v(x n ) %G whenever n E N. For each n,
let x~ be a member of v(xn) \ G. By the weak* compactness of Ex., there
5.4 Smoothness 491
is some sub net (x~) of (x~) that is weakly* convergent to some Xo in Ex-.
For each 0:,
so passing to the limit shows that xi)xo = 1 and therefore that x(j E Sx-.
It follows that x(j E v(xo), so xi) is in the weakly* open set G even though
the net (x~) that is weakly* convergent to xi) lies entirely outside G. This
contradiction finishes the proof.
5.4.29 Corollary. The spherical image map for the unit sphere of a smooth
normed space is norm-to-weak* continuous.
It follows from this corollary that the spherical image map for the unit
sphere of a smooth normed space is norm-to-weak continuous if the space
is reflexive, and is even norm-to-norm continuous if the space is finite-
dimensional. These stronger continuity conditions can be used to define
and characterize strengthened smoothness conditions for normed spaces,
as will be seen in Section 5.6.
Exercises
5.37 (a) Give an example of a rotund Banach space that is not smooth.
Exercise 5.35 might help.
(b) Give an example of a smooth Banach space that is not rotund.
5.38 This exercise is based on the results of Exercise 5.1. Suppose that (Xn)
is a sequence of normed spaces.
(a) Let t'oo((Xn)) be the collection of all sequences (Xn) such that
Xn E Xn for each nand sUPnllXnll is finite. Let II(xn)lloo = sUPnllxnl1
for each member (Xn) of t'oo((Xn)). Show that 00 ((Xn)) is a vec-
tor space when given the obvious vector space operations and that
1111= is a norm on this vector space. Show also that = ((Xn)) is a
Banach space if and only if each Xn is a Banach space.
(b) Suppose that 1 S p < 00 and that q is conjugate to p. For each
member (x;,) of q((X;;)), let
5.41 Let l.T be the nonreflexive rotund Banach space constructed 1Il Exam-
ple fl. 1.8 and studied further in Example 5.1.22. Usc Exercise 5.40 t.o prove
that i,T is not smooth.
5.42 Suppose that K is a compact Hausdorff space with more than OIle element.
Show that rca(K) is not smooth,
5.43 This exercise assumes some knowledge of inner product spaceb and their
duals; see, for example, [24]. Prove that every inner product space is
smooth. Exercise 5.4 might help.
5.44 (V. L. Smulian, 1939 [222]). Suppose that X is a normed space and that
Xo E Sx" Show that the following are equivalent.
(a) The dCIIlfmt. Xo is a point. of smoothness of Bx.
(bl Whenever (Xn) is a sequence in Sx such that x~Xn -> 1, the se-
quence (Xn) is weakly Cauchy.
5.45 This exercise depends on Exercise 5.44 and on James's theorem from ei-
ther of the optional Sections 1.13 and 2.9. Show that a weakly sequentially
complete normed "pace is reflexive if its dual space is smooth. Conclude
that. there is no norm II lis on 1 equivalent to its standard norm such
that (1, IIII~) * is smooth.
5.46 (D. F. Cu<iia, 1964 [44]). Suppose that X is a normed space and that Q is
the natural map from X into X'*. Then X' is Q(X)-rotund if, whenever
Xi,X2 E Sx' and there is an x" in SQ(X) such that x*'xr = x**x.~ = 1,
it follow" that xr = X2' Also, the space X' is Q(X)-srnooth if, whenever
xi', x~' E SQ(X) and there is an x* in Sx' such that xi'x' = x;*x' = 1,
it follows that xi' = x~*.
(al Prove that X is rotund if and only if X* is Q(X)-smooth.
(b) Prove that X is smooth if and only if X* is Q(X)-rotund.
5.47 Characterize the members (an) of Seo that are points of smoothness of Bco
by means of some property of the terms of (an). Do the same for the
members of Sf l that are points of smoothness of Bt"
5.48 Prove t.hat if X is a Banach space that is not zero- or one-dimensional,
then there is a nonsmooth norm on X equivalent to its original norm.
5.5 Uniform Smoothness 493
5.50 (a) This exercise requires James's theorem from either of the optional
Sections 1.13 and 2.9. Find a characterization of the reflexive Banach
spaces among all Banach spaces in terms of the behavior of the
spherical image map for the unit sphere.
(b) Find a characterization of the reflexive, rotund, smooth Banach
spaces among all Banach spaces in terms of the behavior of the
spherical image map for the unit sphere.
5.51 Suppose that X is a normed space, that Q is the natural map from X
into X**, and that v and v* are the spherical image maps for Sx and Sx*
respectively.
(a) Prove that X is rotund if and only ifQ-1 (v*(v(x))) = {x} whenever
x E Sx. (If A is a set and <1> is a set-valued map with domain A,
then <1>(S) = U{ <1>(8) : 8 E S} whenever S ~ A.)
(b) Prove that X is smooth if and only if v( Q-1 (v*(x*))) = {x*} when-
ever x' E Sx*.
(b) hm
. Ilx + tY11 - Ilxll .
eXlsts whenever x E Sx and Y E: X.
t~O t
(c) lim
.! (11x
2
+ tY11 + Ilx - tyll)
.
- 1
= 0 whenever x E Sx and Y E X.
t~O+ t
Now suppose that X is smooth. Then the expression whose limit is being
taken in (b) converges to its limit uniformly for x in Sx when Y is a fixed
element of S x, or for y in S x when x is a fixed element of S x, or for (x, y)
in S x x S x, if and only if the expression whose limit is being taken in (c)
converges to its limit uniformly under the same condition on x and y.
PROOF. The equivalence of (a) and (b) is just a restatement of the fact
that X is smooth if and only if its norm is Gateaux differentiable. Observe
494 5. Rotundity and Smoothness
Illx + tY~1 -llxll _ G(x, y)1 + Illx - t~~ -llxll _ G(x, y)1
(a) For every positive f there is a positive Of such that (5.1) holds when-
ever 0 < t < Of and x, yESx.
(b) For every positive f there is a positive 0< such that Px(t)/t < f
whenever 0 < t < 8f
Consequently, the formal definition of uniform smoothness about to be
given is equivalent to the one stated above in terms of uniform convergence.
Px(t) = {ot- 1
ifO<t<1;
if t ;::: 1
Here are a few simple properties of the modulus of smoothness that will
shed some light on portions of the preceding definition.
~ (ilx + tlyll + Ilx - tlyll) - 1 = ~ (IIX + hyll - Ilxll _ Ilx - tlyll - Ilxll)
tl 2 t1 -tl
< ~ (IIX + t2yll - Ilxll _ Ilx - t2yll - Ilxll )
- 2 t2 -t2
~ (11x + t2yll + Ilx - t2yll) - 1
t2
from which it follows that PX(tl)/t l :S PX(t2)/t2.
Now fix a positive t. Since 0 :S Px (t) :S t, all that remains to be proved
is that p x (t) ;::: t - 1. However, this is an immediate consequence of the
fact that if x, y E Sx, then
As one might suspect, the preceding result does not generalize to isomor-
phisms. For example, the space i is not uniformly smooth since it is not
even smooth, but it is isomorphic to ~ which will be shown to be uniformly
smooth in Corollary 5.5.17.
Uniform smoothness implies stronger continuity properties for the spher-
ical image map than does smoothness. In fact, uniformly smooth normed
spaces can be characterized among all smooth normed spaces in terms of
the continuity of this map, as is apparent from the following lemma. This
lemma will have a further application in Section 5.6.
and therefore
\llx + tY~I- Ilxll _ G(x, y)\ = \IIX + tY~1 - Ilxll _ Re(v(x))(y)\ < E.
It was seen in Section 5.4 that if the dual space of a normed space X is
either rotund or smooth, then X has the other property, but the presence of
one of the properties in X does not imply that of the other in X . It turns
out that there is a more nearly complete duality between uniform rotundity
and uniform smoothness, since a normed space has either property if and
only if its dual space has the other. Early proofs of this or of results that
readily imply this can be found in [50] and [224]. The proof to be given
here is from [157] and is based on an interesting relationship between the
moduli of rotundity and smoothness of a normed space and of its dual.
and
when t > O.
PROOF. It is easy to check that the formulas work if X = {O}, so it will be
assumed that X =1= {O}. For each positive t,
OS:;fS:;2}
= sup{ tf - 20X(f) : 0 s:; s:; 2}.
Dividing through by 2 gives the first ofthe two claimed formulas. The other
is proved exactly the same way after exchanging the roles of X and X.
500 5. Rotundity and Smoothness
E 8X (E) E Es
----<-<-<8
2 ts - 2 - 2 - ,
while if 1' .. < to ::; 2, then the fact that 8x is nondecreasing assures that
(h(E) 8 (E .. )
- - - - < 1 - -X - - = 0 < 8.
2 ts - ts
It follows that px(t .. )/t s ::; 8, so X* is uniformly smooth. An analogous
argument with the roles of X and X* exchanged proves that X is uniformly
smooth when X* is uniformly rotund.
Now suppose that X* is uniformly smooth and that 0 < I' ::; 2. Let the
positive number t, be such that PX' (t,)/t, ::; 1'/4. It follows that
It follows from the preceding result and the Milman-Pettis theorem that
if X is a uniformly smooth normed space, then X* is uniformly rotund and
therefore reflexive, which in turn implies the reflexivity of X provided that
X is complete.
5.5.15 Example. The Banach space (co, II lib) of Example 5.4.13 is smooth
but not reflexive, and therefore is an example of a smooth Banach space
that is not uniformly smooth.
The main examples of uniformly smooth Banach spaces arise from the
following theorem.
5.5.17 Corollary. Suppose that 1 < p < 00. Then f.p is uniformly smooth,
as is f.; whenever n is a nonnegative integer.
= px(t),
as claimed.
5.5.22 Theorem. If M is a closed subspace of a uniformly smooth normed
space X, then X/M is uniformly smooth.
It is apparent from the results of this section that there are many analo-
gies between the behavior of uniformly smooth normed spaces and that
of uniformly rotund Hormed spaces. In fact, there is a much stronger con-
nection between the two properties than their duality and these analogies
would suggest. It turns out that a normed space has a uniformly smooth
norm equivalent to its original norm if and only if it has a uniformly ro-
tund norm equivalent to its original norm; that is, that (US) } (UR); and
therefore that the (US) norrned spaces are exactly the snperrefiexive ones.
A proof and references can be found in [56, pp. 169-173].
5.5 Uniform Smoothness 503
Exercises
whenever tl, t2 E I and 0 < s < 1. Prove that for every normed space X,
the modulus of smoothness of X is a convex function.
5.53 Prove that for every normed space X, the modulus of smoothness of X
is a continuous function. (It is not at all necessary to base the argument
on the result of Exercise 5.52, but feel free to use that result if you have
proved it.)
5.54 The modulus of smoothness of a normed space X is sometimes defined
by the formula
Prove that X is uniformly smooth if and only if f(t)/t ----+ 0 as t -> 0+.
5.57 (M. M. Day, 1944 [50]). Prove that a normed space X is uniformly rotund
if and only if
5.59 Find an explicit formula, in terms only of functions one might encounter in
a precalculus course, for the modulus of smoothness of Euclidean 2-space.
The result of Exercise 5.15 could help, but it is not crucial.
5.60 This exercise assumes some knowledge of inner product spaces and their
duals; see, for example, [24]. Prove that every inner product space is
uniformly smooth. Exercise 5.16 might help.
5.61 Prove that, for every normed space, the modulus of smoothness of the
space is the same as the modulus of smoothness of its second dual.
some rotundity condition for X lying between simple rotundity and uniform
rotundity. In fact, it turns out that Frechet smoothness of X* is equivalent
to X being strongly rotund, a fact due to SmuIian. The proof to be given
here will closely follow SmuIian's, and is based on four lemmas concerning
Cauchy and weakly* Cauchy norming sequences.
::::: 1- 21tl,
so
O:S; 1 -limRey~(x; y, t) :s; 21tl,
n
= t lim Re z~y.
n
506 5. Rotundity and Smoothness
Also,
:::; tlimRey~(y;y,t).
n
Therefore
Taking absolute values and dividing through by It I proves that (2) holds .
5.6.6 Lemma. (V. L. Smulian, 1939 [221]). Suppose that X is a normed
space, that Z is a norming subspace for X in X*, and that a sequence
in Sz is Cauchy if there is an element of Sx for which the sequence is a
norming sequence. Then X is Frechet smooth.
PROOF. Throughout this proof, references to (1), (2), and (3) are to the
conditions in the conclusion of Lemma 5.6.5. Fix an element x of Sx and a
norming sequence (z~) for x in Sz. For each y in Sx and each nonzero real
number t such that It I :::; 1/4, let (y~(.; y, t)) be a sequence in Sz satisfying
(1), (2), and (3). A moment's thought about (2) and the definition of a
FrE3chet differentiable norm shows that it suffices to prove that
(5.2)
and
for each m. Then Rey;~ (x; y"", t'Tn) -+ 1 by (3) and the fact that (5.3) holds
for each m, so (y;~ (.; Y'Tn' t'Tn)) is a norming sequence for x in Sz. Since
(5.2) and (5.4) hold for each m, it also foHows that
Ili;nRez~y",,-ReY;~(Ym;Y'Tn,t'Tn)1 > ~
5.6 Generalizations of Uniform Smoothness 507
for all sufficiently large m, which together with the fact that (z~) is Cauchy
implies that
(Rez*y)(Rez*x) -limRez~y
+Ilx+t[y- (limn~ez~y)xlil-Ilxli
n
:s (l-IRez*xl) (~+2)
signs = {+1 -1
if s 2: 0;
if s < O.
(Rez'y)(Rez'x) -limRez~y
n
= (Rcz*y)(signRez*x) -IRez'xllimRez~y
n
+ (Rez*y)(Rez'x - sign Rez*x) + (IRez*xl- 1) limRez~y
n
as claimed.
508 5. Rotundity and Smoothness
. Ilx
I1m + tYII - Ilxll
= Rex*y
t~O t
whenever y EX. Therefore
. Ilx
11m + try - (Rex*y)xlll - Ilxll
=0
t~O t
whenever yESx. Fix a positive f and let bE be a positive real number such
that
IIx+t[y- (Rex*y)xlil-Ilxli < -E
~--~--~----~~~~
t - 4
whenever 0 < t :::; b, and y E Sx; the Fnchet differentiability of the norm
of X guarantees the existence of such a bE. It follows from Lemma 5.6.7
that
The example (eo, IIIIF) given above can also be used to show that Frechet
smoothness is different from uniform smoothness, since the Banach space
(co, II I F) is Frechet smooth but cannot be uniformly smooth because it is
not reflexive. In fact, the three properties of uniform smoothness, Frechet
smoothness, and smoothness are distinct even in the presence of reflexivity,
as can be shown by applying Theorem 5.6.9 to some examples by Mark
Smith from [219]. Among Smith's examples are two reflexive normed spaces
(2, II I d and (2, II II w), each formed by giving 2 a norm equivalent to
its original norm, such that (2, II IlL) is (K) but not (UR) and (2, 11llw)
is (R) but not (K). It follows that (2, IIIIL)* is (F) but not (US), while
(2, IHw)* is (S) but not (F).
It is time to move on to the other major generalization of uniform
smoothness to be examined in this section.
t> 0,
(Rez*y)(Rez*x) - Rex*y
By Theorem 5.4.17,
IIx
11m + tyll - IIxll
= R ex *y,
t-->O t
It is worth pointing out again that (jE,y does not depend on the member x
of Sx or the member z* of Sx" Therefore
I(Rez;y)(Rez]"x) - (Rez;y)(Rez;x)1
whenever zr, zi E Sx' and x E Sx. It follows that there is a. 8 ,y such that E
whenever zi, Z2 E Sx*, x E Sx, and Re zix, Re zix ?:: 1 - 48f ,y.
For the moment, fix two elements zi and Z2 of Sx' such that
In any case, it follows that IziY - z2yl < E whenever zi, Z2 E Sx' and
11~(zi + z2)11 ?:: 1- D"y. Let Dx' be the w*UR modulus of X*. Then
such that Itml :::; 1/4 for each m and limm tm = 0, a sequence (xm) in Sx
and a corresponding sequence (x~) in Sx' such that each x~ supports Ex
at X m , and a positive E such that
(5.5)
and
for each m. Since (5.5) and (5.6) hold for each m, it follows that
which together with the fact that (5.7) holds for every m implies that
Corollary 5.6.17, along with the two examples by Mark Smith men-
tioned in the discussion following Theorem 5.6.12, can be used to show that
uniform Gateaux smoothness is genuinely different from both smoothness
and uniform smoothness, even for reflexive normed spaces. Smith showed
in [219J that (t'2, 11llw) is (wUR) but not (UR) and that (2, II IlL) is (R)
but not (wUR). It follows that (2, 11'llw)* is (UG) but not (US), while
(2, IIIIL)* is (S) but not (UG). As was previously mentioned, the space
(2, II'IIL)* is (F) while (2, 11'llw)* is not, so this also shows that neither
of the properties of Frechet smoothness and uniform Gateaux smoothness
implies the other.
One further generalization of uniform smoothness deserves mention be-
fore leaving the subject. It has been shown in Proposition 5.4.24, Corol-
lary 5.4.29, and Theorems 5.6.3 and 5.5.10 that a normed space X is
(a) smooth,
(b) Frechet smooth, or
(c) uniformly smooth
if and only if the spherical image map for Sx is singleton-valued and, when
viewed as a point-valued map, is
(a) norm-to-weak* continuous,
(b) norm-to-norm continuous, or
(c) uniformly norm-to-norm continuous,
5.6 Generalizations of Uniform Smoothness 515
respectively. The following definition fills the one conspicuous gap in this
list of continuity conditions.
The final result of this section then follows immediately from the obser-
vations made just before the preceding definition.
Exercises
5.62 (A. R. Lavaglia, 1955 [159]). Prove that a normed space is Fhkhet smooth
if its dual space is locally uniformly rotund.
5.63 (V. L. Smulian, 1940 [224]). Show that a Banach space is reflexive if
its dual space is both strongly rotund and Frechet smooth. Do not use
James's theorem or any result based on it, such as Exercise 5.33, in your
argument. (In fact, it follows from Exercise 5.33 and Theorem 5.6.9 that
a Banach space is reflexive if its dual space is either strongly rotund or
Frechet smooth.)
5.64 (V. L. Smulian, 1940 [224]). Suppose that X is a weakly sequentially
complete normed space with Schur's property (as is the case for every
normed space isomorphic to i\ j see Example 2.5.24). Prove that if X' is
smooth, then X' is Frechet smooth.
In particular, it follows from this exercise that if there is a norm 11lls*
on 1 equivalent to the standard norm of 1 such tha.t (1, II lis r is
smooth, then (1, II lis' ) is strongly rotund. However, this is a contradic-
tion since it would imply that 1 is reflexive; see Exercise 5.33. See also
Exercise 5.45. As was mentioned in Section 5.4, M. M. Day showed in a
1955 paper [51 J that 00 cannot be equivalently renormed to be smooth,
which is a stronger statement than saying that 1 cannot be equivalently
renormed to make its dual space smooth.
5.65 (a) Suppose that X is a strongly rotund Banach space, that Q is the
natural map from X into X", that x" E Sx'" tha.t x E Sx" and
that x"x' = 1. Prove that there is an x in Sx such that Qx = x".
Do not use any material from any optional sections of this book in
your argument.
516 5. Rotundity and Smoothness
This book is intended primarily as a text for Banach space courses taught
to graduate students in mathematics. The following are the prerequisites
for a course taught from this book at that level.
2. A first course in real analysis. This course should cover the topology
of the real line, continuous real-valued functions on the real line,
properties of the Riemann integral, and convergence of sequences and
series of real numbers and real-valued functions on the real line.
not to be treated, then omit Exercises 1.9 and 1.10 and the first part of
Exercise 1. 7.
Section 1.2. If complex vector spaces are not to be treated, then omit
the reference to en in Example 1.2.5. Omit Example 1.2.6, and replace the
paragraph following it by the single sentence "The roles of subscripts and
superscripts are often exchanged in the notations for the normed spaces
of the next three examples." Cover Appendix C through Definition C.9
before covering Examples 1.2.7-1.2.9, and omit the paragraph following
Example 1.2.7 as well as the last two sentences of Example 1.2.9. In Ex-
ample 1.2.10, replace "Hausdorff space" by "metric space" and omit the
reference to Lp[O, IJ. Omit Example 1.2.11, and in the paragraph follow-
ing it replace "topological space" by "metric space" and omit everything
after the argument that C(K) is complete. Cover Theorem C.lO and Corol-
lary C.ll after covering Definition 1.2.12. Omit Example 1.2.14. Delete the
reference to the Banach spaces Lp[O, 1] in the last sentence before the ex-
ercises. Omit Exercises 1.22, 1.23, and 1.28.
Section 1.3. In Corollary 1.3.4, replace "topological space" by "metric
space." Omit the paragraph immediately preceding Theorem 1.3.10. In
Exercise 1.30, replace "Hausdorff space" by "metric space."
Section 1.4. If complex vector spaces are not to be treated, then omit
part (c) of Exercise 1.48.
Section 1.5. Delete the last sentence of the second paragraph of the
section. In the third paragraph of the section, leave "topological meaning"
as it is, since this reference makes sense in the context of the topology of
a metric space, but change "topological space" to "metric space." In the
collection of definitions following this paragraph, change the first occurrence
of "topological space" to "metric space" and change the last sentence to "In
particular, the set A is of the first or second category in itself if A has that
category as a subset of the metric space (A, d A ), where dA is the restriction
of the metric of X to A x A." In Proposition 1.5.3, change "topological
space" to "metric space." Omit the first sentence of the last paragraph
before the exercises. In Exercises 1.53-1.57, replace "topological space"
or "topological spaces" by "metric space" or "metric spaces" respectively.
Omit Exercises 1.58, 1.62, and 1.63.
Section 1.6. In Definition 1.6.4, replace "topological space" by "metric
space." In the paragraph following that definition, omit all but the last
sentence. In the paragraph preceding Corollary 1.6.6, replace "topologi-
cal space" by "metric space." In Exercise 1.68, replace "topological space"
and "Hausdorff space" by "metric space." Omit Exercise 1.70. In Exer-
cise 1.71, replace "Hausdorff space" by "metric space." In Exercise 1.74,
replace "topological space" by "metric space."
Section 1.7. Omit Example 1. 7.2 and the last sentence of the paragraph
preceding it. In the paragraph following the proof of Theorem 1.7.4, omit
the three references to a topology. In the last sentence of the paragraph
preceding Definition 1.7.10, delete everything following "a quotient map."
520 Appendix A. Prerequisites
Omit the first two paragraphs following the proof of Proposition 1.7.12.
If it cannot be assumed that the student has seen the isomorphism theo-
rems for groups, omit the paragraph preceding Theorem 1.7.14 and skip
Exercise 1.87.
Section 1.8. Omit Exercise 1.92.
Section 1.9. If complex scalars are not to be treated, then omit the
following: beginning with the second sentence of the paragraph preceding
Definition 1.9.2, all material through Proposition 1.9.3; the last paragraph
of the proof of Theorem 1.9.6; the occurrence of "real-" in the paragraph
preceding Definition 1.9.13; the two occurrences of "Re" in the statement of
Proposition 1.9.15; the first sentence of the proof of Proposition 1.9.15; and
the single occurrence of "Re" in Exercise 1.109 as well as its two occurrences
in Exercise 1.110.
Section 1.10. Omit Example 1.10.2 and the two paragraphs following
it. Cover Theorem C.12 before covering Example 1.10.3, and replace the
first three sentences of that example with the following: "Suppose that
1 :S p < 00 and that q is conjugate to p. Then ; can be identified with q
in a natural way; see Theorem C.12." Omit the last two sentences of the
first paragraph of Example 1.10.4. Cover Theorem C.13 before covering
Example 1.10.5, then replace the entire text of that example with the fol-
lowing: "Suppose that n is a positive integer, that 1 :S p :S 00, and that q
is conjugate to p. Then (~)* can be identified with ~ in a natural way;
see Theorem C.13." Omit Example 1.10.6. Omit the last two sentences of
the paragraph following Theorem 1.10.7. If complex scalars are not to be
treated, then omit Exercise 1.113. Omit Exercise 1.121.
Section 1.11. Replace the statement of Theorem 1.11.10 by that of The-
orem C.14, and the proof of Theorem 1.11.10 by the statement "See The-
orem C.14." Omit the paragraph following the proof of Theorem 1.11.10.
Omit Examples 1.11.24 and 1.11.25. Omit Exercises 1.124-1.126, 1.129,
and 1.130.
Section 1.12. In the introductory paragraph of the section and in the
statement and proof of Proposi tion 1.12.9, replace each occurrence of "topo-
logical space" by "metric space." If complex scalars are not to be treated,
then replace the third sentence of the proof of Proposition 1.12.1 by "Let
lQo be the rationals." Omit Examples 1.12.3-1.12.5. In the paragraph fol-
lowing the proof of Corollary 1.12.12, omit the references to LdO, 1], C[O, 1],
L(XJ[O, 1], and rca[O, 1J. Omit Exercises 1.142, 1.143, and 1.145.
Section 1.13. Omit the first sentence of the paragraph following Defi-
nition 1.13.2. Omit Example 1.13.7 and Exercise 1.149. If complex scalars
are not to be treated, then omit the following: all occurrences of the sym-
bol "Re"; the second paragraph of the section; Proposition 1.13.1; the first
sentence of the proof of Theorem 1.13.4; the third sentence of the sec-
ond paragraph following the proof of Theorem 1.13.4; the proof of Theo-
rem 1.13.15; and Exercise 1.146.
Appendix B
Metric Spaces
The only results from topology that are used in a crucial way in Chapter 1
are about metric spaces. These results are standard fare in most general
topology texts, but a development of metric spaces independent of gen-
eral topology, such as can be found in Kaplansky's Set Theory and Metric
Spaces [129] and in many elementary real analysis texts, is sufficient. For
reference, here are the definitions and results from metric space theory
needed for Chapter 1.
For the rest of this appendix, let M, M 1 , ... ,MN be sets with respective
metrics d, d 1 , ... ,d N .
B.2 Proposition. If S is a subset of M, then S is itself a metric space
with the metric inherited from M.
B.50 Definition. For each pair (XI, ... ,XN),(YI, ... ,YN) of elements of
the Cartesian product M1 x ... X MN, let
N 1/2
dp ((X1,". ,XN), (YI, ... , YN)) = (~(dj(XJ' YJ))2)
. (x n)
11m (n) )
l ' ... , X N = ( Xl, .. , X N )
n
B.55 Definition. Two metrics on the same set are (topologically) equiva-
lent if they induce the same topology.
Clearly, a set that is closed with respect to a metric d is also closed with
respect to every metric equivalent to d. Such properties that depend not
on the particular metric but rather only on the topology induced by the
metric are called topological properties. It is easy to see that convergence of
a sequence to a particular limit, compactness, separability, and continuity
are topological properties. Cauchyness, completeness, and boundedness are
not. See Exercise 1.42 for an example of a metric don lR that is equivalent
to the usual metric on lR and yet has the property that lR is bounded and
incomplete under d; the incompleteness implies the existence of a sequence
of reals that is not Cauchy in the usual sense but that is d-Cauchy.
Notice that two metrics on the same set are equivalent if and only if
the identity map on the set, treated as a function between the two metric
spaces, is a homeomorphism.
The spaces fp and f;, where 1 :::; p :::; 00 and n E N, are treated in Chapter 1
as the Lebesgue spaces Lp(0., '., J-t), where 0. is N or the set {I, ... , n} and J-t
is the counting measure on the a-algebra'. of all subsets of 0.. The purpose
of this section is to provide a more elementary development of the spaces
fp and f; paralleling that given in Chapter 1 but requiring only as much
knowledge of analysis as would come from a first undergraduate course in
real analysis without measure theory, and no knowledge of general topology
beyond the basic facts about metric spaces from Appendix B. For the sake
of completeness, a few words will also be said about the trivial normed
space f~.
In the following definition, the pth root of infinity is to be interpreted to
be infinity when 1 :::; p < 00.
C.l Definition. Suppose that 1 :::; P :::; 00 and that X is the vector space of
all sequences of scalars with the usual vector space operations, that is, with
addition of sequences and multiplication of sequences by scalars performed
term by term. For each member (OJ) of X, let
Then the p-norm on X is the function II lip : X --7 [0,00]. The collection
of all sequences (aj) of scalars such that II(aj)llp is finite is denoted by fp
(pronounced "little ell p" ).
530 Appendix C. The Spaces t p and t;, 1 ::; P ::; 00
The p-norms are not really norms on the vector space X of the preceding
definition. For example, if aj = j for each j, then lI(aj)llp = 00 when
1 ::; p ::; 00. It does turn out that each of the spaces fp is a subspace of X
and has the corresponding p-norm as a norm. To show this, the first order
of business is to obtain some classical inequalities.
Notice that 1 ::; q ::; 00 for each conjugate exponent q, and that if q is
conjugate to p, then p is conjugate to q. Notice also that q = 2 when p = 2,
and that 2 is the only value of p that is its own conjugate exponent.
C.3 Lemma. Suppose that 1 < p < 00 and that q is conjugate to p. Then
rP sq
rs < -
- p
+-q
for all nonnegative real numbers rand s.
PROOF. Suppose that 0 < a < 1 and that f(t) = t a - at when t > O. It is a
straightforward calculus exercise to show that f takes on its maximum value
when t = 1, so t a - at ::; 1- a when t > O. Substituting the quotient u/v of
two positive numbers for t, multiplying by v, and rearranging the resulting
inequality shows that
when u, v > O. It is clear that this inequality holds, in fact, when u, v ;::: O.
Finally, let rand s be nonnegative reals, let u = r P , let v = sq, let a = p-l
(so that 1 - a = q-l), and substitute into the above inequality to finish
the proof.
The interpretation of one special case of the following result relies on the
usual convention that 0 . 00 = O.
C.4 Holder's Inequality for Sequences. Suppose that 1 ::; p ::; 00 and
that q is conjugate to p. Then
reals. Now suppose that the conclusion of Holder's inequality holds when
II(aj)l!p = 11(,8j )llq = 1. For the general case, let, = II(aj)llp and 0 =
IIC8j)llq, so that ,,0 E (0,00). It is easy to check that
II (aj) II; 111 (fJj) 11;111 (ajfJj )Ih = II b- 1ajo-l fJj) III
::; I!b-1aj)llp II(o-lfJj)llq
=1.
j j
as required.
Finally, suppose that 1 < p < 00. It follows from the lemma that
::; Lj
(~+
p
IfJjl q )
q
= II(aj)ll~ + l!(fJj)ll~
p q
1 1
=-+-
p q
= 1,
C.5 Minkowski's Inequality for Sequences. Suppose that 1 ::; P ::; 00.
Then
j j
If p= 00, then lak + (3kl ::; lakl + l!Jkl ::; II(aj)lloo + 11({3j)lloo for each k, so
IICaj) + ({3j) 1100 ::; lI(aj)lloo + IIC!Jj)lloo.
Finally, suppose that 1 < p < 00, and let q be conjugate to p. It may be
assumed that II(aj) + ({3j)llp is nonzero and that both II(aj)llp and 1I({3j)llp
are finite. Since
j j
the finiteness of II (OJ) lip and II (,Bj) lip implies that of II (O:j) + (,Bj) lip, and so
0< II(aj) + (,Bj) lip < 00. Next, notice that
j j
Now apply Holder's inequality to the sequences (Iaj I) and (Iaj + ,Bj IP-l),
which yields the inequality
C.6 Theorem. Suppose that 1 S; P S; 00. Then 1!p is a vector space when
the sum of two sequences of scalars and the product of a scalar and a
sequence of scalars are defined in the usual way. The p-norm is a norm on
this vector space.
PROOF. It is easy to check that 0:( aj) E 1!p whenever (aj) E 1!p and a
is a scalar. Also, each sum of two elements of 1!p is in 1!p by Minkowski's
inequality. Since the zero sequence is in 1!p, it follows that 1!p is a subspace
of the vector space of all sequences of scalars with the usual operations.
Minkowski's inequality provides a triangle inequality for the p-norm, while
the other properties that the p-norm must have to be a norm on 1!p follow
quickly from the definitions.
C.lO Theorem. Suppose that 1 ::::: p ::::: 00. Then Pp is a Banach space, as
is e; for each nonnegative integer n.
PROOF. Let ((ajk));:'=l be a Cauchy sequence of elements of ep- Since
la(k)
Jo
- a
Jo
I= lim la(k) - a O) I <
I Jo Jo-
E
for each ]0 when k :::: N E Taking the supremum over all ]0 shows that
II(a;k) - (aj)ll= :::; E when k :::: N(. Then
Now suppose that 1 ::; p < 00. Let f be a positive real number and let N.
be a positive integer such that II(ajkl) - (ojll)lIp ::; f when k, I ;::: N . It
follows that for each positive integer )0,
when k, I ;::: N . Leaving jo fixed and letting I tend to infinity shows that
when k 2: N,. Therefore (OJ) E fp by the argument used for foo, and
II(a;kl) - (aj)lIp -7 0 as k -7 00. This completes the proof for f p.
Suppose that n is a positive integer and that ((a~k), ... , a~k)):l is a
Cauchy sequence in f;.
Then each sequence (o;k)k'=l such that] = 1, ... , n
is a Cauchy sequence of scalars and so has a limit aj. Depending on whether
or not p is finite, either
or
as k -7 00, and so
. (a (k)
hm 1 , . .. , an(k) -_ (a1,., On ) .
k
The following two results are necessary for Examples 1.10.3 and 1.10.5
if the duality theory for p and e e;
is not to be obtained from the more
general duality theory for Lebesgue spaces.
= x* (t 'Y l!3jIQ-1 ej )
J=l
j
(C.3)
which, by taking the supremum over all j, again shows that ({3j) E Cq and
that (C.3) holds. Thus, whether or not p = 1, it follows that ({3j) E Cq and
therefore that T({3i) E C;. For every element (OJ) of Cp ,
(C.4)
PROOF. For each element (31, ... ,{3n) of ~, the map T({31, ... ,{3n) is
clearly a linear functional on the finite-dimensional normed space ;, and
so is automatically bounded. It follows that T is a linear operator from ~
into (1:;)*. It is easy to check that ker(T) = {(o, ... , On,
so T is one-to-
one. Since ~ and (;)* both have dimension n, the operator T maps ~
onto (;)*.
538 Appendix C. The Spaces Cp and C;, 1 :S p :S 00
Let ((31, ... , (370) be an element of e~. All that remains to be shown is that
IIT((31,..,(3n)11 = 11((31, ... ,(3n)llq Suppose that (01, ... ,070 ) E and e;
that (01, ... , On) and ((31, ... , (370) are extended to the respective sequences
(OJ) and ((3j) by letting OJ = (3j = 0 when j > n. Holder's inequality
implies that
70
and so
(C.5)
Now let {e1"'" en} be the standard basis for the vector space lF n and let
11, ... "n be scalars of absolute value 1 such that l(3jl = Ij(3j for each j.
If 1 < p < 00, then
n n
LI(3jlq = L Ij(3jl(3jlq-l
j=l j=1
If p = 1, then
If p = 00, then
n
IICBI, ... ,,Bn)lh = 'Elj,Bj
j=l
These terms could have been used to define net convergence and accu-
mulation, since a net converges to (respectively, accumulates at) a point if
and only if the net is ultimately (respectively, frequently) in U whenever U
is a neighborhood of the point. However, these terms would not have con-
tributed much to the discussion of convergence and accumulation, while
they will greatly streamline the presentation of ultranets that is to follow.
Notice that a net is in a set S frequently if and only if it is not ultimately
in the complement of S. Clearly, every net has the property that if it is
ultimately in a set, then it must be frequently in that set. Ultranets are
nets having the converse property.
542 Appendix D. Ultranets
Suppose that (xc is a net in a set X. If S c::; X, then (xc must frequent
either S or X \ S, from which it follows immediately that an ultranet in X
must be either ultimately in S or ultimately in X \ S. Conversely, suppose
that (x o.) has the property that for each subset S of X, the net is either
Ultimately in S or ultimately in X \ S. Since no net can be frequently in
a set and ultimately in the complement of that set, it follows that (xc is
ultimately in a set whenever it is frequently in that set, and so is an ultranet.
This gives the following result, which is often used to define ultranets.
D.4 Example. Let x be an clement of a set X and let 9J1 x be the collection
of all subsets of X containing x. Let I be the set of all ordered pairs (a, A)
such that A E 9J1 x and a E A, directed by declaring that (a, A) ~ (b, B)
whenever A :.2 B. Let X(a,A) = a whenever (a, A) E I. Then (X(a,A)) is a
net in X. Notice that for each subset S of X, the net (X(a,A)) lies either
entirely inside or entirely outside S from some term onward, depending OIl
whether or not xES. The net (X(a,A)) is therefore an ultranet.
It turns out that every net has a subnet that is an ultranet. The proof of
that to be given here depends on the following technical lemma. Readers
familiar with the theory of filters will recognize this lemma as a statement
about ultrafilters.
Appendix D. Ultranets 543
Notice that the following result is stated only for topological groups.
[1] Naum I. Akhiezer and Mark G. Krein, Some questions in the theory of
moments, Gosudarstv. Naucno-Tehn. Izdat. Ukrain., Kharkov, 1938
(Russian, translated into English as [2]).
[6] Cesare Arzela, Funzioni di linee, Atti della R. Accad. dei Lincei Ren-
diconti della Cl. Sci. Fis. Mat. Nat. (4) 5 [1 0 Sem.] (1889),342-348.
548 References
[7] Giulio Ascoli, Le curve limite di una varietiL data di curve, Atti della
R. Accad. dei Lincei Memorie della Cl. Sci. Fis. Mat. Nat. (3) 18
(1882-1883), 521-586.
[8] Guido Ascoli, Sugli spazi lineari metrici e le loro varieta lineari, Ann.
Mat. Pura Appl. (4) 10 (1932),33-81,203-232.
[9] Louis Rene Baire, Sur les fonctions de variables n~elles, Ann. Mat.
Pura Appl. (3) 3 (1899), 1-122.
[10] Stefan Banach, Sur les operations dans les ensembles abstraits et leur
application aux equations integrales, Fund. Math. 3 (1922), 133-18l.
[16] Stefan Banach and Stanislaw Mazur, Zur Theorie der linearen Di-
mension, Studia Math. 4 (1933), 100-112.
[18] Robert G. Bartle, Nets and filters in topology, Amer. Math. Monthly
62 (1955), 551-557; MR 17, 39l.
[19] ___ , The elements of real analysis, second ed., Wiley, New York,
1976; MR 52 #14179.
[37] Richard D. Bourgin, Geometric aspects of convex sets with the Radon-
Nikodym property, Lecture Notes in Mathematics, vol. 993, Springer-
Verlag, Berlin, 1983; MR 85d:46023.
[46] Mahlon M. Day, The spaces LP with 0 < p < 1, Bull. Amer. Math.
Soc. 46 (1940), 816-823; MR 2, 102, MR 2, 419.
[48] ___ , Some more uniformly convex spaces, Bull. Amer. Math. Soc.
47 (1941), 504-507; MR 2,314.
References 551
[53] ___ , Normed linear spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete, vol. 21, Springer-Verlag, Berlin, 1958; MR 20 #1187.
[54] ___ , Normed linear spaces, second (corrected) ed., Ergebnisse der
Mathematik und ihrer Grenzgebiete, voL 21, Springer-Verlag, Berlin,
1962; MR 20 #1187, MR 26 #2847.
[55] ___ , On the basis problem in normed spaces, Proc. Amer. Math.
Soc. 13 (1962), 655-658; MR 25 #1435.
[56] ___ , Normed linear spaces, third ed., Ergebnisse der Mathematik
und ihrer Grenzgebiete, voL 21, Springer-Verlag, Berlin, 1973; MR 20
#1187, MR 26 #2847, MR 49 #9588.
[59] Joseph Diestel and J. Jerry Uhl, Vector measures, Mathematical Sur-
veys, vol. 15, American Mathematical Society, Providence, R.I., 1977;
MR 56 #12216.
[66] Nelson Dunford and Billy James Pettis, Linear operations on sum-
mable functions, Trans. Amer. Math. Soc. 47 (1940), 323-392; MR I,
338.
[75] Per Enfto, A Banach space with basis constant> 1, Ark. Mat. 11
(1973), 103-107; MR 49 #7736.
[86] Bernard R. Gelbaum, Notes on Banach spaces and bases, An. Acad.
Brasil. Cimc. 30 (1958), 29-36; MR 20 #5419.
[87] John R. Giles, On smoothness of the Banach space embedding, Bull.
Austral. Math. Soc. 13 (1975),69-74; MR 52 #1268.
[88] ___ , Convex analysis with application in the differentiation of
convex functions, Research Notes in Mathematics, vol. 58, Pitman,
Boston, 1982; MR 83g:46001.
[89] Herman H. Goldstine, Weakly complete Banach spaces, Duke Math.
J.4 (1938), 125131.
[90] W. Timothy Gowers, Symmetric block bases of sequences with large
average growth, Israel J. Math. 69 (1990), 129-151; MR 91i:4601O.
[91] ___ , A Banach space not containing co, 1 or a reflexive subspace,
Trans. Amer. Math. Soc. 344 (1994), 407-420; MR 94j:46024.
[92] ___ , A solution to Banach's hyperplane problem, Bull. London
Math. Soc. 26 (1994), 523-530; MR 96a:46025.
554 References
[98] Hans Hahn, tiber Folgen linearer Operationen, Monatsh. Math. Phys.
32 (1922), 3-88.
[100] Paul R. Halmos, "Nicolas Bourbaki," Sci. Amer. 196 (1957), no. 5,
88-99; MR 18, 709.
[101] Felix Hausdorff, Grundzuge der Mengenlehre, Verlag von Veit, Leip-
zig, 1914.
[102] Eduard Reily, tiber Systeme linearer Gleichungen mit unendlich viel-
en Unbekannten, Monatsh. Math. Phys. 31 (1921),60-91.
[103] David Hilbert, Grundzuge einer allgemeinen Theorie der linearen In-
tegralgleichungen, IV, N achr. Kg!. Gesells. Wiss. Gottingen Math.-
Phys. Kl. (1906), 157 -227.
[105] Kenneth Hoffman and Ray Kunze, Linear algebra, second ed.,
Prentice-Hall, Englewood Cliffs, N.J., 1971; MR 43 #1998.
[108] Robert C. James, Bases and reflexivity of Banach spaces, Bull. Amer.
Math. Soc. 56 (1950), 58, Abstract 80.
[115J _ _ , Weakly compact sets, Trans. Amer. Math. Soc. 113 (1964),
129-140; MR 29 #2628.
[121] _ _ , Letter to the editor, Izv. Vyssh. Uchebn. Zaved. Mat. 1961
(1961), no. 6, 186-187 (Russian); MR 25 #3342. This contains a
correction to [120].
556 References
[129J Irving Kaplansky, Set theory and metric spaces, Allyn and Bacon,
Boston, 1972; MR 56 #11795.
[133] ___ , Convex sets in linear spaces, Duke Math. J. 18 (1951), 443-
466; MR 13, 354.
[134J ___ , Convex sets in linear spaces, II, Duke Math . .1. 18 (1951),
875-883; MR 13, 849.
[140] Konrad Knopp, Infinite sequences and series, Dover, New York, 1956;
MR 18, 30.
[142] Vladimir Ya. Kozlov, On bases in the space L2[0, 1], Mat. Sb. (N.S.)
26 (1950),85-102 (Russian); MR 11, 602.
[151] ___ , Sur les integrales singulier-es, Ann. Fac. Sci. Univ. Toulouse
(3) 1 (1909), 25-117.
[157] ___ , Classical Banach spaces II: FUnction spaces, Ergebnisse der
Mathematik und ihrer Grenzgebiete, vol. 97, Springer-Verlag, Berlin,
1979; MR 81c:46001.
[160] R. Daniel Mauldin (ed.), The Scottish book: Mathematics from the
Scottish Cate, Birkhauser, Boston, 1981; MR 84m:00015.
[1611 Stanislaw Mazur, Uber die kleinste konvexe Menge, die eine gegebene
kompakte Menge enthiilt, Studia Math. 2 (1930), 7-9.
[164] Stanislaw Mazur and Wladyslaw Orlicz, Uber Folgen linearer Opera-
tionen, Studia Math. 4 (1933), 152-157.
[165] Stanislaw Mazur and L. Sternbach, Uber die Borelschen Typen von
linearen Mengen, Studia Math. 4 (1933),48-53.
[174] Masahiro Nakamura and Shizuo Kakutani, Banach limits and the
eech compactification of a countable discrete set, Proc. Imp. Acad.
Tokyo 19 (1943), 224-229; MR 7, 306.
[181] Billy James Pettis, A note on regular Banach spaces, Bull. Amer.
Math. Soc. 44 (1938), 420-428.
[182] ___ , On integration in vector spaces, Trans. Amer. Math. Soc. 44
(1938), 277-304.
[183] ___ , A proof that every uniformly convex space is reflexive, Duke
Math. J. 5 (1939), 249~253.
[198) Ira Rosenholtz, Another proof that any compact metric space is the
continuous image of the Cantor set, Amer. Math. Monthly 83 (1976),
646-647; MR 54 #6064.
[199) Haskell P. Rosenthal, A characterization of Banach spaces contain-
ing e1 , Proc. Nat. Acad. Sci. U.S.A. 71 (1974), 2411-2413; MR 50
#10773.
[200) Walter Rudin, Functional analysis, McGraw-Hill, New York, 1973;
MR 51 #1315.
[201) ___ , Principles of mathematical analysis, third ed., McGraw-Hill,
New York, 1976; MR 52 #5893.
[202) ___ , Real and complex analysis, third ed., McGraw-Hill, New
York, 1987; MR 88k:00002.
[203) Anthony F. Ruston, A note on convexity in Banach spaces, Proc.
Cambridge Philos. Soc. 45 (1949), 157-159; MR 10, 197.
[213] Wadaw Sierpinski, Sur les ensembles complets d'un espace (D), Fund.
Math. 11 (1928), 203~205.
[214] ___ , Cardinal and ordinal numbers, Polska Akademia Nauk, Mon-
ografie Matematyczne, vol. 34, Panstwowe Wydawnictwo Naukowe,
Warsaw, 1958; MR 20 #2288.
[218] ___ , Banach spaces that are uniformly rotund in weakly compact
sets of directions, Canad. J. Math. 29 (1977), 963~970; MR 56 #9232.
[220] ___ , A Banach space that is ML UR but not HR, Math. Ann. 256
(1981), 277~279; MR 82h:46032.
[223] ___ , On the principle of inclusion in the space of the type (B),
Mat. Sb. (N.S.) 5 (1939), 317 328 (Russian, English summary);
MR 1, 335.
[230] Stanislaw J. Szarek, A Banach space without a basis which has the
bounded approximation property, Acta Math. 159 (1987), 81-98;
MR 88f:46029.
[233] John W. Tukey, Some notes on the separation of convex sets, Portu-
gal. Math. 3 (1942),95--102; MR 4, 13.
[245] K6saku Yosida, Mean ergodic theorem in Banach spaces, Proc. Imp.
Acad. Tokyo 14 (1938), 292-294.
Fraktur Letters
Here are the uppercase Roman letters, each followed by its Fraktur equiv-
alent:
Other Symbols
Symbols used as the names of particular spaces, such as co, have in most
cases been placed in the Index rather than here. Also, most symbols in-
vented for temporary use, such as those in the discussion of arrow vectors
at the beginning of Section 1.2, have been omitted from this list.
Miscellaneous Symbols
0 .......................... 2, 153 G(x,y) ....................... 485
j(n) .......................... 14 lmo: .......................... 71
(x,!) ......................... 89 K(x*, t) ...................... 274
[f satisfies P] ............... 162 L(x~) ........... ............. 127
I-llf satisfies Pl .... .... 162 v(x) ......................... 489
f'(a) ........................ 311 Rea .......................... 71
'I~ ........................... 203 r,,(x) ........................ 312
(x o,) . . . . . . . . . . . . . . . . . . . . . . . . . 143 Rx .................... . .309
A(IOl) ......................... 119 p(x) ......................... 309
Bx ........................... 10 Sx ............................ 10
B(x,r) ......................... 5 cr(x) ......................... 309
B(x, (x~)) ................... 244 cr(X, J) ...................... 203
B(x*, (x n )) . . . . . . . . . . . . . . . . . . . 235 cr(X,X*) ..................... 212
e ............................. 152 cr(X*, X) ..................... 223
G _ (x, y) ..................... 485 V(x;J ........................ 127
G+(x, y) ..................... 485
Index