You are on page 1of 253

Linear Analysis

LINEAR ANALYSIS
An Introductory Course
Bela Bollobas
University of Cambridge

CAMBRIDGE
UNIVERSITY PRESS
PUBLISHED BY THE PRESS SYNDICATE OF THE UNIVERSITY OF CAMBRIDGE
The Pitt Building, Trumpington Street, Cambridge CB2 1RP, United Kingdom

CAMBRIDGE UNIVERSITY PRESS


The Edinburgh Building, Cambridge CB2 2RU, UK http://www.cup.cam.ac.uk
40 West 20th Street, New York, NY 10011-4211, USA http://www.cup.org
10 Stamford Road, Oakleigh, Melbourne 3166, Australia

First edition © Cambridge University Press 1990


Second edition © Cambridge University Press 1999

This book is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without
the written permission of Cambridge University Press.

First published 1990


Second edition 1999

Typeset in Times 10/13pt

A catalogue record for this book is available from the British Library

ISBN 0 521 65577 3 paperback

Transferred to digital printing 2004


To Mark
Qui cupit, capit omnia
CONTENTS

Preface ix

1. Basic inequalities 1

2. Normed spaces and bounded linear operators 18

3. Linear functionals and the Hahn-Banach theorem 45

4. Finite-dimensional normed spaces 60

5. The Baire category theorem and the closed-graph theorem 75

6. Continuous functions on compact spaces and the Stone-

Weierstrass theorem 85

7. The contraction-mapping theorem 101

8. Weak topologies and duality 114

9. Euclidean spaces and Hilbert spaces 130

10. Orthonormal systems 141

11. Adjoint operators 155

12. The algebra of bounded linear operators 167

vn
viii Contents

13. Compact operators on Banach spaces 186

14. Compact normal operators 198

15. Fixed-point theorems 213

16. Invariant subspaces 226

Index of notation 233

Index of terms 235


PREFACE

This book has grown out of the Linear Analysis course given in Cambridge
on numerous occasions for the third-year undergraduates reading
mathematics. It is intended to be a fairly concise, yet readable and down-
to-earth, introduction to functional analysis, with plenty of challenging
exercises. In common with many authors, I have tried to write the kind of
book that I would have liked to have learned from as an undergraduate. I
am convinced that functional analysis is a particularly beautiful and
elegant area of mathematics, and I have tried to convey my enthusiasm to
the reader.
In most universities, the courses covering the contents of this book are
given under the heading of Functional Analysis; the name Linear Analysis
has been chosen to emphasize that most of the material in on linear func-
tional analysis. Functional Analysis, in its wide sense, includes partial
differential equations, stochastic theory and non-commutative harmonic
analysis, but its core is the study of normed spaces, together with linear
functionals and operators on them. That core is the principal topic of this
volume.
Functional analysis was born around the turn of the century, and within
a few years, after an amazing burst of development, it was a well-
established major branch of mathematics. The early growth of functional
analysis was based on 19th century Italian function theory, and was given
a great impetus by the birth of Lebesgue's theory of integration. The
subject provided (and provides) a unifying framework for many areas:
Fourier Analysis, Differential Equations, Integral Equations, Approxima-
tion Theory, Complex Function Theory, Analytic Number Theory, Meas-
ure Theory, Stochastic Theory, and so on.

IX
x Preface

From the very beginning, functional analysis was an international sub-


ject, with the major contributions coming from Germany, Hungary,
Poland, England and Russia: Fisher, Hahn, Hilbert, Minkowski and
Radon from Germany, Fejer, Haar, von Neumann, Frigyes Riesz and
Marcel Riesz from Hungary, Banach, Mazur, Orlicz, Schauder, Sierpinski
and Steinhaus from Poland, Hardy and Littlewood from England, Gel-
fand, Krein and Milman from Russia. The abstract theory of normed
spaces was developed in the 1920s by Banach and others, and was
presented as a fully fledged theory in Banach's epoch-making monograph,
published in 1932.
The subject of Banach's classic is at the heart of our course; this
material is supplemented with a body of other fundamental results and
some pointers to more recent developments.
The theory presented in this book is best considered as the natural
continuation of a sound basic course in general topology. The reader
would benefit from familiarity with measure theory, but he will not be at a
great disadvantage if his knowledge of measure theory is somewhat shaky
or even non-existent. However, in order to fully appreciate the power of
the results, and, even more, the power of the point of view, it is advisable
to look at the connections with integration theory, differential equations,
harmonic analysis, approximation theory, and so on.
Our aim is to give a fast introduction to the core of linear analysis, with
emphasis on the many beautiful general results concerning abstract spaces.
An important feature of the book is the large collection of exercises, many
of which are testing, and some of which are quite difficult. An exercise
which is marked with a plus is thought to be particularly difficult. (Need-
less to say, the reader may not always agree with this value judgement.)
Anyone willing to attempt a fair number of the exercises should obtain a
thorough grounding in linear analysis.
To help the reader, definitions are occasionally repeated, various basic
facts are recalled, and there are reminders of the notation in several
places.
The third-year course in Cambridge contains well over half of the con-
tents of this book, but a lecturer wishing to go at a leisurely pace will find
enough material for two terms (or semesters). The exercises should cer-
tainly provide enough work for two busy terms.
There are many people who deserve my thanks in connection with this
book. Undergraduates over the years helped to shape the course;
numerous misprints were found by many undergraduates, including John
Longley, Gabor Megyesi, Anthony Quas, Alex Scott and Alan Stacey.
Preface xi

I am grateful to Dr Pete Casazza for his comments on the completed


manuscript. Finally, I am greatly indebted to Dr Imre Leader for having
suggested many improvements to the presentation.

Cambridge, May 1990 Bela Bollotas

For this second edition, I have taken the opportunity to correct a number of
errors and oversights. I am especially grateful to R. B. Burckel for providing
me with a list of errata.

B. B.
/. BASIC INEQUALITIES

The arsenal of an analyst is stocked with inequalities. In this chapter we


present briefly some of the simplest and most useful of these. It is an
indication of the size of the subject that, although our aims are very
modest, this chapter is rather long.
Perhaps the most basic inequality in analysis concerns the arithmetic
and geometric means; it is sometimes called the AM-GM inequality.
The arithmetic mean of a sequence a = (at,..., an) of n reals is

Ma)
- =,!,
if each at is non-negative then the geometric mean is

V=l

where the non-negative nth root is taken.


Theorem 1. The geometric mean of n non-negative reals does not
exceed their arithmetic mean: if a = (ai,...,an) then
G(a)^A(a). (1)
Equality holds iff at = • • • = an.
Proof. This inequality has many simple proofs; the witty proof we shall
present was given by Augustin-Louis Cauchy in his Cours d'Analyse
(1821). (See Exercise 1 for another proof.) Let us note first that the
theorem holds for n = 2. Indeed,
(Q\~a2) — #i — 1a\a2 + a2 ^ 0;
Chapter 1: Basic inequalities

so
(a1 + a2)2
with equality iff ax = a2.
Suppose now that the theorem holds for n = m. We shall show that
it holds for n = 2m. Let ai,...,am,bi,...,bm be non-negative reals.
Then

ra m

2m
If equality holds then, by the induction hypothesis, we have ax =
' ' ' ~ am~ ^\~ ''' = bm. This implies that the theorem holds when-
ever n is a power of 2.
Finally, suppose n is an arbitrary integer. Let
n
1
n <2k = N and a =

Set #„+! = • • • = aN = a. Then

so

1= 1

with equality iff flj = • • • = aN, in other words iff ^ = • • • = # „ . D

In 1906 Jensen obtained some considerable extensions of the AM-GM


inequality. These extensions were based on the theory of convex func-
tions, founded by Jensen himself.
A subset D of a real vector space is convex if every convex linear
combination of a pair of points of D is in D, i.e. if x,y E D and
0 < f < l imply that tx + (l-t)y E D. Note that if D is convex,
xl9...,xneD, tl9...,tn>0 and 2 " = 1 h = 1 then J ^ ^ e D .
Chapter 1: Basic inequalities

Indeed, assuming that a convex linear combination of n-\ points of D


is in D, we find that

x2 =
i

and so
n

1=1

Given a convex subset D of a real vector space, a function / : D —> R


is said to be convex if
/ 0 * + (1 -t)y) ^ tf(x) + (1 -fl/OO (2)
whenever x,y E D and 0 < f < 1. We call / strictly convex if it is con-
vex and, moreover, f(tx + (l-t)y) = r/(jc) + (l-f)/()0 and 0 < t < 1
imply that x = y. Thus / is strictly convex if strict inequality holds in
(2) whenever x ^ y and 0 < t < 1. A function / is concave if - / is
convex and it is strictly concave if - / is strictly convex. Clearly, / is
convex iff the set {(x,y) E DxR: y ^ f(x)} is convex.
Furthermore, a function/: D —» IR is convex (concave, ...) iff its res-
triction to every interval [a,b] = {ta + (l-t)b : 0 ^ t ^ 1} in D is con-
vex (concave, . . . ) . Rolle's theorem implies that if/: (a,b) —> (R is dif-
ferentiable then / is convex iff / ' is increasing and / is concave iff / ' is
decreasing. In particular, if / is twice differentiable and / " ^ 0 then / is
convex, while if / " ^ 0 then / is concave. Also, if / " > 0 then / is
strictly convex and if/" < 0 then / is strictly concave.
The following simple result is often called Jensen's theorem; in spite
of its straightforward proof, the result has a great many applications.
Theorem 2. L e t / : D —» R be a concave function. Then
n I n \
y t fix) $ n T t x \ (3)

whenever x1,...,xnE D, tx,..., tn E (0,1) and 2" = 1 *i = 1- Further-


more, if / is strictly concave then equality holds in (3) iff xx = • • • = xn.
Proof. Let us apply induction o n n . As for n — 1 there is nothing to
prove and for n = 2 the assertions are immediate from the definitions,
let us assume that n ^ 3 and the assertions hold for smaller values of n.
Chapter 1: Basic inequalities

Suppose first that / is concave, and let


n
xl9...,xn ED, tu...,tn E (0,1) with 2 4 = I-

For i = 2,...,/i, set 4' = 4 / ( l - * i ) , so that 2" = 2 '/ = 1- Then, by


applying the induction hypothesis twice, first for n — 1 and then for 2, we
find that
n n

l l
i=l i=2

.1=2

L-'I) 2 to
i=2

If / is strictly concave, n ^ 3 and not all JC,- are equal then we may
assume that not all of x2,..., xn are equal. But then

d-^i) £ ^ / w < a - ^ (
i=2 \i=2 /

so the inequality in (3) is strict. •

It is very easy to recover the AM-GM inequality from Jensen's


theorem: logjc is a strictly concave function from (0,°°) to R, so for
ax,..., an > 0 we have
H n n
a
1 V* * V* i
- 2, iog«,-«log 2, - .
i=l i=l

which is equivalent to (1). In fact, if tx,..., tn > 0 and 2 " 4 = 1 then

2 fjlogx,- ^ l o g 2 ^ i , (4)
1=1 1=1

with equality iff x\ = •# • = JCW , giving the following extension of


Theorem 1.

Theorem 3. Let al9...,an^0 and pl9...,pn > 0 with 2 " = 1 A = 1-


Then
Chapter 1: Basic inequalities

ft af'^ ipiOi, (5)


1=1 i= l

with equality iff ax = • • • = an.


Proof. The assertion is trivial if some at is 0; if each at is positive, the
assertion follows from (4). •

The two sides of (5) can be viewed as two different means of the
sequence ai,...,an: the left-hand side is a generalized geometric mean
and the right-hand side is a generalized arithmetic mean, with the vari-
ous terms and factors taken with different weights. In fact, it is rather
natural to define a further extension of these notions.
Let us fix px,...,pn > 0 with S" = 1 Pi = 1- the pt will play the role of
weights or probabilities. Given a continuous and strictly monotonic
function cp : (0,o°) —> R, the <p-mean of a sequence a = (a1,...,an)
(at > 0) is defined as

Note that M^ need not be rearrangement invariant: for a permutation it


the <p-mean of a sequence ai,...,an need not equal the <p-mean of the
sequence a^x) > • • • > a7r(«) • Of course, if px = • • • = pn = l/n then every
<p-mean is rearrangement invariant.
It is clear that

min at ^ M9(a) ^ max ^ .

In particular, the mean of a constant sequence (ao,...,ao) is precisely


a0.
For which pairs <p and ^ are the means M^ and Af^, comparable?
More precisely, for which pairs <p and </> is it true that M^ia) ^ M^a)
for every sequence a = («!,...,fl w ) (^ > 0)? It may seem a little
surprising that Jensen's theorem enables us to give an exact answer to
these questions (see Exercise 31).
Theorem 4. Let p1,...,pn > 0 be fixed weights with 2 n = 1 Pi ~ 1 a n d let
<P,I/J: (0,o°) -» IR be continuous and strictly monotone functions, such
that (p4f~l is concave if cp is increasing and convex if cp is decreasing.
Then
Chapter 1: Basic inequalities

for every sequence a = (ai,...,an) (at > 0). If cpif/ l is strictly concave
(respectively, strictly convex) then equality holds iff a\ = • • • = an.

Proof. Suppose that cp is increasing and <p^~l is concave. Set bt = i/K^/)


and note that, by Jensen's theorem,

1
/' 1 is strictly concave and not all a, are equal then the inequality
If (pi//'
above is strict since not all bt are equal.
The case when cp is decreasing and (pip~l is convex is proved analo-
gously. •

When studying the various means of positive sequences, it is con-


venient to use the convention that a stands for a sequence (a1,...,an),
b for a sequence (b1,...,bn) and so on; furthermore,

a-1 = - = (flf 1 ,...,fl / T 1 ), a + x = (a1+x,...,an+x) (x E U+),


aa
ab = (a1bu...,anbn), abc = {axbxcx,...,anbncn),
and so on.
If <p(t) = tr (-00 < r < oo, r =£ 0) then one usually writes M r for M^.
For r > 0 we define the mean Mr for all non-negative sequences: if
a =

Note that if px = • — = pn = 1/n then Mx is the usual arithmetic mean


A, M2 is the quadratic mean and M_ x is the harmonic mean. As an
immediate consequence of Theorem 4, we shall see that Mr is a continu-
ous monotone increasing function of r.
In fact, Mr(a) has a natural extension from (-°°,0) U(0,oo) to the
whole of the extended real line [-00,00] such that Mr(a) is a continuous
monotone increasing function. To be precise, put
Chapter 1: Basic inequalities

n
MJ(a) = max ai9 M^a) = min ai9 M0(a) = J] a?1.

Thus M0(a) is the weighted geometric mean of the at. It is easily


checked that we have Mr(a) = {M-r{a~l)}~1 for all r (-<» ^ r ^ 00).

Theorem 5. Let a = (a1,...,an) be a sequence of positive numbers, not


all equal. Then Mr(a) is a continuous and strictly increasing function of
r on the extended real line -0° ^ r ^ 00.

Proof. It is clear that Mr(a) is continuous on (-o°,0)U(0,o°). To


show that it is strictly increasing on this set, let us fix r and s, with
-00 < r < s < 00, r ^ O and 5 ^ 0 . If 0 < r then Jr is an increasing
function of t > 0, and fr^ is a concave function, and if r < 0 then £r is
decreasing and f^s is convex. Hence, by Theorem 4, we have
Mr{a) <Ms(a).
Let us write A(a) and G(fl) for the weighted arithmetic and geometric
means of a = (a1,...,an), i.e. set

A(a) = Mx{a) = £ A-*,- and G(n) = Afo(fl) = ft */"••


1=1 1=1

To complete the proof of the theorem, all we have to do is to show


that

MM = lim Mr(a), M^a) = lim Mr(a), G(a) = limAfr(fl).


^ 0

The proofs of the first two assertions are straightforward. Indeed, let
1 ^ m ^ n be such that am — M^{a). Then for r > 0 we have

Mr(a) ^ \pm am) ' — pm am;

so lim infr^oo Mr(a) ^ am ~ Moo(ci). Since Mr(a) ^ MaXa) for every r,


we have l i m ^ ^ Mr(fl) = MJ^a), as required. Also,

Af_oo(fl) = {MQ0(a~ )}~ = {limM r (a~ )}~ = lim Mr(a)-


y—»oo f—> — oo

The final assertion, G(fl) = lim r ^ 0 Mr(a), requires a little care. In


keeping with our conventions, for — oo < r < oo (r ^ 0) let us write
ar = ( « [ , . . . , < ) . Then, clearly,

Also, it is immediate that


Chapter 1: Basic inequalities

d rlogfl,
lim -(af-1) = log a,
7rQ
=

and so

lim - {A (a^-l } =logG(a) (6)


r->0 r
Since
logt^ t-1
for every t>0, if r > 0 then
1 1
]noGi 6 V
<
r r r
Letting r —> 0, we see from (6) that the right-hand side tends to log G(a)
and so

lim logAfr(tf) = lim - log Aia^} = logG(a),


r—»0+ r—>0+ T

implying
lim Mr{a) = G(a).

Finally,

lim Mr{a) = Hm {M^a" 1 )}" 1 = GCa"1)"1 = G(a). •


r->0- 0+

The most frequently used inequalities in functional analysis are


due to Holder, Minkowski, Cauchy and Schwarz. Recall that a
hermitian form on a complex vector space V is a function <p : Vx F -> C
such that (p(\x + /xy,z) = \(p(x,z)+fjL(p(y,z) and <p(y,x) = <p(x,y) for all
x,y,z EV and \,fji EC. (Thus <p(*,A>> + /iz) = A<P(JC,>')+/Z^(JC,Z).) A
hermitian form <p is said to be positive if cp(x, x) is a positive real number
foralljcE V(x + 0).
Let <p(-, -) be a positive hermitian form on a complex vector space V.
Then, given x,y E V, the value

is real and non-negative for all A E C. For x ^ 0, setting A =


"~<p(x9y)/(p(x9x)9 we find that

and the same inequality holds, trivially, for x = 0 as well. This is the
Chapter 1: Basic inequalities

Cauchy-Schwarz inequality In particular, as

<£>(*,)>) =

is a positive hermitian form onC",

and so
n

V=l •n i= l
1*1*
1/2

(i. ;r-
w>
n 1/2 \l/2/ „ \l/2
2
S Ivl2
V=l
(7)
Our next aim is to prove an extension of (7), namely Holder's ine-
quality.
Theorem 6. Suppose

p,q > 1 and - + - = 1.

Then for complex numbers a1,...,an,b1,...,bn we have


q
2 |«*|') "(I \bk\<) (8)
.*=i / U=i /
with equality iff all ak are 0 or \bk\q = t\ak\p and akbk = el6\akbk\ for
all k and some t and 6.
Proof. Given non-negative reals a and b, set xx = ap, x2 = bq,
Pi = V/7 andp 2 = 1/?• Then, by Theorem 3,

afc = J C M 2 ^ plXl +p2x2 =J + J> (9)


with equality iff ap = bq.
Holder's inequality is a short step away from here. Indeed, if
n
1
t=l
then by homogeneity we may assume that

But then, by (9),


10 Chapter 1: Basic inequalities

Kl" \bt\i
«1 |

Furthermore, if equality holds then

\ak\P=\bk\" and =2
implying akbk = el6\akbk\. Conversely, it is immediate that under these
conditions we have equality in (8). D

Note that if Mr denotes the rth mean with weights pt — n~l


(i = l , . . . , w ) a n d f o r a = ( a l , . . . , a n ) a n d b = ( b l 9 . . . 9 b n ) w e p u t
ab = ( a 1 b l 9 . . . 9 a n b n ) 9 \a\ = (\ax |,..., \an |) a n d | 6 | = ( 1 ^ 1 , . . . , \bn | ) ,
then Holder's inequality states that if p~1 + q~1 = 1 with p,q > 1, then
Mx{\ab\)^Mp(\a\)Mq{\b\).
A minor change in the second half of the proof implies that (8) can be
extended to an inequality concerning the means M1? Mp and Mq with
arbitrary weights (see Exercise 8).
The numbers p and q appearing in Holder's inequality are said to be
conjugate exponents (or conjugate indices). It is worth remembering that
the condition n~x + q~l = 1 is the same as
(p-l)q = p or (q — l)p = q.
Note that 2 is the only exponent which is its own conjugate. As we
remarked earlier, the special case p = q = 2 of Holder's inequality is
called the Cauchy-Schwarz inequality.
In fact, one calls 1 and oo conjugate exponents as well. Holder's ine-
quality is essentially trivial for the pair (I, 00 ):
M1(\ab\)^M1(\a\)MO0(\b\)9
with equality iff there is a 6 such that \bk\ = Afoo(|6|) and akbk =
cl6\akbk\ whenever ak ^ 0.
The next result, MinkowskVs inequality, is also of fundamental
importance: in chapter 2 we shall use it to define the classical lp spaces.
Theorem 7. Suppose 1 ^ p < oo and al9...,an,bi,...,bn are complex
numbers. Then
i \VP / n \l/P
Chapter 1: Basic inequalities 11

with equality iff one of the following holds:


(i) all ak are 0;
(ii) bk = tak for all k and some t ^ 0;
(iii) p = 1 and, for each k, either ak = 0 or bk = tkak for some tk ^ 0.
Proof. The assertion is obvious if p = 1 so let us suppose that
1 < p < oo? not all ak are 0 and not all bk are 0. Let q be the conju-
gate of p: p~l + q~l — 1. Note that

£ k+ft*l*« £ k+**lp"1kl + £ k+W 1 !**!-


fc=l k=l k=\
Applying (8), Holder's inequality, to the two sums on the right-hand
side with exponents q and p, we find that

l/q
l
/«(/ n \UP I n \Vp]

I t ) i?,'»') )
Dividing both sides by

we obtain (8). The case of equality follows from that in Holder's ine-
quality. •

Minkowski's inequality is also essentially trivial for p = oo, i.e. for the
Moo mean:

with equality iff there is an index k such that \ak\ = Afoo(|a|),


\bk\ =MJ\b\)ui&\ak + bk\ = \ak\ + \bk\.
The last two theorems are easily carried over from sequences to
integrable functions, either by rewriting the proofs, almost word for
word, or by approximating the functions by suitable step functions.
Readers unfamiliar with Lebesgue measure will lose nothing if they take
/ and g to be piece wise continuous functions on [0,1].
12 Chapter 1: Basic inequalities

Theorem 8. (Holder's inequality for functions) Let p and q be conju-


gate indices and let / and g be measurable complex-valued functions on
a measure space (X, ^,/JL) such that \f\p and \g\q are integrable. Then
fg is integrable and
/ r \\/ni r \\la

u
Theorem 9. (Minkowski's inequality for functions) Let 1 ^ p < °° and
let / and g be measurable complex-valued functions on a measure space
p p p
(X,&,IJL) such that \f\ and \g\ are integrable. Then \f+g\ is integr-
able and

*,)• •
Exercises
All analysts spend half their time hunting through the
literature for inequalities which they want to use but can-
not prove.
Harald Bohr

f ]
1. Let An = JC = (Xi)1: 2 JC,: = n and JC, ^ 0 for every i\ C R".
(i) Show that g(x) = I]" = 1 *i is bounded on i4n and attains its
supremum at some point z = (z,)i E >!„.
(ii) Suppose that x E. A and JCX = minjcf < x2 = maxjc,-. Set y1 =
yi ~ 2(^1+^2) an<^ yi = xi f° r 3 ^ 1 ^ n. Show that y =
(yi)1 e >4n and g(y) > g(x). Deduce that zt = 1 for all 1.
(iii) Deduce the AM-GM inequality.
2. Show that if </f: (a,b) —> (c,d) and <p : (c, d) —> R are convex func-
tions and <p is increasing then ((p°i//)(x) = <p(i//(x)) is convex.
3. Suppose that / : (#,&) -» (0,oo) j s such that log/ is convex. Prove
that / is convex.
4. L e t / : (a,b) -> (c,6) and 9 : (c,6) -> R be such that <p and <p~lof
are convex. Show that / is convex.
5. Let {fy : y G F} be a family of convex functions on (a, 6) such that
f(x) = s u p r e r /7(JC) < 00 for every x E (a, 6). Show that f(x) is
also convex.
6. Suppose that / : (0,1) —» R is an infinitely differentiable strictly
convex function. Is it true that/"(;t) > 0 for every x E (0,1)?
Chapter 1: Basic inequalities 13

7. Let p,q > 1 be conjugate exponents. By considering the areas of


the domains
Di = {(x,y): 0 ^ x ^ a and 0 ^ y ^ JC^7"1}

and

prove inequality (8) and hence deduce Holder's theorem. [Note


that (p-l)(q-l) = 1, soifjy = xp~l then x = yq~\]
8. Prove the following form of Holder's theorem for the means
n
/ \1/r
MM) = A*r((«i)i) = 2 Pitf) (r > 0).
V=i /
If p,q > 1 are conjugate exponents, that is p~l + q~l = 1, and
{ax,...,an) and {bi,...,bn) are complex sequences then
n
2 Piatbi

with equality as in Theorem 6.


9. Deduce from the inequality in the previous exercise that if
a = (ai,...,an) is a positive sequence, that is, a sequence of posi-
tive reals, not all equal, then Mrr{a) is a strictly log-convex function
of r, i.e. rlog Mr(a) is a strictly convex function of r.
10. Show that for a fixed positive sequence a, logMx/r{a) is a mono-
tone decreasing convex function of r > 0, that is, if r~l =
1
+ 4~1) then
MM) ^ {Mp(a)Mq(a)}1/2.
11. Show that MX/M) is a monotone decreasing convex function of r.
12. Show that if 0 < < ? < / ? < r then
(r~p)/(r-q) I n \{p-q)/{r-q)
2 )
for all positive reals a1,a29...,an,/i1,ii2,...9tLn.
13. Deduce from the previous exercise that if 0 < q < p < r and
f(x) > 0 is continuous on (a,b) (or just measurable) then
rb ( rb \(r-p)/(r-q) / b \{p-q)/(r-q)
14 Chapter 1: Basic inequalities

14. Show that if p,q,r > 0 are such that p~x + q~x = r~l then

Mr(ab) *£ Mp{a)Mq{b)
for all positive sequences a = (aly...,an) andfe= (b1,...,bn).
Prove also that if p,q,r,s > 0 are such that p ~ + q~ + r~ = s'1
1 1 1

then
Ms{abc) ^Mp{a)Mq(b)Mr(c)
for all non-negative sequences a, b and c.
State and prove the analogous inequality for k sequences.
15. Let a and b be positive sequences. Show that if r < 0 < s < t and
r~l+s~l = t'1 then
Mt(ab) ^ Mr(a)Ms(b).
Deduce that if 0 < r < 1 then
Mr(a + b) ^ Mr(a) + Mr(b).
16. Let <Pp,q(x,y) = xl/pyx/q {x,y ^ 0). Show that <Pp q is concave iff
p,q ^ 1 and (p-l)(q-l) ^ 1 and that it is convex iff p,q ^ 1 and

17. Deduce from the result in the previous exercise that Mi(ab) and
Mp(a)Mq(b) are comparable if (p-l)(q — 1) 5* 1. To be precise, if
p,q > 1 and (/?-l)(<?-l) ^ 1 then

M1(ab) ^ Mp(a)Mq(b)
and if p,q < 1 and (/?-l)(^-l) ^ 1 then

18. Mr(a + x)-x = Mr(a1+x,a2 + x,...,an+x)-x is a decreasing func-


tion of x ^ 0 and tends to M^a) as JC -> ». Show also that if
r ^ 1 then Mr(a + jc)-jc is an increasing function of x ^ 0 and it
also tends to Mi(a) as JC —> oo.
19. Prove that if r > 1, « > 0 and a + 1 then

ar-\ >r{a-\)
and

r
Deduce that if a, & > 0 and a =fi b then
Chapter 1: Basic inequalities 15

rbr~\a-b) < ar-br < rar~\a-b)


if r < 0 or r > 1, and that the reverse inequalities hold if
0 < r < 1.
20. Prove Chebyshev's inequality: if r > 0, 0 < ax ^ a2 ^ • • • ^ «„
and 0 < bx ^ fe2 ^ • • • ^ bn then
Mr(a)Mr(b) <Mr{ab)
unless all the at or all the bt are equal. Prove also that the ine-
quality is reversed if a — (a^i is monotone increasing and b =
(bj)" is monotone decreasing. [HINT: Note first that it suffices to
prove the result for r = 1. For r = 1 the difference of the two
sides is 2*,- 2 tjajbj ~ 2'*<*/ 2*/fy•]
21. For fl1,...,fln>0(fl1^a2) ' e t ^ be the arithmetic mean of the
(2) products of the form n* =1 % (1 ^ h < '" < k ^ n)- s h o w
that the sequence TT0, TT\ ,..., 7rw is strictly log-concave:

for all fc (1 ^ A: ^ A Z - 1 ) . [HINT: A p p l y i n d u c t i o n o n n. For


n = 2 t h e assertion is just t h e A M - G M inequality for t w o t e r m s .
N o w let n ^ 3 a n d d e n o t e by ir'r t h e a p p r o p r i a t e average for
a1,...,an_1. Check that
—r

and deduce that for


AZ2(7Tr-l7Tr + 1 - 7 r r 2 ) = X + 5 ^ + C ^ 2

we h a v e

v4 < - T T ; 2 , ^ < 27r;_ 1 7r;, C < -Tr;2^.]

22. D e d u c e from t h e inequality in t h e previous exercise t h e following


considerable extension of t h e A M - G M inequality, already k n o w n
to N e w t o n : if ax,..., an a r e positive, n ^ 2 a n d ax ^ a2 t h e n

23. Let / : IR+ -» R be a strictly convex function with /(0) ^ 0. Prove


that if a1,...,an ^ 0 and at least two at are non-zero then
16 Chapter 1: Basic inequalities

24. L e t / : (0,<») —> (0,o°) be a monotone increasing function such that


f(x)/x2 is monotone decreasing. For a,b > 0 set

and g(a,b) = a2

Prove that if ai9bi > 0 then

(*)
25. Prove that if f,g: (0,oo)x(0,oo) -* (0,oo) satisfy (*) for all ai9bt > 0
then they are of the form given in the previous exercise.
26. Show that

for all real ai9bi with a2 + b2 > 0.


27. Let bi,b2,...,bn be a rearrangement of the positive numbers
ax, <z 2 ,...,fl/i.Prove that

28. Let/(;c) ^ 0 be a convex function. Prove that

/2djc^|max/ /ck.
Jo Jo
Show that f is best possible.
29. L e t / : [0,#] —> R be a continuously differentiable function satisfy-
ing /(0) = 0. Prove the following inequality due to G. H. Hardy:

[HINT: Note that

and so
Chapter 1: Basic inequalities 17

30. Show that Theorem 4 characterizes comparable means, i.e. if


M9(a) ^ M$(a) for all a = (a;)" (at > 0) then cpi//'1 is concave if cp
is increasing and convex if <p is decreasing.
31. In a paper the authors claimed that if xk > 0 for k = 1,2, . . ., n then
V > *! (2x2 - xx) (3x3 - 2x 2 ). . . (nxn - (n - 1) xB_ J.
Show that this is indeed true if 2x2 > xl9 3x 3 > 2 x 2 , . . . , nxn > (n — 1)
xn_ 1? and equality holds if and only if Xj = ... = xn. Show also that the
inequality need not hold if any of the n — 1 inequalities kxk > (k — 1)
xfc_x fails.

Notes

The foundation of the theory of convex functions is due to J. L. W. V.


Jensen, Sur les fonctions convexes et les inegalites entre les valeurs moy-
ennes, Acta Mathematica, 30 (1906), 175-93. Much of this chapter is
based on the famous book of G. H. Hardy, J. E. Little wood and G.
Polya, Inequalities, Cambridge University Press, First edition 1934,
Second edition 1952, reprinted 1978, xii + 324 pp. This classic is still in
print, and although its notation is slightly old-fashioned, it is well worth
reading.
Other good books on inequalities are D. S. Mitrinovic, Analytic Ine-
qualities, Springer-Verlag, Berlin and New York, 1970, xii + 400 pp.,
and A. W. Marshall and I. Olkin, Inequalities: Theory of Majorization
and Its Applications, Academic Press, New York, 1979, xx + 569 pp.
2. NORMED SPACES AND BOUNDED
LINEAR OPERATORS

In this long chapter we shall introduce the main objects studied in linear
analysis: normed spaces and linear operators. Many of the normed
spaces encountered in practice are spaces of functions (in particular,
functions on M, i.e. sequences), and the operators are often defined in
terms of derivatives and integrals, but we shall concentrate on the
notions defined in abstract terms.
As so often happens when starting a new area in mathematics, the
ratio of theorems to definitions is rather low in this chapter. However,
the reader familiar with elementary linear algebra and the rudiments of
the theory of metric spaces is unlikely to find it heavy going because the
concepts to be introduced here are only slight extensions of various con-
cepts arising in those areas. Moreover, the relatively barren patch is
rather small: as we shall see, even the basic definitions lead to fascinat-
ing questions.
A normed space is a pair (V, || • ||), where V is a vector space over R
or C and || • || is a function from V to IR+ = {r E R: r ^ 0} satisfying
(i) ||jt:|| = 0 iff JC = 0;
(ii) ||AJC|| = |A11|*|| for all x E V and scalar A;
(iii) \\x + y\\ ^ W + l b l l f o r a l l j c ^ E F .
We call ||JC|| the norm of the vector x: it is the natural generalization
of the length of a vector in the Euclidean spaces Rn or Cn. Condition
(iii) is the triangle inequality: in a triangle a side is no longer than the
sum of the lengths of the other two sides.
In most cases the scalar field may be taken to be either R or C, even
when, for the sake of simplicity, we specify one or the other. If we
want to emphasize that the ground field is C, say, then we write 'com-
plex normed space', 'complex lp space', 'complex Banach space', etc.

18
Chapter 2: Normed spaces and bounded linear operators 19

Furthermore, unless there is some danger of confusion, we shall identify


a normed space X = (V, ||-||) with its underlying vector space V, and
call the vectors in V the points or vectors of X. Thus x EL X means that
x is a point of X, i.e. a vector in V. We also say that || • || is a norm on
X.
Every normed space is a metric space and so a topological space, and
we shall often make use of some basic results of general topology.
Although this book is aimed at the reader who has encountered metric
spaces and topological spaces before, we shall review some of the
basic concepts of general topology. A metric space is a pair (X,d),
where X is a set and d is a function from XxX into U+ = [0,°°) such
that (i) d(x,y) = 0 iff x = y, (ii) d(x,y) = d(y,x) for all x,y E X and
(iii) d(x,z) ^ d(x,y) + d(y,z) for all x,y,z E X. We call d(x9y) the dis-
tance between x and y; the function d is a metric on X. Condition (iii)
is again the triangle inequality.
A topology r on a set X is a collection of subsets of X such that (i)
0 E r and X E r, (ii) r is closed under arbitrary unions: if Uy E r for
y E F then (J e r Uy E T, and (iii) r is closed under finite intersec-
tion: if Ul9...9 Un E T then fl" = 1 UtET. The elements of the collec-
tion r are said to be open (in the topology T). A topological space is a
pair (Jf, T) , where X is a set and r is a topology on X. If it is clear that
the topology we take is r then we do not mention r explicitly and we
call X a topological space.
If Y is a subset (also called a subspace) of a topological space (X, r)
then {YOU: U E T} is a topology on 7, called the subspace topology or
the topology induced by r. In most cases every subset Y is considered
to be endowed with the subspace topology.
Given a topological space X, a set N C X is said to be a neighbour-
hood of a point JC E ^ if there is an open set U such that x E U C N.
A subset of X is closed if its complement is open. Since the intersection
of a collection of closed sets is closed, every subset A of X is contained
in a unique minimal closed set A = {x E X: every neighbourhood of x
meets A}, called the closure of >1.
It is often convenient to specify a topology by giving a basis for it.
Given a topological space (X, r), a basis for r is a collection er of subsets
of X such that a C r and every set in r is a union of sets from a.
Clearly, if a C ^(X), i.e. o- is a family of subsets of X, then o- is a basis
for a topology iff
(i) every point of X is in some element of cr\
(ii) if Bi,B2 E a then B1DB2 is a union of some sets from a.
20 Chapter 2: Normed spaces and bounded linear operators

A neighbourhood base at a point x0 is a collection v of neighourhoods


of Jt0 such that every neighbourhood of x0 contains a member of v.
There are numerous ways of constructing new topological spaces from
old ones; let us mention here the possibility of taking products, to be
studied in some detail in Chapter 8. Let (X,a) and (Y,T) be topological
spaces. The product topology on XxY = {(x,y): x E X, y E Y} is the
topology with basis {UxV: U E a, V E T}. Thus a set W C J x y is
open iff for every (*,)>) E W there are open sets U C X and V C 7
such that (x,y) E UxV C W.
If d is a metric on ^ then the open balls
D(x,r) ={yEX: d(x,y) < r} (x E X, r > 0)
form a basis for a topology. This topology is said to be defined or
induced by the metric d; we also call it the topology of the metric space.
Not every topology is induced by a metric; for example,
T = {[/CR: U = 0 or the complement U\U of U is countable}
is a topology on IR and it is easily seen that it is not induced by any
metric.
Given topological spaces (X1,r1) and (X2,T2), a map f:X1^X2 is
said to be continuous if f~l{U) E T1 for every U E r 2 , i.e. if the inverse
image of every open set is open.
A bijection / from Xx to X2 such that both / and f~l are continuous
is said to be a homeomorphism; furthermore, (XI,TI) and (X2,T2) are
said to be homeomorphic if there is a homeomorphism from Xx to A"2.
A sequence (xn)™ in a topological space (^,r) is said to be convergent
to a point JC0 E A", denoted xn —» JC0 or lim^-^oo *„ = JC0 , if for every
neighbourhood A^ of JC0 there is an n0 such that xn E N whenever
n^ n0. Writing S for the subspace {n~l: n = l,2,...}U{0} of R with
the Euclidean topology, we see that lim^^oo xn = x0 iff the map
/ : S—» X, given by fin'1) = xn and/(0) = xo is continuous.
The topology of a metric space is determined by its convergent
sequences. Indeed, a subset of a metric space is closed iff it contains
the limits of its convergent sequences.
If a and r are topologies on a set X and a C r then a is said to be
weaker (or coarser) than r, and T is said to be stronger {ox finer) than a.
Thus <7 is weaker than r iff the formal identity map (X, r) —> (X, o) is
continuous.
The topological spaces occuring in linear analysis are almost always
Hausdorff spaces, and we often consider compact Hausdorff spaces. A
Chapter 2: Normed spaces and bounded linear operators 21

topology T on a set X is a Hausdorff topology if for any two points


x,y E X there are disjoint open sets Ux and Uy such that x G Ux and
y E. Uy. A topological space (^, r) is compact if every open cover has a
finite subcover, i.e. if whenever X= (J £ r t/ y , where each t/y is an
open set, then X = [J Uy for some finite subset F of F. A subset A
of a topological space (X, r) is said to be compact if the topology on A
induced by r is compact. Every closed subset of a compact space is
compact, and in a compact Hausdorff space a set is compact iff it is
closed.
It is immediate that if AT is a compact space and f:K-*Ris continu-
ous then / is bounded and attains its supremum on K. Indeed, if we
had f(x) < s = sup{f(y): y E K} for every x E K then
U {xGK:f(x)<r}
r<s

would be an open cover of K without a finite subcover.


Every normed space X is a metric space with the induced metric
d(x,y) = \\x—y\\. Conversely, given a metric d on a vector space X,
setting ||JC|| = d(x,0) defines a norm on X iff d(x,y) = d(x + z,y + z) and
d{\x,\y) = \\\d(x,y) for all x,y,zE.X and scalar A. The induced
metric in turn, defines a topology on X, the norm topology. We shall
always freely consider a normed space as a metric space with the
induced metric and a topological space with the induced topology, and
we shall use the corresponding terminology.
Let X be a normed space. By a subspace of X we mean a linear sub-
space Y of the underlying vector space, endowed with the norm on X
(to be pedantic, with the restriction of the norm to Y). A subspace is
closed if it is closed in the norm topology. Given a set Z C X, the sub-
space spanned by Z is

linZ = | 2 hkzk : zk E Z, A^ scalar, n = 1,2,...


U
it is called the linear span of Z and it is the minimal subspace containing
Z.
A normed space is complete if it is complete as a metric space, i.e.
if every Cauchy sequence is convergent: if (xn)i C X is such that
d(xn,xm) = \\xn—xm\\ ^> 0 as min{Az,ra} —> & then (xn)™ converges to
some point x0 in X (i.e. d(xn,x0) = \\xn-x0\\ -* 0). A complete
normed space is called a Banach space. It is easily seen that a subset of
a complete metric space is complete iff it is closed; thus a subspace of a
Banach space is complete iff it is closed.
22 Chapter 2: Normed spaces and bounded linear operators

A metric space is separable if it contains a countable dense set, i.e. a


countable set whose closure is the whole space. A normed space is
separable if as a metric space it is separable. Most normed spaces we
shall consider are separable.
Given a normed space X, the unit sphere of X is
S(X) = {xEX: \\x\\ = 1}
and the (closed) unit ball is
B(X) = {x EX: \\x\\ ^ 1}.
More generally, the sphere of radius r about a point x0 (or centre x0) is
Sr(x0) = S(xo,r) ={xEX: \\x-xo\\ = r)
and the (closed) ball of radius r about x0 (or centre JC0) is
Br(x0) = B(xo,r) ={xEX: \\x-xo\\ ^ r}.
Occasionally we shall need the open ball of radius r and centre x0,
Dr(x0) = D(xo,r) ={xEX: \\x-xo\\ < r}.
Note that the sets x + tB(X) (t > 0) form a neighbourhood base at the
point x.
The definition of the norm implies that Br(x0) is closed, Dr(x0) is
open, and for r > 0 the interior of Br(x0) is lntBr(x0) = Dr(x0). The
closure of Dr(x0) is Dr(xo) = Br(x0). Furthermore, the boundary
dBr(x0) of Br(x0) is the sphere Sr(x0).
The definition also implies that if X is a normed space then the map
XxX^> X given by (x,y) >-> x + y is uniformly continuous. Similarly,
the map RxX-* X (or CxX-* X) given by (A,JC) •-» AJC is continuous.
Note that the norm function || • || from X to R+ is also continuous.
Let us give a host of examples of normed spaces. Most of these are
important spaces, while some others are presented only to illustrate the
definitions. The vector spaces we take are vector spaces of sequences or
functions with pointwise addition and multiplication: if x = (x^ and
y = (yt)i then kx + ny = (A^ + /x^-)5°; if / = f(t) and g = g(t) are func-
tions then (A/+/xg)(r) = \f(t)+iig(t).
In the examples below, and throughout the book, various sets are
often assumed to be non-empty when the definitions would not make
sense otherwise. Thus S =£ 0 in (ii), T ^ 0 in (iii), and so on.
Chapter 2: Normed spaces and bounded linear operators 23

Examples 1. (i) The n-dimensional Euclidean space: the vector space is


Un or Cn and the norm is

IWI = ( 2
V=i
where x = (xi,...9xn). The former is a real Euclidean space, the latter
is a complex one.
(ii) Let S be any set and let 9b(S) be the vector space of all bounded
scalar-valued functions on S. F o r / E 3^(S) let

This norm is the uniform or supremum norm.


(iii) Let L be a topological space, let X — C(L) be the vector space
of all bounded continuous functions on L and set, as in example (ii),
11/11= SUpJ/(*)|.
(iv) This is a special, but very important, case of the previous example.
Let K be a compact Hausdorff space and let C(K) be the space of con-
tinuous functions on K, with the supremum norm
||/|| = 11/11. = sup{|/(*) | : * £ * } .
Since K is compact, \f(x)\ is bounded on K and attains its supremum.
(v) Let X be Un or Cn and set
n
IMIi = 2 I*,-1.
i=i

This is the space /"; the norm is the lt-norm.


Also,
|JC||OO = max I*, |

is a norm, the /oo-norm; the space it gives is /£.


(vi) Let 1 ^ p < oo. For x = (x1,...,xn) E Un (or C") put

This defines the space l£ (real or complex); the norm is the lp~norm.
The notation is consistent with that in example (v) for /" and also, in a
natural way, with that for l£ (see Exercise 6). Note that 1% is exactly
the n-dimensional Euclidean space.
24 Chapter 2: Normed spaces and bounded linear operators

(vii) Let X consist of all continuous real-valued functions f(t) on R


that vanish outside a finite interval, and put

Hi = [ I/WI d*.
J — oo

(viii) Let X consist of all continuous complex-valued functions on


[0,1] and for / E X put
v l/2
2
\\fh=(fo m A .
(ix) For 1 ^ p < 00 the space lp consists of all scalar sequences
x = (x1, x2,...) for which

The norm of an element x E lp is

The space lx consists of all bounded scalar sequences with


||x||oo = sup |x,-|

and c0 is the space of all scalar sequences tending to 0, with the same
norm ||*||<x. As remarked earlier, we have both real and complex forms
of these spaces.
(x) For 1 =s$ p < 00 the space Lp(0,1) consists of those Lebesgue
measurable functions on [0,1] for which

Note that L^O, 1) is precisely the space of integrable functions.


Strictly speaking, a point of Lp(0,1) is an equivalence class of func-
tions, two functions being equivalent if they agree almost everywhere.
Putting it another way, f\ and f2 are equivalent (and so are considered
to be identical) if | | / i - / 2 | | p = 0.
(xi) By analogy with the previous examples, the astute reader will
guess that L^O, 1) consists of all essentially bounded Lebesgue measur-
able functions on [0,1], i.e. those functions / for which
Chapter 2: Normed spaces and bounded linear operators 25

H/IIco = esssup|/(0| <°°


Recall that the essential supremum esssup|/(f)| is defined as
inf{||/|5||oo: 5 C [0,1] and [0,1]\5 has measure 0}
= infjfl : the set {/: \f(t)\ > a} has measure 0},
where f\S is the restriction of f to S and so

Once again, strictly speaking L^O, 1) consists of equivalence classes of


functions bounded on [0,1].
(xii) Lp(R), Loo(IR), Lp(a,oo), Loo(tf,o°), etc., are defined similarly to
the spaces in (x) and (xi). Of these, once again, the Hilbert spaces
L2(R), L2(fl,°°), etc., are the most important.
(xiii) Let C (AI) (0, 1) consist of the functions f(t) on (0,1) having n con-
tinuous and bounded derivatives, with norm

\fk\t)\:O<t<\\
=o J
(xiv) Let X consist of all polynomials

nt) = i cktk
k=0
of degree at most n, with norm

11/11= 2 (*+i)kt|.
k=0
(xv) Let X be the space of bounded continuous functions on (—1,1)
which are differentiable at 0, with norm

(xvi) Let X consist of all finite trigonometric polynomials

fit) = £ ckek\ (n = 1,2,...)


k = -n
with norm

111
/1 = ( 2 k*
26 Chapter 2: Normed spaces and bounded linear operators

(xvii) Let V be a real vector space with basis (e1,e2), and for x E V
define

| + |ft| + |c|:fl,ft,c6 R and x = ae

(xviii) Let V be as in (xvii) and set

Ml = inff|fl| + |6| + |c| + |d|:fl, b, c, dE R


A ,, c(e ^e )-^d(e -e2)}
and x = aex + be2 + — 1—27 - — £ 1 — [.
y2 J
(xix) Let V be an n-dimensional real vector space and let
v1, v2,..., vm be vectors spanning V. For JC E V define

Ml = inf 2 kJ : Q E R and x = 2 Q

Let us prove that the examples above are indeed normed spaces, i.e.
that the underlying spaces are vector spaces and the functions || • || satisfy
conditions (i)-(iii). It is obvious that conditions (i) and (ii) are satisfied
in each example so we have to check only (iii), the triangle inequality.
Then it will also be clear that the underlying space is a vector space.
Furthermore, the triangle inequality is obvious in examples (ii)-(v),
(vii), (xi) and (xii)-(xv). In the rest, with the exception of the last
three examples, the triangle inequality is precisely Minkowski's inequal-
ity in one of its many guises. Thus Theorem 1.7 (Minkowski's inequal-
ity for sequences) is just the triangle inequality in l£\

P^\\x\\p + \\y\\P (i)


whenever x,y E l£\ this clearly implies the analogous inequality in lp\
Theorem 1.9 (Minkowski's inequality for functions) is precisely the tri-
angle inequality in L p (0,1):

l / + * l , « l / l , + ll*llp (2)
whenever f,g E L p (0,1), etc. In turn, the triangle inequality implies
that the underlying spaces are indeed vector spaces.
Examples (xvii)-(xix) are rather similar, with (xix) being the most
general case; we leave the proof to the reader (Exercise 7). •

The spaces lp (1 ^ p ^ °°) are the simplest classical sequence spaces,


and the spaces C(K) and Lp(0,1) (1 ^ p ^ oo) are the simplest classical
Chapter 2: Normed spaces and bounded linear operators 27

function spaces. These spaces have been extensively studied for over
eighty years: much is known about them but, in spite of all this atten-
tion, many important questions concerning them are waiting to be set-
tled. In many ways, the most pleasant and most important of all
infinite-dimensional Banach spaces is / 2 , the space of square-summable
sequences. This space l2 is the canonical example of a Hilbert space.
Similarly, of the finite-dimensional normed spaces, the space 1% is cen-
tral: this is the n-dimensional Euclidean space.
The spaces in Examples (i)-(vi), (ix)-(xiv) and (xvii)-(xix) are com-
plete, the others are incomplete. Some of these are easily seen, we
shall see the others later.
Let us remark that if X is a normed space then its completion X as a
metric space has a natural vector space structure and a natural norm.
Thus every normed space is a dense subspace of a Banach space. We
shall expand on this later in this chapter.
Having mentioned the more concise formulations of Minkowski's ine-
quality (inequalities (1) and (2)), let us draw attention to the analogous
formulations of Holder's inequality. Suppose p and q are conjugate
indices, with p = 1 and q = o° permitted. If x = (xk)™ E lp and
y = totf e /, then

2 1*^*1 *II4,IMI,. (3)


k=l
Similarly, if / E Lp(0,1) and g E Lq(0,1) then fg is integrable and

\fg\dx<\\f\\P\\g\\q- (4)

It is easy to describe the general form of a norm on a vector space V


in terms of its open unit ball. Let X = (V, || • ||) be a normed space and
let D = {x E V: \\x\\ < 1} be the open unit ball. Clearly || • || is deter-
mined by D: if x ± 0 then ||*|| = inf{t: t > 0, x E tD}.
The set D has the following properties:
(i) ifx,y E D and |A| + |/i| ^ 1 then AJC + py E D;
(ii) if x E D then x + eD C D for some e = e(x) > 0;
(iii) for x E V (x ^ 0) there are non-zero scalars A and /JL such that
AJC E D and fix £ D.
A set D satisfying (i) is said to be absolutely convex: this property is a
consequence of the triangle inequality. Property (ii) follows from the
fact that D is an open ball, while (iii) holds since \\x\\ < °° for every x
and \\y\\ = Oiffy = 0.
28 Chapter 2: Normed spaces and bounded linear operators

Conversely, if D C V satisfies (i)-(iii) then


q(x) = inf{t: t > 0, x E tD}
defines a norm on V, and in this norm D is the open unit ball. The
function q(x) is the Minkowski functional determined by D. Note that
in conditions (i)-(iii) we do not assume any topology on V.
When studying normed spaces, it is often useful to adopt a geometric
point of view and examine the geometry, i.e. the 'shape', of the unit ball.
Much of the material presented in this book concerns linear function-
als and linear operators. Let X and Y be normed spaces over the same
ground field. A linear operator from X to Y is a linear map between the
underlying vector spaces, i.e. a map T: X-* Y such that
r(A1^1+A2JC2) = A1r(x1)+A27X*2)
for all Xi,x2 E X and scalars Ax and A2. The vector space of linear
operators from X to Y is denoted by i£{X, Y). The image of T is
Im T = {Tx : x E X} and the kernel of T is Ker T = {x : Tx = 0}.
Clearly Ker T is a subspace of X and Im T is a subspace of Y. Further-
more, T is a vector-space isomorphism iff Ker T = {0} and Im T = Y.
A linear operator T E !£(X, Y) is bounded if there is an N > 0 such
that
|| 7x|| ^tf||jt|| for all x EX.
We shall write 3h(X, Y) for the set of bounded linear operators from X
to Y; ^(X) = ^(X,X) is the set of bounded linear operators on X.
Clearly <&(X,Y) is a vector space. The operators in £(X, Y)\^(X, Y)
are said to be unbounded. A linear functional on X is a linear operator
from X into the scalar field. We write X' for the space of linear func-
tionals on X, and X* for the vector space of bounded linear functional
on X. It is often convenient to use the bracket notation for the value
of a functional on an element: for xEXf and fEXf we set <f,x) =
(x,f) = f(x). This is in keeping with the fact that the map XxX' -> R
(or C) given by (*,/) •-» f(x) is bilinear.
Theorem 2. Let X and Y be normed spaces and let T: X -> Y be a
linear operator. Then the following conditions are equivalent:
(i) T is continuous (as' a map of the topological space X into the topo-
logical space Y);
(ii) 7 is continuous at some point x0 E X;
(iii) r is bounded.
Chapter 2: Normed spaces and bounded linear operators 29

Proof. The implication (i) => (ii) is trivial.


(ii) => (iii). Suppose T is continuous at x0. Since Txo + B(Y) is a
neighbourhood of T(x0), there is a 8 > 0 such that
x = * 0 + y G xo + 8B(X) => Tx = Txo+Ty E
Hence \\y\\ ^ 8 implies ||7>|| ^ 1 and so ||rz|| ^ S ' l z l l for all z.
(iii) => (i). Suppose ||7JC|| ^ N\\x\\ for all x E X. Then if ||jc-.y|| <
e/N we have || Tx - Ty\\ < e. D

Two normed spaces X and Y are said to be isomorphic if there is a


linear map T:X—>Y which is a topological isomorphism (i.e. a
homeomorphism). We call X and 7 isometrically isomorphic if there is
a linear isometry from Jf to Y, i.e. if there is a bijective operator
r e » ( * , Y) such that T~l E 9J(y,J0 and ||7i|| = ||JC|| for all x E X.
Two norms, || - J|x and || -1| 2 , on the same vector space V are said to be
equivalent if they induce the same topology on V, i.e. if the formal iden-
tity map from Xx = (V, || • ||x) to X2 = (V, || • ||2) is a topological isomor-
phism. As an immediate consequence of Theorem 2, we see that if a
linear map is a topological isomorphism then both the map and its
inverse are bounded.
Corollary 3. Let X and Y be normed spaces and let T E £(X, Y). Then
Tis a topological isomorphism iff T E <3l(X, Y) and T~l E 9&(Y,AT).
Two norms, || • \\x and || • ||2, on the same vector space V are equivalent
iff there are constants c, d > 0 such that
cIMIi ^ IHI2 ^ 4*111
for all x EV. D

It is immediate that equivalence of norms is indeed an equivalence


relation, and Corollary 3 implies that if a normed space is complete then
it is also complete in every equivalent norm.
The equivalence of norms has an intuitive geometrical interpretation
in terms of the (open or closed) unit balls of the normed spaces. Let
H'IIJ and ||-||2 be norms on V, with closed unit balls Bx and B2> Then
|| • Id and || • ||2 are equivalent if and only if B2 C cBx and Bx C dB2 for
some constants c and d. Indeed, these relations hold if and only if
IMl! ^ c|M|2 and ||*||2 ^ d\\x\\x for all x E V.
It is clear that <&(X, Y) and X* are vector spaces: SB(AT, Y) is a sub-
space of £(X, Y) and X* is a subspace of S£(X, R). In fact, they are also
normed spaces with a natural norm. The operator norm or simply norm
30 Chapter 2: Normed spaces and bounded linear operators

on ^(X, Y) is given by
= ini{N > 0: ||7*|| ^ N||JC|| for all x <E X} = sup{||7;t||: \\x\\ ^ 1}.
Although this gives us the definition of the norm of a bounded linear
functional as well, let us spell it out: the norm offEX* is
Il/H = inf{N > 0: |/(x)|< ATM for all x(EX} = sup{|/(*)|: ||*|| < 1}.
Note that in these definitions the infimum is attained so
Iftfl « U71IIMI and | / ( * ) | « Il/H W for all x.
The terminology is justified by the following simple result.
Theorem 4. The function || • ||, defined above, is a norm on ^(X, Y). If
Y is complete then so is ^(X, Y). In particular, X* is a complete
normed space.
Proof. Only the second assertion needs proof. Let (7^)^° be a Cauchy
sequence in ^(X, Y). Then (Tnx)™ is a Cauchy sequence in Y for every
x E X and so there is a unique y E Y such that Tnx —> y. Set Tx = y.
To complete the proof, all we have to check is that T E 38(AVY) and
T —> T
Given JC1?JC2 E X and scalars Ax and A2, we have

= Ax lim

and so r ( J )
Furthermore, given 6 > 0, there is an n0 such that ||r n —rm|| < e if
n,m ^ n0. Then for JC E A' and m ^ n0 we have

= lim\\(Tn-Tm)x\\
«Umsup||rn-rOT|||W|
Therefore
Chapter 2: Normed spaces and bounded linear operators 31

and so T £ 9S(Z, Y). Finally, from the same inequality we find that
||r-rm||=£€. •

If we extend the operator norm to the whole of i£{X, Y) by putting


= sup{||7jt||: ||;t|| ^ 1} = oo for an unbounded operator, then
, Y) consists of the operators in !£(X, Y) having finite norm.
The Banach space X* is called the dual of X. For T G <&(X, Y) and
g E y* define a function T*g : X -> (R (or C) by
(T*g)(x) =g(Tx).
Then T*g is a linear functional on X and T* : Y* —> X* is easily seen to
be a linear map. Furthermore, T* is not only in !£(Y*,X*) but is, in
fact, a bounded linear operator since
\(T*g)(x)\ = \g(Tx)\ ^ \\g\\\\Tx\\ ^ \\g\\\\T\\\\x\\,
so that
llll II1I ||g|| and
The operator T* is the adjoint of T. In fact, as we shall see later
(Theorem 3.9), | | r || = ||r||.
The definition of the adjoint looks even more natural in the bracket
notation: T*g is the linear functional satisfying

Given maps T: X-> Y and 5 : Y-» Z, we can compose them: the


map ST: X-* Z is given by (ST)(x) = (SoT)(x) = S(T(x)). If 5 and T
are linear then so is ST; if they are bounded linear operators then so is
ST.
Theorem 5. Let X, Y and Z be normed spaces and let TEL 3& (X, Y),
S E 3$ (Y, Z). Then S77: X -* Z is bounded linear operator and
\\ST\\ ^ \\S\\\\T\\.

In particular, ^(X) = 3^(X,X) is closed under multiplication (i.e. under


the composition of operators).
Proof. Clearly
|| |||| |||||||| |||||||||W|. D
32 Chapter 2: Normed spaces and bounded linear operators

The last result shows that if X is a Banach space then ^(X) is a unital
Banach algebra: it is an algebra with a unit (identity element) which is
also a Banach space, such that the identity has norm 1 and the norm of
the product of two elements is at most the product of their norms.
Examples 6. (i) Define T: /2" - • /2m by putting
Tx = (jtj ,...,*/, 0,..., 0), where / = min{n, m).
Then T E »(/2",/2m) and ||r|| = 1.
(ii) If A is any endomorphism of W then A 6 <&(l£,l£) for all /? and
q (1 ^/?,tf ^°o).
(iii) Define S E Sft^) by

This is the right shift operator.


The left shift T E &(/p) is defined by
Tx = (x2,x3,...).

Clearly 5 is an injection but not a surjection, T is a surjection but not


an injection, ||5|| = ||r|| = 1, and TS = I (the identity operator) but
STj= I: K e r S r = {(JC 1? O,...) : xx E R}.
(iv) Let C{k) = C(k)(0,1) be the normed space of functions on (0,1)
with k continuous and bounded derivatives, as in Examples l(xiii). Let
D: C{n) -> Cin~1} be the differentiation operator: Df = f. Then
\\D\\ = 1.
(v) Let T: C{n) - • C(o) be the formal identity map Tf = f. Then
imi = i-
(vi) In order to define linear functionals on function spaces, let us
introduce the following notation. Given sequences x = (xk)™ and
y = (y/c)i> ^ t (x,y) = ^=1xkyk and let x-y be the sequence \xkyk)™.
Thus if x*y E /x then (x,y) is well defined.
Inequality (3), i.e. Holder's inequality for sequences, can now be res-
tated once again in the following concise form: if p and q are conjugate
indices, with 1 ^ p,q ^ a>, x E lp, and y E lq, then x-y E lx and
Ik-ylla^lWlplMI,. (5)
Inequality (5) implies that for y E lq the function fy(x) = (x,y) is a
bounded linear functional on lp and \\fy\\ ^ ||y||^. In fact, ||/y|| = ||y||9,
and for 1 ^ p < °° the correspondence y »-^> fy is a linear isometry which
identifies lq with /* (see Exercise 1 of the next chapter). The dual of /«,
Chapter 2: Normed spaces and bounded linear operators 33

contains l\ as a rather small subspace (/«> and /£ are not separable but /x
is); however, /x is the dual of c 0 , the closed subspace of /„ consisting of
sequences tending to 0 (again, see Exercise 1 of the next chapter),
(vii) Let p and q be conjugate indices and g E Lq(0,1). Define, for

•r fg.dx.

Then, by Holder's inequality (4), <pg is a bounded linear functional on


Lp(0,1) and H^ll ^ \\g\\r In fact, \\<pg\\ = \\g\\r and for 1 ^ p < oo the
correspondence g •-» <pg is a linear isometry which identifies Z^(0,1) with

(viii) Let C[0,1] be the space of continuous functions with the


supremum norm: ||/|| = \\f\\oo = max{|/(0|: 0 ^ t ^ 1}. Let 0 ^ t0 ^ 1
and define Sh: C[0,l] - • R by Sto(f) = f(to). Then 8,o is a bounded
linear functional of norm 1.
(ix) Let X be the subspace of C[0,1] consisting of differentiable func-
tions with continuous derivative. Then D : X —> C[0,1], defined by
Df = / ' , is an unbounded linear operator.
(x) Let V be the vector space of all scalar sequences x = (xk)™ with
finitely many non-zero terms. Set Jfp = (V, \\ • || p ), where

and

For 1 ^ r,5 ^ oo let 7 ^ : Xr -» Z, be the formal identity map: Trsx =


x. Then Tr 5 is a linear operator for all r and s. I f l ^ r ^ s ^ o o then
Tr>5 is bounded: in fact, ||rr>5|| = 1, but if r > s then Trs is unbounded
(see Exercise 27). •

As promised earlier in the chapter, let us say a few words about com-
pletions. Every metric space has a unique completion, and if the metric
is induced by a norm then this completion is a normed space. Before
examining the completion of a normed space, let us review briefly the
basic facts about the completion of a metric space.
A metric space X is said to be the completion of a metric space X if X
is complete and X is a dense subset of X. (A subset A of a topological
space T is dense if its closure is the whole of T.) The completion is
34 Chapter 2: Normed spaces and bounded linear operators

unique in the sense that if X is a dense subset of the complete metric


spaces Y and Z then there is a unique continuous map <p : Y —> Z whose
restriction to X is the identity, and this map is an isometry onto Z. This
is easily seen since if y E Y then y is the limit (in Y) of a sequence
(xn)i C X. Then (xn)i is a Cauchy sequence in Z and so it has a limit
z E Z. Since <p is required to be continuous, we must have <p(y) = z
and so <p, if it exists, is unique. On the other hand, if we define a map
<p : Y' —» Z by setting <p(y) = z, then this map is clearly an isometry
from Y onto Z.
The completion of a metric space is defined by taking equivalence
classes of Cauchy sequences. To be precise, given a metric space X, let
[X] be the set of Cauchy sequences of points of X:
[X] = {(*„)? :xnEX, lim sup d(xn,xm) = 0}.
Define an equivalence relation ~ on [X] by setting (xn)i ~ O^,)" if
d(xn,yn) -> 0, and let A" = [A^/~ be the collection of equivalence
classes with respect to ~ . For x,y E X set
rf(i,jO = limd(xn,yn),
n—»°°
where (jcn)f E JC and 0*,,)" E y. It is immediate that d : XxX -> [0, c») is
well defined, i.e. that limn-^ood(xn,yn) is independent of the representa-
tives chosen. Furthermore, if (xn)i E x, (yn)i E y and (zn)°? E z then
d(x, z) = lim d(xn ,zn) ^ lim sup{d(^, yn)
n—
= lim d(xn ,yn)+ lim d(>^n, zn)

Finally, d(i,i) = 0, and if d(x,y) = 0 then (xj™ ~ (y n )", so that


x = y.
The metric space (X, d) is complete. Indeed, let ( i ^ ) be a Cauchy
sequence in X and let (x^k))n=i E x{k) (k = 1,2,...). Let /i* be such that
d{x^k\x^)<2-k if n,m^nk.
Set JC^ = x^k). It is easily checked that 0^)5° E [X] and that £(*} tends to
the equivalence class of the sequence (xk)i.
Writing [x] for the equivalence class of the constant sequence
JC,JC, , wefindthat the map X—> X, given by x »-> [JC], is an isometry.
Thus, with a slight abuse of terminology, X can be considered to be a
subset of X. Clearly, Z is a dense subset of X, since if 005° is a
Chapter 2: Normed spaces and bounded linear operators 35

representative of x E X and d(xn,xm) < e whenever n,m ^ n0 then


d(x,xnQ) = d(x,[xno\) ^ 6.
If X is not only a metric space but also a normed space then its com-
pletion has a natural Banach-space structure.
Theorem 7. For every normed space X there is a Banach space X such
that X is dense in X. This space X is unique in the sense that if A" is a
dense subspace of a Banach space X then there is a unique continuous
map (p : X—* X such that the restriction of <p to X is the identity; furth-
ermore, this map <p is a linear isometry from X to X
Proof. Considering X as a metric space, with the metric induced by the
norm, i.e. with d(x,y) = \\x-y\\, let X be its completion. Given
x,y E X, let xn —• x and >>„ -> >>, where xn ,yn E X Then, for scalars A
and fi, the sequence (hxn + /JLyn)i is a Cauchy sequence. Now, if X is to
be given a normed-space structure such that the norm on X induces the
metric on X, then
lim (\xn + fiyn) = A lim x n + /x lim yn

Thus \x + fAy must be defined to be the limit of the Cauchy sequence


(hxn + /jLyn)™. It is easily checked that with this definition of addition
and scalar multiplication X becomes a vector space. Furthermore,
||*|| = d(x,0) = limn_,00||jcn|| is a norm on this vector space, turning X
into a Banach space.
The remaining assertions are clear. •
In fact, Theorem 7 is also an immediate consequence of Theorem 3.10. The
space X in this result is called the completion of the normed space X. In view of
Theorem 7, if X is a dense subspace of a Banach space Y then Y is often
regarded as the completion of X. For example, the space C[0,1] of continuous
functions on [0,1] is the completion of the space of piecewise linear functions
on [0, 1] with supremum norm
ll/lloo = sup{|/(x)| : 0 ^ x < 1} = max{|/(x)| : 0 ^ x ^ 1}.
The same space C[0,1] is also the completion of the space of polynomi-
als and of the space of infinitely differentiable functions. Similarly, for
1 ^ p < oo, Lp(0,1) is the completion of the space of polynomials with
the norm

\\f\\P = ^ \f(x)\p dx^j ",


and also of the space of piecewise linear functions with the same norm.
36 Chapter 2: Normed spaces and bounded linear operators

When working in Banach spaces, one often considers series. Given a


normed space X, a series

j ^ xn (xn E X)
is said to the convergent to x E X if
TV
xn —> x as iV

i.e. if
N
lim x- 2 *„ = 0.

We also say that x is the swra of the xn and write

x=

A series 2°°_i *n *s sa id to the absolutely convergent if Y'


oo. In a Banach space A^ every absolutely convergent series T^=n=1
1
xn is
convergent. Indeed, let yN = 2 ^ = 1 xn be the Mh partial sum. We
have to show that (y^)\ is a Cauchy sequence. Given e > 0, choose an
n0 such that 2 T = II** II < *. Then for n0 ^ N < M we have
M M
< e.
n=N+l
Hence ( ^ ) r is indeed a Cauchy sequence and so converges to some
x E X, so that JC = 2°° =1 *n •
In an incomplete space X the assertion above always fails for some
absolutely convergent series 2 " = 1 xn: this is often useful in checking
whether a space is complete or not.
Before showing this, let us mention a very useful trick when dealing
with Cauchy sequences in metric spaces: we may always assume that we
are dealing with a 'fast' Cauchy sequence. To be precise, given a Cau-
chy sequence (xn)i, a priori all we know is that there is a sequence
(rtfc)f, say nk = 2k, such that d(xn,xm) < \/k if n,m ^ nk. However,
we may always assume that our sequence is much 'better' than this:
d(xn,xm) < 2~n if m ^ n, or d(xn,xm) < 2~~2 if m ^ n, or whatever
suits us. Indeed, let f(n) > 0 be an arbitary function. Set e1 = /(I) and
choose an nx such that d(xn,xm) < e1 whenever n,m ^ nx. Then set
e2 = /(2) and choose an n2 > nx such that d(xn,xm) < e2 whenever
Chapter 2: Normed spaces and bounded linear operators 37

n,m ^ n2. Continuing in this way, we find a sequence nx < n2 <


such that d(xnk,xn{) < f(k) if / ^ k. Setting yk = xHk, we find that the
subsequence (yk)\ is such that d(yk,yi) < f(k) whenever k < I. Now if
we can show that (yk)\ converges to some limit y, then (xn)i tends to
the same limit y. Indeed, if d(xn,xm) < e for n,m^n0 then for
n ^ n0 we have
d(xn,y) = d(xn, limy^) ^ limsup d(xn,yk) ^ e,

since yk = xnk and nk ^ n0 if k is sufficiently large. Thus, in proving


that a Cauchy sequence (xn)™ in a metric space is convergent, we may
indeed assume that d(xn,xm) < 2~n, say, if m ^ n.
Theorem 8. A normed space is complete if and only if every absolutely
convergent series in it is convergent.
Proof. We have already seen that in a Banach space every absolutely
convergent series is convergent. Suppose then that in our space X every
absolutely convergent series is convergent. Our aim is to show that
every Cauchy sequence is convergent.
Let (xn)i be a Cauchy sequence in X. As we have just seen, we may
assume that d(xn,xm) = \\xn— xm\\ < 2~n for n ^ m. Set x0 = 0 and
yk = xk-xk.x (k ^ 1). Then xn = Tk=1 yk and \\yk\\ < 2~k+1. There-
fore the series 2Z=i ^ ^s absolutely convergent and has partial sums
xl9x2,... . By assumption, 2 ^ = 1 yk *s convergent, say to a vector x,
and so *„ —• *. •

Absolutely convergent series in Banach spaces have many properties


analogous to those of absolutely convergent series in U. For example, if
2°° xn is absolutely convergent to x and ni,n2,... is a permutation of
1,2,... then 2,™ xn is also absolutely convergent to x. However, as
we shall see in a moment, unlike in R, this property does not character-
ize absolutely convergent series in a Banach space.
The use of series often enables one to identify a Banach space with a
sequence space endowed with a particular norm. This identification is
made with the aid of a basis, provided that the space does have a basis.
Given a Banach space X, a sequence (e,-)" is said to be a (Schauder)
basis of X if every x E X can be represented in the form x = 2°°=1 ^iei
(A, scalar) and this representation is unique.
For 1 ^ p < oo the space lp has a basis; the so-called standard or
canonical basis (e,-)5°, where et• = (0,..., 0,1,0,...) = (S lf , 82i,...): the
38 Chapter 2: Normed spaces and bounded linear operators

/th term is 1 and the others are 0. It is easily seen that if x =


(JC 1? JC2,...) E lp then x = ^°°=1 xtet and this representation is unique
(see Exercise 11). In particular, 2°°=i xtei ls convergent (in lp) if and
only if 2JLi \xi\p < °°- This implies that in l2 every rearrangement of
2°° =1 en/n *s convergent (to the same sum), but the series is not abso-
lutely convergent.
To conclude this chapter, let us say a few words about defining new
spaces from old. We have already seen the simplest way: every (alge-
braic) subspace Y of a normed space X is a normed space, with the res-
triction of the norm of X to Y as the norm. If X is complete then Y is
complete if and only if it is closed. As the intersection of a family of
subspaces is again a subspace, for every set 5 C X there is a unique
smallest subspace containing S, called the linear span of S and denoted
by linS: it is the intersection of all subspaces containing S and also the
set of all (finite) linear combinations of elements of 5.
linS = PI {W: W is a subspace of X and S C W]

A,-^: si,...,sn E 5; n = 1,2,... |.


J
Similarly, the closed linear span of 5, denoted linS, is the unique smal-
lest closed subspace containing S; it is the intersection of all closed sub-
spaces containing 5, and is also the closure of lin 5, the linear span of S.
Let us turn now to quotient spaces. Given a vector space X and a
subspace Z, define an equivalence relation ~ on X by setting x ~ y if
x-y E Z. For x E X, let [x] be the equivalence class of x: putting it
another way, [x] = x + Z. Then X/~ = {[x]: x E X) is a vector space,
with vector-space operations induced by those on X: A[x]+/i[y] =
[\x + fiy]. Note that [x] = 0 iff x E Z. Now if X is a normed space
and Z is a closed subspace of A!' then we can define a norm || • ||0 on X/Z
by setting
HMIIo = inf{|M|: y ~ x} = inf{||x + z||: z £ Z}.
It is easily checked that ||-||0 is indeed a norm on X/Z: the homo-
geneity and the triangle inequality are obvious and ||[0]|| 0 = 0. All that
remains is to show that if ||[JC]|| = 0 then [x] = 0. To see this, note that
if IIMil = 0 then ||jt-z n || -* 0 for some sequence (zn) C Z. Hence
zn -» x and so, as Z is closed, x E Z.
We call X/Z, endowed with || • ||0, the quotient normed space of X by
Z, and call || • ||0 the quotient norm. Throughout this book, a quotient
Chapter 2: Normed spaces and bounded linear operators 39

space of a normed spaces will always be endowed with the quotient


norm.
Given normed spaces X and Y, and a bounded linear operator
T: X-^ Y, the kernel Z = Ker T = T~l{G) of T is a closed subspace of
A", and 7 induces a linear operator To : X/Z -> Y. Analogously to
many standard results in algebra, we have the following theorem.
Theorem 9. Let X and Y be normed spaces, T E <3l(X, Y) and
Z = Ker T. Let T o : X/Z -> Y be the linear map induced by T. Then
To is a bounded linear operator from the quotient space X/Z to Y, and
its norm is precisely || r||.
Proof.

showing that ||ro|| *


Conversely, let e > 0 and choose an x E X such that ||JC|| = 1 and
||7i|| > | | 7 i | - 6 . Then ||W|| ^ 1 and ||r 0 M|| = ||7i|| > ||r|| - e .
Hence ||r o ||> ||7i|-€. D

In fact, the quotient norm || • ||0 on X/Z is the minimal norm on X/Z
such that if T E a (AT, Y) and Ker TD Z, then the operator To : X/Z ->
Y induced by 7 has norm at most ||r|| (see Exercise 13).
Suppose that X and Y are closed subspaces of a normed space Z,
with ^rn Y = {0} and X+Y = Z. If the projections / ^ : Z -> X, and
/?y: Z - ^ Y, given by Px(x>y) — x an<3 PY&^y) = y> a r e bounded (i.e.
continuous) then we call Z a </irecf 5«m of the subspaces X and Y. It is
easily seen that Z is a direct sum of its subspaces A" and Y iff the topol-
ogy on Z (identified with X@Y = {(x,y): x E X, y G Y}) is precisely
the product of the topologies on X and Y. Note that, if Z is a direct
sum of X and Y, Z' is a direct sum of X' and Y, and X is isomorphic to
X' and Y is isomorphic to Y, then Z is isomorphic to Z'. If Z is a
direct sum of X and Y then the projection px: Z-* X induces an iso-
morphism between Z/Y and X.
Conversely, given normed spaces X and Y, there are various natural
ways of turning X® Y, the algebraic direct sum of the underlying vec-
tor spaces, into a normed space. For example, for 1 =^ p ^ <»5 we may
take the norm ||(*,.y)||p = IKIMUMDIlp- Thus for 1 ^ p < oo we take
ll(^y)ll P = (IWI p +IMI p ) 1/p , and for p = oo we define \\(x,y)\\aa =
max{||jc||, ||_y||}. It is easily seen that all these norms are equivalent;
indeed, each induces the product topology on X@Y. The normed
40 Chapter 2: Normed spaces and bounded linear operators

space (X@Y9 || • ||p) is usually denoted by ^ © ^ Y. Considering X and Y


as subspaces of X@p Y, we see that X@p Y is a direct sum of X and Y.
Finally, given a family {|| • ||r : y E f\ of norms on a vector space V, if

\\x\\ = :

for every x E V, then || • || is a norm on V. Note that the analogous


assertion about the infimum of norms does not hold in general (see
Exercise 15).
Having got a good many of the basic definitions under our belts, we
are ready to examine the concepts in some detail. In the next chapter
we shall study continuous linear functional.

Exercises

1. Show that in a normed space, the closure of the open ball Dr(x0)
(r > 0) is the closed ball Br(x0) and the boundary dBr(x0) of Br(x0)
is the sphere Sr(x0). Do these statements hold in a general metric
space as well?
2. Let Bx D B2 3 • • • be closed balls in a normed space X, where
Bn = B(xn,rn) ={xEX: \\xn-x\\ ^ rn) (rn > r > 0).
Does
00

n Bn ^ 0
71 = 1

hold? Is there a ball B(x,r) r > 0, contained in f|^=1 Bn ?


3. Prove or disprove each of the following four statements. In a com-
plete normed space every nested sequence of closed bounded sets has a
non-empty intersection.
4. Let X = (V, || • ||) be a normed space and W a subspace of V. Sup-
pose | • | is a norm on W which is equivalent to the restriction of
|| • || to W. Show that there is a norm || • ||x on V that is equivalent
to || • || and whose restriction to W is precisely | • |.
5. Let || • || and | • | be two norms on a vector space V and let W be a
subspace of V that is || • ||-dense in V. Suppose that the restrictions
of || • || and | • | to W are equivalent. Are || • || and | • | necessarily
equivalent?
6. For 1 ^ p ^ oo, let ||-|| p be the /p-norm on Un. Show that if
1 ^ p < r ^ oo then \\x\\p ^ ||*||r- For which points x do we have
equality?
Chapter 2: Normed spaces and bounded linear operators 41

Prove that for every e > 0 there is an N such that if N < p < »
then

7. Show that the space defined in Examples 1 (xix) is indeed a


normed space.
8. Let 1 ^ p,q,r ^ « be such that p~x + q~l + r~x = 1, with 1/°°
defined to be 0. Show that for x,y9 z E Cn we have

\\4Ph\\Mr-
i= l
State and prove the analogous inequality for s vectors from Cn.
9. Show that lp is a Banach space for every p, (1 ^ p ^ <») and that
c 0 , the set of all sequences tending to 0, is a closed subspace of /«,.
Show also that lp (1 ^ p < <») and c 0 are separable Banach spaces
(i.e. each contains a countable dense set) while /«, is not separable.
10. Let/?, 9 and r be positive reals satisfying p~l + q-1 = r"1. Show
that for / E L p (0,1) and g E L^(0,1) the function fg belongs to
L r (0,1) and

11. Let ^ = ( 0 , . . . , 0 , 1 , 0 , . . . ) = ( S ^ S a , . . . ) E / p , where l ^ p < o o .


Show that (ei,*2,...) is a basis of / p , called the standard basis, i.e.
every x E lp has a unique representation in the form
00

x = 2 A,*,.
12. For JC = (*,-)" E /^, the support of x is supp* = {i: x( + 0}. Let
1 ^ p < oo and let /x, / 2 , . . . E /p be non-zero vectors with disjoint
supports. Show that X = lin(/-)5° is isometric to lp: in fact, the
map Jf-> /p given by/j »-> ||^|| p ^ defines a linear isometry.
13. Let Z be a closed subspace of a normed space X Show that the
quotient norm on the vector space X/Z is the minimal norm such
that if Y is a normed space, T E <&(X, Y) and Z C Ker T then the
norm of the induced operator To : X/Z —» Y is at most || T\\.
14. A seminorm on a vector space V is a function /?: V—> IR+ such
that p(Ajc) = |A|/?(JC) and pOt+y) ^p(x)+p(y) for all vectors
*,)> E V and scalar A. [Thus a seminorm p is a norm if p(x) = 0
implies x — 0.]
Let {p r : y £ / ] be a family of seminorms on a vector space V
such that
42 Chapter 2: Normed spaces and bounded linear operators

0 < p(x) = sup{py(x): y E / j < oo


for every x E V (x =fi 0). Show that /?(•) is a norm on V.
15. Let Hjclli and ||JC||2 be norms on a vector space V. Is ||JC|| =
min{||jc||1, ||JC||2} necessarily a norm?
16. Consider the vector space c0 of complex sequences x = (xn)™ tend-
ing to 0. For a sequence (xn)i E c 0 , let (**)5° be the decreasing
rearrangement of (\xn |)*. Formally, JC* = x if
|{*: k | > * } | <n^\{k: \xk\ & x}\.
Let 6j ^ 62 ^ ' ' ' > 0 be such that 2 " = 1 &/ = °°- Define, for
x = (xn)i E c 0 ,

Let
do = {xEco: \\x\\' <oc}.
Show that || • ||' is a norm on d0. Is this space complete?
17. For jt e /x set

Ml' = 2
n l
~ n=2
Show that || ||' is a norm on lx and that lx is complete in this
norm. Is ||JC||' equivalent to the ^-norm ||JC||x = 2°° =1 \xn\7
18. Show that on every infinite-dimensional normed space X there is a
discontinuous (unbounded) linear functional. [By Zorn's lemma X
has a Hamel basis, i.e. a set {jcr : y E r} C X such that every
j c £ l has a unique representation in the form x = 2 " ^ixyi
(%ef;l ^/^;»6N).]
19. Prove that if two norms on the same vector space are not
equivalent then at least one of them is discontinuous on the unit
sphere in the other norm. Can each norm be discontinuous when
restricted to the unit sphere of the other?
20. Give two norms on a vector space such that one is complete and
the other is incomplete.
21. Find two inequivalent norms || * || ^ and |||| 2 on a vector space V
such that (V, || • \\t) and {V, || • ||2) are isometric normed spaces.
22. Let Y be a closed subspace of a Banach space X and let x be an
element of X. Is the distance of x from Y attained? (Is there a
pointy E Ysuch that \\x-yo\\ = inf{||jc-y||: y E Y}?)
Chapter 2: Normed spaces and bounded linear operators 43

23. Let / i , / 2 , . . . , / n be linear functionals on a vector space V. Show


that there is a norm || • || on V such that each /• is continuous on
(V, ||-||). Can this be done for infinitely many linear functionals?
And what about infinitely many linearly independent linear func-
tionals?
24 + . Let X be a Banach space. Suppose (xn)i C X is a sequence such
that every x EL X has a unique representation in the form x =
2°° =1 Anxw. Prove that the set {xn}™ consists of isolated points.
25. Check the assertion in Examples 6(x).
26. Let Y be a closed subspace of a normed space X. Show that if
Xj Y and Y are separable then so is X.
27. Let y be a closed subspace of a normed space X. Show that if any
two of the three spaces X, Y and X/Y are complete then so is the
third.
28. Let Y be a subspace of a normed space X. Show that Y is a closed
subspace if and only if its unit ball, B(Y), is closed in X. Show
also that if Y is complete then Y is closed.
29. Prove the converse of one of the assertions of Theorem 4: if
$&(X, Y) is complete 4heh so is Y.
30 + . Let V be a vector space with basis (*„)". [Thus every x E V
(x ^ 0) has a unique expression in the form x = 2* = 1 Kxnt
(«! < n2 < • • • < Wfc and A, ^ 0).] Show that there is no complete
norm on V.
31^. Let C be a closed convex set in a normed space X such that
C+B{X) D B1+€(X) for some e > 0. Does it follow that IntC ^
0 , i.e. that C contains a ball of positive radius?

Notes

Abstract normed spaces were first defined and investigated by Stefan


Banach in 1920 in his Ph.D. thesis at the University of Leopol (i.e. the
Polish town of Lw6w, now in the Soviet Union). Much of this thesis
was published as an article: Sur les operations dans les ensembles
abstraits et leur applications aux equations integrates, Fundamenta
Mathematica, 3 (1922), 133-81. A few years later Banach wrote the
first book wholly devoted to normed spaces and linear operators:
Thiorie des Operations Liniaires, Warsaw, 1932, vii + 254 pp. Banach
gave an elegant account of the work of many mathematicians involved
in the creation of functional analysis, including Frederic (Frigyes) Riesz,
whom he quoted most frequently, just ahead of himself, Maurice
44 Chapter 2: Normed spaces and bounded linear operators

Frechet, Alfred Haar, Henri Lebesgue, Stanislaw Mazur, Juliusz


Schauder and Hugo Steinhaus. It is perhaps amusing to note that, when
writing about Banach spaces, Banach used the term 'espace du type

This beautiful book of Banach has had a tremendous influence on


functional analysis; it is well worth reading even today, expecially in its
English translation: Theory of Linear Operators (translated by F. Jel-
lett), North-Holland, Amsterdam, 1987, ix 4- 237 pp. This edition is
particularly valuable because the second part, by A. Pelczyriski and Cz.
Bessaga {Some aspects of the present theory of Banach spaces, pp.
161-237), brings the subject up to date, with many recent results and
references.
Another classic on functional analysis is F. Riesz and B. Sz.-Nagy,
Functional Analysis (translated from the 2nd French edition by L. F.
Boron), Blackie and Son Ltd., London and Glasgow, 1956, xii + 468
pp. This volume concentrates on the function-theoretic and measure-
theoretic aspects of functional analysis, so it does not have too much in
common with our treatment of linear analysis.
There are a good many monographs on normed spaces and linear
operators, including the massive treatise by N. Dunford and J.
Schwartz, Linear Operators, in three parts; Interscience, New York;
Part I: General Theory, 1958, xiv + 858 pp.; Part II: Spectral Theory,
Self Adjoint Operators in Hilbert Space, 1963, ix + 859-1923 pp. + 7
pp. Errata; Part III: Spectral Operators, 1971, xix + 1925-2592 pp.,
L. V. Kantorovich and G. P. Akilov, Functional Analysis in Normed
Spaces, Pergamon Press, International Series of Monographs on Pure
and Applied Mathematics, vol. 45, Oxford, 1964, xiii + 771 pp., A. N.
Kolmogorov and S. V. Fomin, Introductory Real Analysis (translated
and edited by R. A. Silverman), Prentice-Hall, Inc., Englewood Cliffs,
N. J., 1970, xii + 403 pp., Mahlon M. Day, Normed Linear Spaces,
Third Edition, Ergebnisse der Mathematik und Ihrer Grenzgebiete, vol.
21, Springer-Verlag, Berlin, 1973, viii + 211 pp. and J. B. Conway, A
Course in Functional Analysis, Graduate Texts in Mathematics, vol. 96,
Springer-Verlag, New York, 1985, xiv + 404 pp.
3. LINEAR FUNCTIONALS AND THE
HAHN-BANACH THEOREM

Given a normed space X = (V, ||-||), let us write X' for the algebraic
dual of X, i.e. for the vector space V of linear functionals on V. Thus
X*, the space of bounded linear functionals on X, is a subspace of the
vector space X'.
We know from the standard theory of vector spaces that every independent
set of vectors is contained in a (Hamel) basis (see Exercise 18 of Chapter 2). In
particular, for every non-zero vector wE V there is a linear functional/E V
with/(w) 7* 0. Equivalently, V is a large enough to distinguish the elements of
V: for all x,yE V(x ^ y) there is a functional/E V such that/(x) ¥^f(y\ Even
more, the dual V" of V is large enough to accommodate V: there is a natural
embedding of V into V" which is an isomorphism if V is finite-dimensional.
But what happens if we restrict our attention to bounded linear func-
tionals on a normed space XI Are there sufficiently many bounded
linear functionals to distinguish the elements of XI In other words,
given an element x E X (x ^ 0) is there a functional / E X* such that
f(x) ^ 0? As X is a normed space, one would like to use X* to obtain
some information about the norm on X. So can we estimate ||JC|| by f(x)
for some / E X* with ||/|| = 1? To be more precise, we know that for
every x E X,
\\x\\^sup[\f(x)\:feB(X*)}.
But is the right-hand-side comparable to ||JC||?
As a matter of fact, so far we do not even know that for every non-
zero normed space there is at least one non-zero linear functional, i.e.
we have not even ruled out the utter indignity that X* = {0} while
X = (V> II' ID *s large> saY V i s infinite-dimensional.
The main aim of this chapter is to show that, as far as the questions
above are concerned, Candide and Pangloss were right, tout est pour le
mieux dans le meilleur des mondes possibles', indeed, everything is for

45
46 Chapter 3: Linear functionals and the Hahn-Banach theorem

the best in the world of bounded linear functionals. Before we present


the result implying this, namely the Hahn-Banach theorem, we shall
point out some elementary facts concerning linear functionals.
Let us show that / E X' is bounded if and only if f(B) is not the
entire ground field. Let B = B(X) = B(0,1) be the closed unit ball of
X. If |A| ^ 1 then KB = 5(0, |A|) C B. Hence for f E X' we have
sup|/(2?)| = SUP{|/(JC)| : ||*|| ^ 1} = Il/H, where ||/|| = » if / is
unbounded, and \f(B) = f(\B) C f(B). Consequently f(B) is either
{A: |A| < ||/||} or {A: |A| ^ ||/||}. Similarly, if D = D(X) = D(0,l) is
the open unit ball of X then f(D) = {A: |A| < ||/||} for all f E X'
(f ± 0). Hence / E X' is bounded iff f(B) is not the entire ground
field, as claimed.
It is often useful to think of a (non-zero) linear functional as a hyper-
plane in our vector space. An ajfine hyperplane or simply a hyperplane
H in X is a set

where x0 E X and Y C X is a subspace of codimension 1, i.e. a sub-


space with dimX/Y = 1. We say that H is a translate of Y. Given a
non-zero functional / E X', let K(f) = f~l(0) ={xEX: f(x) = 0} be
the null space, i.e. the kernel, of / and let
I(f)=f-\l)={xEX:f(x) = l}.
Let us recall the following simple facts from elementary linear
algebra.
Theorem 1. Let X be a (real or complex) vector space.
(a) If / E X' and /(JC0) i=- 0, then K(f) is a subspace of codimension
1. Moreover, if f(x0) ^ 0 then every vector x E X has a unique
representation in the form x = y + kx0, where y E K(f) and A is a
scalar.
Furthermore, /(/) is a hyperplane not containing 0.
(b) If f,g E X'-{0} then/ = A g iff K(f) = K(g).
(c) The map / *-> /(/) gives a 1-1 correspondence between non-zero
linear functionals and hyperplanes not containing 0. •

Continuous (i.e. bounded) linear functionals are easily characterized


in terms of K(f) or / ( / ) . Note that ||/|| ^ 1 if and only if |/(JC) | < 1 for
all x with ||JC|| < 1, i.e. if and only if /(/) is disjoint from the open unit
ballD(0,l) = {JCE X: \\x\\ < 1}.
Chapter 3: Linear functionals and the Hahn-Banach theorem 47

A subset A of a topological space T is nowhere dense if its closure has


empty interior.
Theorem 2. Let X be a (real or complex) nOrmed space.
(a) Let / E X\ (f + 0). If / is continuous (i.e. / E X*) then K(f)
and /(/) are closed and nowhere dense in X. If / is discontinuous
(i.e. unbounded) then K(f) and /(/) are dense in X.
(b) The map / *-> /(/) gives a 1-1 correspondence between non-zero
bounded linear functionals and closed hyperplanes not containing
0.
Proof, (a) Suppose that / is continuous. Then K(f) and /(/) are
closed, since they are inverse images of closed sets. If f(x0) ^ 0 then
/(*) i1 f(* + €Xo) f° r € > 0- Hence K(f) and /(/) have empty interiors,
and so are nowhere dense.
Suppose now that K(f) is not dense in X, say B(xo,r)nK(f) = 0 for
some x0 E X and r > 0. Then f(B(xo,r)) = f(xo) + rf(B(X)) does not
contain 0, so f(B(X)) is not the entire ground field. Hence, as we have
seen, / is bounded.
Since /(/) is a translate of K(f), it is dense iff K(f) is dense.
(b) This is immediate from (a) and Theorem 1 (c). •

In fact, B(xo,r)DK(f) = 0 implies a bound on the norm of / ,


namely the bound \\f\\ ^ \f(xo)\/r. Indeed, otherwise \f(x)\ >
|/(JC0) | ||jc||/r for some * E X and so

xf(x0)
y =X

and/(>>) = 0, contradicting our assumption.


Now we turn to one of the cornerstones of elementary functional
analysis, the Hahn-Banach theorem which guarantees that functionals
can be extended from subspaces without increasing their norms. This
means that all the questions posed at the beginning of the chapter have
reassuring answers. Although the proof of the general form of the
Hahn-Banach theorem uses Zorn's lemma, the essential part of the
proof is completely elementary and very useful in itself.
Let Y C X be vector spaces and let f E X' and g £ 7'. If f(y) =
g(y) for all y E Y (i.e. f\Y, the restriction of / to 7, is g) then / is an
extension of g. We express this by writing g C / . A function
p : X^> U* = RU{°°} on a real vector space X is said to be a convex
functional if it is positive homogeneous, i.e. p(tx) = tp{x) for all t ^ 0
48 Chapter 3: Linear functionals and the Hahn-Banach theorem

and x E X, and is a convex function (as used in Chapter 1), i.e. if


x,y E X, and 0 ^ t ^ 1 then p(tx + (l-t)y) ^ tp(x) + (l-t)p(y). By
the positive homogeneity of p, the second condition is equivalent to
p(x+y) ^ p(x) +p{y) for all x9y E X, i.e. to the subadditivity of p. As
customary for the operations on R* = RU{°o}, we use the convention
that oo + 5 = oo + oo = oo for all 5 E R; O.oo = 0; and t.°° = oo for t > 0.
Note that a norm is a convex functional, as is every linear functional.
Furthermore, if X = (V, || • ||) is a normed space then a linear functional
f E X' is dominated by the convex functional JV||JC|| iff / E X* and
Il/H ^ N. As usual, a function <p : 5 -» R is said to dominate a function
«A : 5 -> R if i//(s) ^ <p(s) for all s E S.
Lemma 3. Let p be a convex functional on a real vector space X and let
/o be a linear functional on a 1-codimensional subspace Y of X Sup-
pose that /o is dominated by p, i.e.
/oO0 ^ P(y) for all y E y.
Then / 0 can be extended to a linear functional f E X' dominated by p:
f(x) ^ p(x) for all xEX. (1)
Proof. Fix z E X (z ^ Y) so that every J C E Z has a unique representa-
tion in the form x = y + tz, where y E Y and f E R. The functional /
we are looking for is determined by its value on z, say f(z) = c. To
prove (1), we have to show that for some choice of c we have
f{y + tz)
in other words
(2)
for all y E Y and t E R.
For t > 0 inequality (2) gives an upper bound on c, and for
t = — s < 0 it gives a lower bound. Indeed, for t > 0, (2) becomes

for all y E Y. For s > 0 we have /00>) - JC ^ p(>>-sz) and so

s
for all y E Y. The former holds iff
Chapter 3: Linear functionals and the Hahn-Banach theorem 49

for all / E Y, and the latter holds iff

for all y" E Y. Hence there is an appropriate c iff


My") -Piy"-z) ^ p{y' + z) -fo(y') (3)
for all y',y" E Y. But (3) does hold since

completing the proof. D

The following theorem is a slight strengthening of Lemma 3.


Theorem 4. Let Y be a subspace of a real vector space ^ such that X is
the linear span of Y and a sequence z x , z 2 , • • • • Suppose / 0 E Y' is dom-
inated by a convex functional /? on X. Then / 0 can be extended to a
linear functional f E: X' dominated by p.
If X is a real normed space and / 0 E Y* then / 0 has an extension to a
functional / on X such that ||/|| =
Proof. Set Xn = lin{ Y,z x ,..., zn}. By Lemma 3 we can define linear
functionals / 0 C fx C f2 C • • • such that /„ E X'n and each /„ is dom-
inated by p . Define / : X-> R by setting /(*) = /„(*) if JC E Z n . Then
f E Xr extends / 0 and it is dominated by p.
The second part is immediate from the first. Indeed, / 0 is dominated
by the convex functional p(x) = iV||jt|| where N = ||/0||. Hence there
is an / E X' extending f0 and dominated by p. But then f(x) ^
p{x) = N\\x\\ for all x E X, so that ||/|| ^ N = ||/0||, implying ||/|| =
ll/oll. •
The restriction on Y in Theorem 4 is, in fact, unnecessary. As we
shall see, this is an easy consequence of Zorn's lemma, the standard
weapon of an analyst which ensures the existence of maximal objects.
For the sake of completeness, we shall state Zorn's lemma, but before
doing so we have to define the terms needed in the statement.
A partial order or simply order on a set P is a binary relation ^ such
that (i) a ^ a for every a E P, (ii) if a ^ b and b ^ c for a, b,c E P
then a ^ c, and (iii) if a ^ b and b ^ a for some a,b E P then a = b.
Briefly, ^ is a transitive and reflexive binary relation on P. We call the
50 Chapter 3: Linear functionals and the Hahn-Banach theorem

pair (P, ^) a partially ordered set; in keeping with our custom concern-
ing normed spaces and topological spaces, (P, ^) is often abbreviated to
P. A subset C of P is a chain or a totally ordered set if for all a, b E C
we have a ^ b or b ^ a. An element m 6 P is a maximal element of P
if m ^ a implies that a = m; furthermore, we say that b is an upper
bound for a set 5 C P if 5 ^ 6 for all s E S. It can be shown that the
axiom of choice is equivalent to the following assertion.
Zorn's lemma. If every chain in a non-empty partially ordered set P has
an upper bound, then P has at least one maximal element. •

The fact that Theorem 4 holds for any subspace Y of X is the cele-
brated Hahn-Banach extension theorem.
Theorem 5. Let Y be a subspace of a real vector space X and let
/ 0 £ Y'. Let p be a convex functional on X If / 0 is dominated by p on
Y, i.e. fo(y) ^ /?(>>) for every y E Y, then / 0 can be extended to a linear
functional / E X' dominated by p.
If X is a real normed space and / 0 E Y* then / 0 has a norm-
preserving extension to the whole of X: there is a functional f E X*
such t h a t / 0 C / a n d Il/H = ||/0||.
Proof. Consider the set 9 = {fy : y E f} of all extensions of f0 dom-
inated by p: for each y there is a subspace Yy and a linear functional
fy E Yy such that Y C Yy, f0 C fy and / r is dominated by p. Clearly
the relation ' C is a partial order on 9 (fy is 'less than or equal to' fs iff
fy C / 5 ) . If 9^0 = {/y : y E Fo} is a non-empty chain (i.e. a totally
ordered set) then it has an upper bound, namely f EY\ where
Y = \JyGro Yy and f(y) = fy(y) if y E Yy (y E T o ). Therefore, by
Zorn's lemma, there is a maximal extension. But by Lemma 3 every
maximal extension is defined on the whole of X.
The second part follows as before. •

With a little work one can show that norm-preserving extensions can
be guaranteed in complex normed spaces as well. A complex normed
space X can be considered as a real normed space; as such, we denote it
by XR. We write Xfa for the dual of XR. It is easily checked that the
mapping r: X* -» j $ defined by r(f) = R e / (i.e. r(f)(x) = Re/(jc) for
x E X) is a one-to-one norm-preserving map onto X&. The inverse of r
is the map c : Xfa —> X* defined by c(g)(x) = g(x) - ig(ix). This enables
us to deduce the complex form of the Hahn-Banach extension theorem.
Chapter 3: Linear functionals and the Hahn-Banach theorem 51

Theorem 6. Let Y be a subspace of a complex normed space X and let


/o E y*. Then / 0 has a norm-preserving extension to the whole of X:
there is a functional/ E Z* such that focf and ||/|| = ||/0||.
Proof. By Theorem 5, we can extend r(/ 0 ) to a functional g on XR
satisfying ||g|| = ||r(/ 0 )|| = ||/0||. The complex functional / = c(g) E X*
extends / 0 and satisfies ||/|| = ||/0||. •

The Hahn-Banach theorem has many important consequences; we


give some of them here.
Corollary 7. Let X be a normed space, and let x0 E X. Then there is a
functional / E S(X*) such that f(x0) = ||JCO||. In particular, ||JCO|| ^ C iff
\g(xo)\ ^C for all g<ES(X*).
Proof. We may assume that x0 ^ 0. Let Y be the 1-dimensional sub-
space lin{jc0} and define/ 0 E Y* by f^Xx0) = X\\xo\\. Then ||/0|| = 1 and
its extension / , guaranteed by the* Hahn-Banach theorem, has the
required properties. •

Corollary 8. Let A" be a normed space, and let x0 E X. lff(x0) = 0 for


a l l / E X* thenjco = 0. •

The functional / whose existence is guaranteed by Corollary 7 is said


to be a support functional at x0. Note that if x0 E S(X) and / is a sup-
port functional at x0 then the hyperplane /(/) is a support plane of the
convex body B{X) at x0; in other words: JC0 E B{X) n / ( / ) and /(/) con-
tains no interior point of B{X). The norm on X is said to be smooth if
every x0 E S(X) has a unique support functional.
Corollary 7 implies that the map 28(X, Y) -> <&(Y*,X*), given by
T>-+ T*, is an isometry, as remarked after Theorem 2.4, when we
defined the adjoint.
Theorem 9. If X and Y are normed spaces and T E <&(X, Y) then
T* E & ( Y * , J n and | | r | | = ||r||.
Proof. As usual, we may and shall assume that X and Y are non-trivial
spaces: X + {0} and Y + {0}. We know that ||r*|| ^ ||r||. Given e > 0,
there is an x0 E 5(Z) such that ||7jto|| ^ ||7i|-*. Let g E 5(y*) be a
support functional at Tx0: g(Tx0) = \\Txo\\. Then
(T*g)(x0) = g(Tx0) =
so that \\rg\\ ^ \\T\\-e and | | r | | ^ \\T\\-e. D
52 Chapter 3: Linear functionals and the Hahn-Banach theorem

Given a vector space V with dual V and second dual V" = (V)',
there is a natural embedding V-* V" defined by v »-> i;", where v" is
defined by *;"(/) = f{v) for f E V. Rather trivially, this embedding is
an isomorphism if V is finite dimensional. If X is a normed space with
dual X*, second dual X** and x E X, then we write x for the restriction
of x" io X*\ x is the linear functional on X* given by x(f) = /(JC) for
f E X*. In other words, with the bracket notation,

for all / E X*. Since | i ( / ) | = |/(*)| ^ ||/||||JC||, we have x E X** (not


just x E (X*)'), and moreover ||jc|| ^ ||;t||. The Hahn-Banach theorem
implies that, in fact, we have equality here.
Theorem 10. The natural map x i-* x is a norm-preserving isomorphism
(embedding) of a normed space X into its second dual X**.
Proof. For x E X (x + 0), let / be a support functional at x: ||/|| = 1
and /(*) = \\x\\. Then |*(/)| = |/(*)| = ||JC|| and | i ( / ) | ^ ||i||||/|| =
|||||W|^||i||. D

In view of Theorem 10 it is natural to consider X as a subspace of


X**. If, under the natural identification, X is the whole of X**, i.e.
X = X**, then X is said to be reflexive. We know that X** is complete
even when X is not, so a reflexive space is necessarily complete. How-
ever, a Banach space need not be reflexive. For example, lp is reflexive
for 1 < p < oo, and /x and /«, are not reflexive. Also, as we shall see,
every finite-dimensional normed space is reflexive.
Writing C(L) for the Banach space of bounded continuous functions
on a topological space L with the uniform norm, it is easily seen that
every Banach space X is a closed subspace of C(L) for some metric
space L. Indeed, put L = B(X*), and for x E X define fx = x\L.
Then, by Theorem 10, the map X-+ C(L), given by x •-> fx, is a linear
isometry onto a closed subspace of C(L). We shall see in chapter 8 that
considerably more is true: instead of C(L) we may take C(K), the space
of all continuous functions on a compact Hausdorff space K with the
uniform norm.
To conclude this chapter, let us present a strengthening of the
Hahn-Banach theorem. This time we wish to impose not only an upper
bound but also a lower bound on our linear functional to be found, the
upper bound being a convex functional, as before, and the lower bound
a concave functional.
Chapter 3: Linear functionate and the Hahn-Banach theorem 53

Given a real vector space X, a function q: X^> R, = IRU{-<»} is said


to be a concave functional if -q(x) is a convex functional, i.e. if q is
positive homogeneous and q(x+y) ^ q(x) + q(y), i.e., superadditive.
Given a convex functional p and a concave functional q, our aim is to
find a linear functional f E X' such that
?W^/W^pW (4)
for all xEX.
What condition does / have to satisfy on a subspace Y in order for /
to be extendable? One's first guess is surely that
q(y) ^ f(y) ^ p(y)
for all y E Y. While this condition is undoubtedly necessary, it need
not be the whole story. Indeed, as f{y) = f(x+y)—f(x) and —f(x) ^
-q(x), we must have
f(y)^p(x+y)-q(x) (5)
for all y E Y and x E X. Inequality (5) is stronger than (4): putting
x = 0 in (5) we find that f(y) ^ p(y) for all y E Y, and setting x = -y
we see that f(y) ^ -q(-y), i.e. q(-y) ^ f(-y) for all y E Y. Of
course, if Y = A" then conditions (4) and (5) are equivalent.
The following strengthening of the Hahn-Banach theorem shows that
(5) is sufficient to guarantee the existence of an extension from Y to the
whole of X.
Theorem 11. Let p be a convex functional and q a concave functional
on a real vector space X, let Y be a subspace of X and let f0 E Y' be
such that

for all y E Y and x E X. Then f0 has an extension / E X' such that


q(x) ^ f(x) ^ p{x)
for every x E X.
Proof. The heart of the matter is the analogue of Lemma 3: once we
have managed to extend f0 to a slightly larger (i.e. one dimension
larger) subspace, the rest follows as before.
Pick a vector z E X (z £ Y) and set Z = lin{Y, z}. Let us show that
/o has an extension fx to Z satisfying
54 Chapter 3: Linear functionah and the Hahn-Banach theorem

for all u E Z and x E X, As in the proof of Lemma 3, we have to


show that there is a suitable choice c for/i(z), i.e. that there is a c £ R
such that

and
fi(y'-z)=fo(/)-c^p(x'+y'-z)-q(x')
for all y,y' E Y and *,*' E X Such a c exists if and only if

for all y,y' E Y and JC, x' E X. But this inequality does hold, since
fo(y)+fo(y')=fo(y+y')
^p(x+x'+y+yr)-q(x+x')
z)+p{xl+yt-z)-q{x)-q{x').
Hence / 0 has a suitable extension fa to Z.
This assertion implies the analogue of Theorem 4, and an application
of Zorn's lemma gives our theorem in its full generality. •

Corollary 12. Let p be a convex functional and q a concave functional


on a real vector space X such that q is dominated by p: for all x E X
we have q(x) ^ /?(*). Then there is a linear functional f EV such that
4(*) ^ f{x) ^ p(x) for all xEX.
Proof. Let 7 = (0) C X. Then the trivial linear functional f0 on Y
satisfies the condition in Theorem 11, namely

for all x E X. Hence / 0 has an extension to a linear functional f E Xf


such that q(x) ^ f(x) ^ p(x) for all JC E X •

The following separation theorem is an easy consequence of Corollary 12.


Theorem 13. Let A and B be disjoint non-empty convex subsets of a real
vector space X. Suppose that for some OLEA and every xEX there is an
e(x) > 0 such that a + tE A for all r, \t\^e (x). Then 4 and B can be
separated by a hyperplane, i.e. there is a non-zero linear functional/E JT and
a real number c such that/(x) ^ c ^f(y) for all xE A and yEB.
Chapter 3: Linear functionals and the Hahn-Banach theorem 55

Proof. We may assume that a = 0, i.e., for every xEX there is an


6 = e (x) > 0 such that [ - e x , 6 x ] c A Define functions p and q on X by
setting, for x E X,

p(x) = inf{t ^ 0: x E ^ } ; #(x) = supfr ^ 0: x E tB}.


It is easily checked that p: X -* R is a convex functional and
^ : X^>" R* is a concave functional. Furthermore, as tAiltB = 0 for
f > 0, we have #(JC) =^ /?(*). Hence, by Corollary 12, there is a non-zero
linear functional f E X' such that #(*) ^ f{x) ^ /?(*) for all x E X. To
complete the proof, note that if x E A and y E B then

Hence we may take c = 1. •

As the last result of this chapter, we shall show that the separation
theorem gives a pleasant description of the closed convex hull of a set in
a normed space. The convex hull coS of a set 5 in a vector space X is
the intersection of all convex subsets of X containing 5, so it is the
unique smallest convex set containing S. Clearly,

co5 = I 2 ttXi : ^ e 5 , ^ 0 ( / = l,...,/i), £ ti? = l ( n = 1,2,...)

If Jf is a normed space then the c/ased cowvex Aw// co 5 of 5 is the inter-


section of all closed convex subsets of X containing 5, so it is the unique
smallest closed convex set containing 5. As the closure of a convex set
is convex, coS is the closure of co5.
The following immediate consequence of the separation theorem
shows that co 5 is the intersection of all closed half-spaces containing 5.
It is, of course, trivial that coS is contained in this intersection.
Theorem 14. Let 5 be a non-empty subset of a real normed space X.
Then co5 = {x E X: f(x) ^ sup f(s) for all / E X*}.

Proof. Suppose that x0 £ coS. Then fi(jc0,r)flco5 = 0 for some


r > 0. L e t / E X' (f j= 0) separate the convex sets B(xo,r) and coS:

for all x E B(xo,r) and y E co5. Since the restriction of / to B(xo,r) is


bounded above,/E X* and since/# 0,/(xo) > c. •
56 Chapter 3: Linear functional and the Hahn-Banach theorem

Throughout the book, we shall encounter many applications of the


Hahn-Banach theorem and its variants. For example, the last result
will be used in chapter 8.

Exercises

1. Let p and q be conjugate indices, with 1 ^ p < o°. Prove that


lp = lq> Show also that CQ = l1. [Note that this gives a quick
proof of the fact that for 1 ^ p ^ « the space lp is complete.]
2. Let c be the subspace of L consisting of all convergent sequences.
What is the general form of a bounded linear functional on c?
3. Let p and q be conjugate indices, with 1 ^ p < «>. Prove that the
dual of L p (0,l) is Lq(O,l).
4. Check that for 1 < p < °°, the space lp is reflexive. Check also
that ll9 loo and c0 are not reflexive.
5. Let X and Y be normed spaces, and set Z = X@Y, with norm
Ml = \\(x>y)\\ = MI + IMI- What is Z*?
6. Let X1,X2,... be normed spaces and let X = \\yn=1 Xn be the
space of all sequences x = (xi,x2,...) which are eventually zero:
xn E Xn and xn = 0 if n is sufficiently large, with pointwise opera-
tions and norm
,1/2
Ml = '
What is X*?
7. Let I b e a Banach space. Show that if X*** = X* then X is
reflexive, i.e. X** = X.
8. A linear functional / E /i> is positive if f(x) ^ 0 when every
x = (xn)i E loo with *„ ^ 0 for all «. Show that every positive
linear functional on /«> is bounded.
9. Show thatpo(jt) = Po((xn)i) = limsupn_>ooxn is a convex functional
on loo and deduce the existence of a linear functional / : fc —» R
such that
liminf*n ^ /(*) ^ limsupjcn

for every x = (*„)" E /«,.


10. Let Z b e a complex normed space and let Y be a subspace of Xu.
(Thus if y1,y2E Y and r l 9 r 2 E IR then rlyx + r2y2 E Y.) Let
/o E y* (i.e. let / 0 be a bounded real linear functional on Y) such
Chapter 3: Linear functionals and the Hahn-Banach theorem 57

that if yi9y29^iyi^^2y2 E ^ f° r some A1?A2 E C and yi,y2 e ^


then /0(Aiyi+A2<y2) = Ai/oO>i)+A2/o()>2). s h o w t h a t /o need not
have an extension to a bounded complex linear functional on the
whole of X, i.e. it need not have an extension to a functional
/ £ JT.
11 + . Let X be a Banach space, Y a 1-codimensional subspace of ^ and
X1 and A^2 dense subspaces of X. Is A ^ f l ^ dense in XI Is
ZiflY dense in Y? What are the answers if Xx and X2 have codi-
mension 1?
12. Let ^ be a normed space, Y a dense subspace of X and Z a closed
finite-codimensional subspace of X. Is Zfl Y dense in Z?
13. Let Y be a subspace of a normed space X. Show that the closure
of Yis
Y = n i K e r / : / E X*, Y C Ker/}.
14. Let # be a closed convex set in a real normed space X. Show that
every boundary point of K has a support functional: for every
x0 E d# there is an / E X* such that / ^ 0 and supxeA^/(jc) =

15. Let i be a set of points in a real normed space X, and let


f0: A^>U. Show that there is a functional / E B(^T*) such that
/(*) = fo(a) f° r all 0 E i4 if and only if

aGF

for every finite subset F of A and for every function A : F—» R.


16. Let Y be a subspace of a normed space X and let x E X. Show
that
d(x, Y) = inf{||* ~yII: y E Y> ^ 1
if and only if there is a linear functional / E S^*) such that
Y C K e r / a n d / ( * ) = 1.

The results in the final exercises, all due to Banach, enable us to


define finitely additive 'integrals' of large classes of functions and to
attach a 'limit' to every bounded sequence.

17. Let T = R/Z be the circle group, i.e. the additive group of reals
modulo the integers. Let X = 3^(1") be the vector space of
bounded real-valued functions on T. For
58 Chapter 3: Linear functionals and the Hahn-Banach theorem

f = f(t)EX and «!,...,«„ E R

set

1
r(f;al9...,an) = sup
p - J /(*+«,-)
- o o < f < o o A?

and define

= i n f { r ( / ; a ! , . . . , a n ) : n = 1,2,...; a l 9 . . . 9 a n E R}.

Show that /? is a convex functional and deduce that there is a gen-


eralized 'integral' / ( / ) on 9b(J) such that for f,g E %(T) and
\,H,t0 E R we have

(i) / ( A / + M * ) = A/(/)+/t/(g); (ii) / ( / ) ^ 0 if/^ 0;

(iii) /(/(>+fo)) = /(/W); (iv) / ( / ( - 0 ) = /(/(r));


(v) 7(1) = 1 (i.e. the integral of the identically 1 function is 1).

18. Let X be the vector space of all real-valued functions f(t) on R


such that limsup,_>oo|/(f)| < ». For

f = f(t)EX and au...,anEU

set

r(/;«!,...,an) = limsup- Y f(t+at)


/-co „ ^

and define

= i n f { K / ; a i , . . . , a n ) : / i = 1,2,...; a 1 ? . . . , a n E R}.

Show that p is a convex functional, and deduce that there is a gen-


eralized 'limit' LIM^oo f(t) on X such that

(i)
f-»00 t—*co

(ii) LIM/(O 0 if liminf/(r) =

(iii) LIM/(H LIN


t-*a

(iv) LIM1 = 1.
t—*»
Chapter 3: Linear functionals and the Hahn-Banach theorem 59

19. For x = (xn)i E L and nx< • • • < nk put

'*Kx;«i,...,/!fc) = limsup -
and define
p(x) — 'mi{Tr{x,ni,...,nk): nx < • • • < nk; k = 1,2,...}.
Show that p is a convex functional, and deduce the existence of a
linear functional L: /« -» R such that

for every x E /<». The value L(JC) is said to be a Banach limit or a


generalized limit of the sequence (*„) and is usually denoted by
IM^oo xn o
Show that
(i)

(ii) liminfjcn ^ LIM;cn ^ limsupjcn;

(iii) n+1 n

Let m E M be fixed and define xn = n/m- [n/m\. What is


LIMjcn?
20. Show that if in Theorem 13 we drop the condition on A then the
assertion is no longer true. [Hint. Let AT be a vector space with basis
f" )
el9 e2,. . ., and let A = - B = < £ f2, x 2 ,: rw > 0, n = 1,2,... V. Check
that/(i4) =/(B) = R for all/e A", f # 0.]

Notes
The Hahn-Banach theorem for real normed spaces was proved by H.
Hahn, Uber lineare Gleichungen in linearen Rdumen, J. fur die reine
und angewandte Mathematik, 157 (1927), 214-29; and by S. Banach,
Sur les fonctionnelles lineaires //, Studia Math., 1 (1929), 223-39. The
complex version, namely Theorem 6, was proved by H. F. Bohnenblust
and A. Sobczyk, Extensions of functionals on complex linear spaces,
Bull. Amer. Math. Soc, 44 (1938), 91-3.
Many dual spaces were first identified by F. Riesz, in Sur les
operations fonctionnelles lineaires, Comptes Rendus, 149 (1909), 974-7,
and in Untersuchungen uber Systeme integrierbarer Funktionen,
Mathematische Annalen, 69 (1910), 449-97.
4. FINITE-DIMENSIONAL NORMED SPACES

As the next cautious step in our exploration of normed spaces and


operators on them, we look at the 'smallest' normed spaces, namely the
finite-dimensional ones. As far as the crude classification of norms is
concerned, these spaces are very simple indeed: any two norms on a
finite-dimensional vector space are equivalent. This can be proved in
many different ways; the proof we give here is based on a lemma about
the space /f.
Lemma 1. The closed unit ball of /" is compact.
Proof. We shall show that B(l") is sequentially compact. Let (e^i be
the standard basis of /", so that
n n
e =
2J ^i i 2J | A | | .
i= l i= \

Given (xk)~ C B = B(l?) with xk = Ti=l tik)ei9 we have |A{*> | ^ 1 for


all i and k. As a bounded sequence of complex numbers has a conver-
gent subsequence, by repeatedly selecting subsequences, we can find a
subsequence (xnk)i such that A^1** converges to some scalar A,- for every
i ( l ^ i ^ ri).
Setting x = 2" = 1 ^iei w e find that
n

lim HJC^-JCH = lim 2 lA/"**—Af-| = 0

and so xnk~^> x. Since B is closed, x E. B. •


Theorem 2. On a finite-dimensional vector space any two norms are
equivalent.

60
Chapter 4: Finite-dimensional normed spaces 61

Proof. Let V be an n-dimensional vector space with basis (e,-)?. Let


|| • || j be the /i-norm on V given by

|A,|

and let || • || be an arbitrary norm on V. It suffices to show that || • || and


H'lli are equivalent.
Let Xx = (VjHd), Sx = S(XX) and let f: S1-^U be defined by
f(x) = \\x\\. The set Sx is a closed subset of the compact set B(X{) and
therefore it is compact. Furthermore

\f(x)-f(y)\^\\x-y\\ =

(max|k-||) £ k-y.-l

= (max \x-y\\x
l i

so / is a continuous function on the compact set St. Hence / attains its


infimum m and supremum M on S1. Since /(x) = ||JC|| > 0 for all
x E Si, we have m > 0. By the definition of / , for any x E V we have
mWxh^Wx^MWxh. D

This theorem has several easy but important consequences.


Corollary 3. Let X and Y be normed spaces, with X finite-dimensional.
Then every linear operator T.X-+Y is continuous. In particular,
every linear functional on X is continuous.
Proof. Note that ||JC||' = ||JC|| + ||7JC|| is a norm on X; since ||-|| and ||-||'
are equivalent, there is an N such that ||JC||' =^ A^||A:|| for all x and so

Corollary 4. Any two finite-dimensional spaces of the same dimension


are isomorphic.
Proof. If dim^f=dimY then there is an invertible operator
TE£(X,Y). As both T and T'1 are bounded, X and Y are iso-
morphic. D
62 Chapter 4: Finite-dimensional normed spaces

Corollary 5. Every finite-dimensional space is complete.


Proof. If a space is complete in one norm then it is complete in every
equivalent norm. Since, for example, the space /« is complete, the
assertion follows. •

Corollary 6. In a finite-dimensional space a set is compact iff it is closed


and bounded. In particular, the closed unit ball and the unit sphere are
compact.
Proof. Recall that B(l") is compact, and that a closed subset of a com-
pact set is compact. •

Corollary 7. Every finite-dimensional subspace of a normed space is


closed and complete.
Proof. The assertion is immediate from Corollary 5. •

In fact, as proved by Frederic Riesz, the compactness of the unit ball


characterises finite-dimensional normed spaces. We shall prove this by
making use of the following variant of a lemma also due to Riesz.
Theorem 8. Let Y be a proper subspace of a normed space X.
(a) If Y is closed then for every e > 0 there is a point x E S{X) whose
distance from Y is at least 1 — e:
d(x,Y) =inf{\\x-y\\:yEY}^l-e.
(b) If Y isfinite-dimensionalthen there is a point x E S(X) whose dis-
tance from Y is 1.
Proof. Let z E X\Y and set Z = lin{Y,z}. Define a linear functional
/ : Z -* U by f(y + \z) = A for y E Y and A E R. Then / is a bounded
linear functional since Ker/ = Y is a closed subspace of Z. By the
Hahn-Banach theorem, / has an extension to a bounded linear func-
tional FEX* with ||F|| = H/ll > 0. Note that Y C KerF.
(a) Let x E S(X) be such that
F(x)^(l-e)\\F\\.
Then for y E Y we have

U
\\F\\ Ml
Chapter 4: Finite-dimensional normed spaces 63

(b) If Y is finite-dimensional then F attains its supremum on the com-


pact set S{X) so there is an x E S{X) such that
Fix) = \\F\\.

But then for y E Y we have

11 n L
\\F\\ \\F\\
Corollary 9. Let X\ C X2 C • • • be finite-dimensional subspaces of a
normed space, with all inclusions proper. Then there are unit vectors
Xi,x2,... such that xn E Xn and d{xn,Xn_i) = 1 for all n ^ 2.
In particular, an infinite-dimensional normed space contains an infinite
sequence (xn) of 1-separated unit vectors (i.e. with \\xn— xm\\ ^ 1 for
n ± m).
Proof. To find xnE Xn, apply Theorem 8 (b) to the pair (X, Y) =
(Xn9Xn.x). D

Theorem 10. A normed space is finite-dimensional if and only if its unit


ball is compact.
Proof. From Corollary 6, all we have to show is that if X is infinite-
dimensional then its unit ball B(X) is not compact. To see this, simply
take an infinite sequence x x , JC2 , - - - E B(X) whose existence is
guaranteed by Corollary 9: H*,-—Jty-|| ^ 1 for / ^ /'. As this sequence
has no convergent subsequence, B(X) is not compact. •

Theorem 10 is often used to prove that a space under consideration is


finite-dimensional: the compactness of the unit ball tells us precisely
this, without giving any information about the dimension of the space.
The above proof of Theorem 10 is based on the existence of a
sequence of unit vectors (xn)i such that |JC/—jcy-|| ^ 1 for all i ^ j . Let
us show that, in fact, we can do better: we can make sure that the ine-
qualities are strict. All we need is the compactness of the unit ball of a
finite-dimensional normed space.
Lemma 11. Let jti,...,* n _i be linearly independent vectors in a real
normed space X of dimension n ^ 2. Then there is a vector xn E S(X)
such that Ita-xJI > 1 for all i (1 ^ i < n).
Proof. We may assume that d\mX = n. Let / E S(X*) be such that
f{xi) = 0 for 1 ^ / < n. (In other words, we require
64 Chapter 4: Finite-dimensional normed spaces

K(f) = \in{xl9...,xn-1}
and so we have precisely two choices for / : a functional and its
negative.) Furthermore, let g E X* be such that g(xt) = 1 for every /
(1 ^ i < n). Since S(X) is compact, so is
K = {x E S(X): f(x) = 1}?0.
Let xn E K be such that g(xn) = mm{g{x): x E K}. Then for 1 ^ i < n
we have ||* n -X;|| ^ f(xn-xt) = 1. Since g(*n-.*,•) = #(*„)-1 < g(*n)>
the choice of xn tells us that * „ - * ; £ A', so that ||^-jc f || ^ 1. Hence
Il^-X/H > 1 for every i (1 ^ / < n). D

Theorem 12. Let XX<ZX2C- • • • be subspaces of a real normed space,


with dimXn = n. Then there is a sequence xi,x2,... of unit vectors
such that ||X|-Jty|| > 1 if i j= j and \m{xx,...,xn} = Xn for n = 1,2,... .

There is another elegant way of finding 1-separated sequences of unit
vectors. This time we shall rely on the finite-dimensional form of the
Hahn-Banach theorem. Let us choose vectors JC1?JC2,... E S(X) and
support functional x\,xl,... E S(X*) as follows. Pick xlES(X1) and
let x* E S(X*) be a support functional at xx: **(*!) = 1. Suppose
k<n, dimXn = n, xl9...,xk E S(Xn), xt,...,xi E S(X*)9 xf(Xi) = 1
and X*(XJ) = 0 for i < j . Then

W = f| AT(jcf) = 0 Kerjcf
i= l i= \

has dimension at least n-k (in fact, precisely n-k) and so we can pick
a vector xk+1 E Wf)S(Xn). Let JC| +1 be a support functional at xk+1,
i.e. let*£ + 1 E 5(^r*) and *| + 1 (JC* + 1 ) = 1.
This implies that if Xx C A"2 C • • • are subspaces of X ( d i m ^ = n)
then there are sequences (*„)? C S(X) and (JC*)^ C 5(AT*) such that
Un{xl9...,xn} = ATn, A:*(^) = 1 and x*(^m) = 0 for n < m. In particu-
lar, J t a - j J I 2* lfor/i ^ m.
The canonical bases of lp and lq = /*, where p~l + q~l = \, are
examples of pairs of sequences of this type. In fact, with en =
(0,..., 0,1,0,...) Elp and e* = (0,..., 0,1,0,...) Elq, where the l's
occur in the nth places, we find that \\en\\ = \\e*\\ = e*(en) = 1 and
e e
n( m) = 0 for all n j= m, not only when n < m. A pair of sequences
satisfying these conditions is called a normalised biorthogonal system.
To be precise, given a normed space X, a biorthogonal system on X
Chapter 4: Finite-dimensional normed spaces 65

consists of vectors Xi,..., xn and bounded linear functional x* ,..., x*


such that xf(xf) = 8tj (i.e. xf(xj) = 0 if i ± j and xf{xt) = 1). Such a
system is normalised if ||jt/|| = ||JC*|| = 1 for all i.
Theorem 13. Let X be an n-dimensional real normed space. Then there
is a normalised biorthogonal system (*,•)? ? (*f)i on X.
Proof. Let X = (R n ,||-||) a n d s = S(X). For uu...,unE Un, write
v(ul9...,un) for the n-dimensional Euclidean volume of the n-simplex
with vertices 0 9 U i , . . . , u n . L e t x 1 , . . . , x n E 5 b e such that v ( x i , . . . , x n )
is maximal.
We claim that these vectors will do. What are the functionals (JC*)" ?
Clearly xf has to be defined by **(*,-) = 1 and X*(XJ) = 0 for; ^ /. But
does xf have norm 1? Indeed it does, since ||JC*|| > 1 implies that
xf(yt) > 1 for some yt E 5 and so
v(x1,...,xi-1,yi,xi+1,...,xn) > v(xl9...9xn). •

The sequence (xt)i whose existence is guaranteed by Theorem 13 is


called an Auerbach system.
In order to see what the existence of an Auerbach system really
means, it is worthwhile to reformulate Theorem 13 in geometric terms.
Let Xi,...,xn be a basis of X. Define two norms on X: the /x norm and
the /oo norm defined by taking xi,...,xn as canonical basis. Thus for
x = 2 " = 1 A,*,- set
n
IMIi = 2 |A,-| and IML = max|A;|.
Let X1 and X^ be the normed spaces defined by these norms. Then
Theorem 13 claims precisely that x1,...,xn can be chosen in such a way
that
B(XX) CB(X) CB(XJ. (1)
Clearly B(X1) is the convex hull of the 2n vectors ±* 1 , ± J C 2 , . . . , ±xn
and B{Xn) is the n-dimensional parallelepiped whose 2n vertices are
2 ^ ^ - ( 6 , 6 {-1,1}).
Let us return to the opening statement of this chapter: the isomorphic
classification of finite-dimensional normed spaces is trivial, with two
spaces being isomorphic if and only if they have the same dimension.
Based on this, one could come to the hasty verdict that there is nothing
to finite-dimensional normed spaces: they are not worth studying. As it
66 Chapter 4: Finite-dimensional normed spaces

happens, this would not only be a hasty verdict but it would also be
utterly incorrect. There are a great many important and interesting
questions, only the isomorphic classification is not one of them. All
these questions, many of which are still open, concern the metric pro-
perties of the finite-dimensional normed spaces.
Perhaps the most fundamental question is the following: given two n-
dimensional normed spaces, how close are they to being isometric? Let
us formulate this question precisely.
Let X and Y be isomorphic normed spaces, i.e. let X and Y be such
that there is a bounded linear operator T E ^(X, Y) which has a
bounded inverse. The Banach-Mazur distance between X and Y is
d(X, Y) = inf{||711IIT'1]]: T E <&(X, Y), T~l E
If X and Y are not isomorphic then one defines their Banach-Mazur
distance to be oo.
Note that d(X,Y) ^ 1, d(Y,X) = d(X,Y) = d(X*,Y*) for any two
spaces, and
d(X,Z)^d(X,Y)d(Y,Z)
for any three spaces. Thus in some sense it would be more natural to
measure the 'distance' between X and Y by logd(X, Y).
What does d(X, Y) < d mean in geometric terms? It means precisely
that there is an invertible linear operator T E 2S(Z, Y) such that

— B(Y) C TB(X) C c2B(Y)

for some cx and c2 satisfying cxc2 < d, in other words that


SB(Y)cB(X)CcSB(Y)
for some 5 E ^(Y,X) and c < d. Equivalently, we may demand that
B(Y) C TB(X) C cB(Y)
for some c < d. Thus the distance is less than d if after a linear
transformation the unit ball of one of the spaces is sandwiched between
the unit ball of the other space and c times that unit ball, where c < d.
Corollary 4 tells us that if dimA" = dim Y = n then X and Y are iso-
morphic and so d(X, Y) < oo. But how large can d(X, Y) be? A com-
pactness argument implies immediately that d(X, Y) ^ f(n) for some
function / : N - > R . In fact, the following simple consequence of
Theorem 13 gives a bound on d(X, Y).
Chapter 4: Finite-dimensional normed spaces 67

Corollary 14. Let Jf = (V, || • ||) be an «-dimensional normed space.


Thend(XJ?) ^ n.
Proof. Let (JC,)I be an Auerbach system in X so that, in particular, (x^)\
is a basis of V. Define a norm || -1| x o n Z by setting

i ii 1=1
Then Xx = (V, \\-\\i) is isometric to /"; so we have to show that
iXM^n.
We claim that the formal identity / : Xx -» X is such that ||/|| ^ 1 and
"1!! ^ n. Indeed, for x = 2" = 1 A,-*,- w e have
n n
P(x)\\ = 2 A,-*; 2
1=1
i=\
and so ||/|| ^ 1. (As \\Jxx\\_ = IMx = 1, in fact ||/|| = 1.)
In order to estimate H/"1!!, let (x*)" be the other half of the normal-
ised biothogonal system (JC,-)I, (xf)1; in other words, let xf be defined
by x*(Xj) = 8tj. Given x = 2" = 1 A,-^,-, choose et E {-1,1} such that
c^ = |Af| for i = 1,...,/!. Set / = 2" = 1 ^ * e X*. Then ||/|| ^ n
and

and so

implying H/"1!! ^ n. D

The alert reader may have noticed that the formal proof above is, in
fact, unnecessary, because the result is a trivial consequence of (1), the
relation equivalent to Theorem 13. Indeed, B(Xoo) C nB(Xi) and so
BiXJ C B(X) C nB(Xx). Thus d(X,Xx) ^ n. But Xx is isometric to
/f, so that </(*,/!") ^AZ.
An immediate consequence of Corollary 14 is that d(X, Y) ^ n2 for
any two «-dimensional spaces, but this trivial estimate is far from being
best possible. In 1948 Fritz John proved an essentially best-possible
upper bound for d(X, Y) by first bounding the distance of an
n-dimensional space from /2", rather than from /". Before giving this
result, let us think a little about the Banach-Mazur distance from a
Euclidean space.
68 Chapter 4: Finite-dimensional normed spaces

An easy compactness argument implies that if dimX = dimY = n


then d(X, Y) is attained: there is an operator 5 E ^(Y,X), say, such
that
SB(Y)CB(X)cdSB(Y)
where d = d(X,Y). Now if Y = /2" = (Un, ||-||) then fl(Y) is the
Euclidean unit ball

Z) = {(*,)? ER": £ JC, 2 ^


For a linear operator 5 E X(Rn, Un), the image SZ> of the Euclidean ball
D is an ellipsoid centred at 0; conversely, for every ellipsoid E C Un,
with centre at 0, there is a linear map S E £(Un, Un) such that SD = E.
Thus d(Z, /2rt) ^ d if and only if there is an ellipsoid E, with centre 0,
such that
E C B(X) C d£.
The following elegant and important theorem of John shows that one
can obtain a good upper bound for d(X,l%) be taking the ellipsoid E of
minimal volume containing B(X).
Theorem 15. (John's theorem) Let X= (Un, \\-\\) be a normed space
with unit ball B = B(X). Then there is a unique ellipsoid D of minimal
(Euclidean) volume containing B. Furthermore,
n~1/2D C B C D.
In particular, d(XJ2) ^ n1/2.
Proof. A simple compactness argument shows that the infimum of the
volumes of ellipsoids containing B is attained. Furthermore, every ellip-
soid of minimal volume is centred at 0.
Let us first prove uniqueness. Suppose that D and D' are ellipsoids
of minimal volume containing B. If T E £(Un) is invertible then T(D)
and T(D') are elliposids of minimal volume containing T(B). Hence we
may assume that

D = Ix E Un: £ xf ^ 1 and Dr = Ix E

where ax > 0 (i = 1,..., n).


Chapter 4: Finite-dimensional normed spaces 69

Denoting by vn the volume of D, we have volD' = vn \Yl=1 at, and


so ]1" =1 at = 1. Let E be the ellipsoid

E=

Then
B C DC\D' C E
and
(vol£) 2
< 1
(volD)2

since n"_i fli = 1 a n d not every at is 1. This contradicts the assumption


that D was an ellipsoid of minimal volume containing B.
Let us turn to the proof of n~l^2D C B. Suppose that this is not the
case, so that B has a boundary point in the interior of n~l^2D. By tak-
ing a support plane of B at such a point and rotating B to make this
support plane parallel to the plane of the axes x2, x3,..., xn, we may
assume that

for some c > n1/2.


For a > b > 0 define an ellipsoid Eab by
n
=
En h
i=2
1
so that (volD)/(vol£fl^) = a*"" . If JC E 5 then x e DHP and so
a2x{+b2 2 x 2 = (^-fc2)*!2 + 62 2 ^ 2
i=2 «=1

5
C

Hence B CEab and vol£fl ^ < volZ) whenever

^-^+Z>2^1 and ^"-^l.


c
Thus, to complete the proof, it suffices to show that these inequalities
70 Chapter 4: Finite-dimensional normed spaces

are satisfied for some choice of 0 < b < a. This is a trivial matter
which, unfortunately, looks a little untidy.
Put 0 < e < £, a = (l + e + 26 2 )"- 1 and b = 1-e. Then abn~l > 1.
Also,
+ b2= v / +(1_e)2/1_J_

if 6 > 0 is sufficiently small. •

Theorem 15 can be reformulated in terms of inscribed ellipsoids: there


is a unique ellipsoid D of maximal volume contained in B and
D C B C n1/2D. Furthermore, if B is not a unit ball but only a
bounded convex body in Rn then some translate Bf of B satisfies
D C B' C nD for the ellipsoid D of maximal volume contained in B'
(see Exercise 18).
Corollary 16. Let X and Y be n-dimensional normed spaces. Then
d(X9Y)^n.
Proof. d(X9 Y) ^ d(XJ?)d(l2, Y) ^ n1/2n1/2 = n. D

Although the estimate in Corollary 16 does seem somewhat crude,


since the Banach-Mazur distance between X and Y is estimated by
going through l$9 it is, in fact, close to being best possible: in 1981 the
Russian mathematician E. D. Gluskin proved that there is a constant
c > 0 such that for every n there are n -dimensional normed spaces X
and Y with d(X, Y) ^ en. At the time this result was extremely surpris-
ing; the proof was based on a probabilistic argument which has become
an important tool in the so-called local theory of Banach spaces, the
theory concerning finite-dimensional spaces.

Exercises

1. Let xly...,xn be non-zero elements in a normed space. For


A = (A,-)? E Un set
f n
€ , e {-1,1} ( / = ! , . . . , /
Chapter 4: Finite-dimensional normed spaces 71

Show that || ||o is a norm on Un. Show also that for A = (A,-)?=i
and; = (1)?=1 E Rn we have
l|A||o=ll/llol|A|U=||;||omax|Aj.

2. Let A", = (V;, ||-||) (i = l,...,n) be normed spaces. For 1 «£


p « oo and x = (*,)" £ F = Vj© • • • © Vn set

IkL = ll(lkll)7IL.
Show that || • ||p is a norm on V. Show also that any two of these
norms are equivalent. [The space (V, ||*||p) is usually denoted by
Xx ®pX2®p" • © p -Yn; it is a direct sum of Xx,..., Xn.]
n
3. Let | | be a symmetric norm on R , i.e.

2 A,e, ^ | = |(0,..., 0,1,0,...,0)| = 1

and let Xx,..., Xn be normed spaces. Show that


W(xl9x29...9xn)\\ =
is a norm on X = Xx® • • • ©Jf n , inducing the product topology on
this space. Check that each Xt is a closed subspace of X and that
A' is complete iff each Xt is complete.
4. Let *!,...,*„ be unit vectors in a normed space such that
II
A/jeJ ^ c 2 max |A,-|
i i^/^/i

for all A,-,...,\ n G R. Show that for k2 ^ n there are unit vectors

such that t/j.fl Uv = 0 for i ^ i' and

2 A.-M,- c max JAj

for all \ly...,\k G R.


5 + . Let X be a finite-dimensional real normed space, x0 E X and
T E #pf). Can {Tnx0}i be dense in JT? [HINT: First prove it for
complex normed spaces.]
+
6 . Let Y be a proper subspace of a finite-dimensional normed space
X. Can one always find a vector z 0 E ^ (z 0 ^ 0) such that
72 Chapter 4: Finite-dimensional normed spaces

Ibll for all y E Y? Or a subspace Z CX such that


Y+Z = X and \\y + z\\ & \\y\\ for all y E 7 and z E Z ?
7. Let Y be a finite codimensional closed subspace of a normed space
X and let 71 E S£(X,Z), where Z is a normed space. Show that if
the restriction of T to Y is continuous then so is T.
8. Let xl9...,xn be unit vectors in a normed space. Let 0 < e < \
and suppose that

max

for all A 1 ? ...,A n Prove that

l-e) max

for allA 1? ...,A rt E R.


9. Let X and Y be n-dimensional normed spaces. Show that the
Banach-Mazur distance between X and Y is attained, i.e. there is
an operator To E &(X, Y) such that T$l E &(Y, A") and
') = inf{|| 711| T - 1
10. Show that if 2 =£ p *£ °° then

11. Let X be a normed space with unit ball B = B(X) and suppose
that B can be covered by a finite number of translates of \B\ there
are xx,..., xn E ^ such that

B C U (*, + §£) = U *(*«,§)•


1=1 1=1

By making use of Corollary 7 but no subsequent result, prove that


X is finite-dimensional, and deduce Theorem 10.
12. Let Y be a finite-dimensional subspace of an infinite-dimensional
normed space X. Show that for every e > 0 there is an x E S(X)
such that \\y\\ ^ (l + €)||x+;y|| for every y E Y.
13. A sequence (xn)i in a normed space is said to be a basic sequence
if xn 7^ 0 for every n, and there is a constant K such that
m n
.^^ k k .^^ k k
k=l k=l

for all 1 ^ m ^ AZ and scalars \1,..., \n. The minimal # satisfying


Chapter 4: Finite-dimensional normed spaces 73

this is the basis constant of (xn)i. Deduce from the result in the
previous exercise that every infinite-dimensional normed space con-
tains a basic sequence.
14. Let (xn)i be a basic sequence in a Banach space. Show that (xn)i
is a Schauder basis of its closed linear span, i.e. every x E lin(jcn)^°
has a unique representation in the form x = S L i ^kxk> w ^h the
series convergent in norm.
15. Deduce from Corollary 9 that a Banach space cannot have a count-
ably infinite algebraic basis, i.e. if X = lin(en)~, where e1,e2,...
are independent, then X is incomplete.
16. Let X and Y be finite-dimensional normed spaces and set
NK = {T<= B(X, Y): ||T||, || 3—x||" ^ K}.
Prove that NK is a compact subset of B(X, Y) (with the operator
norm).
17 + . Let X be an infinite-dimensional separable Banach space such that
for every subspace Y C X (dimY = Az) there is an operator
Tn e B(Y,l?) such that H^HH^"1!! ^ K. Prove that there is an
operator T E B(XJ2) such that || 3i| ||y~ x || ^ K.
18. Let B C Rn be a bounded convex body (i.e. Int£ ^ 0 ) . Show
that there is an ellipsoid D of centre 0 such that some translate B'
of B satisfies D C B ' C nD.
19. Prove that if a linear functional is continuous on a closed finite-
codimensional subspace then it is a bounded linear functional.
20. Let Xx = (Vx,|H|) and X2 = (V2,|H|) be subspaces of a normed
space X = (V, \\ -||), with Xt finite-dimensional. Show that if
V = Vi@V2 then Jf is the direct sum of Xx and X2, i.e. the pro-
jections px.: X = Xi + X2~~* X; (i = 1,2) are continuous.

Notes

Fritz John's theorem was proved in the paper Extremum problems with
inequalities as subsidiary conditions, in Courant Anniversary Volume,
Interscience, New York, 1948, pp. 187-204. The original paper of
E. D. Gluskin, The diameter of the Minkowski compactum is roughly
equal to n, Funct. Anal. Appl., 15 (1981), 72-3, is an all-too-concise
account of his celebrated results.
Excellent accounts of much of the excitement concerning finite-
dimensional spaces can be found in the following volumes: V. D. Mil-
man and G. Schechtman, Asymptotic Theory of Finite Dimensional
74 Chapter 4: Finite-dimensional normed spaces

Normed Spaces, Lecture Notes in Mathematics, vol. 1200, Springer-


Verlag, Berlin, Heidelberg, 1986, viii + 156 pp., G. Pisier, Volume of
Convex Bodies and Geometry of Banach Spaces, Cambridge University
Press, 1989, xv + 250 pp., and N. Tomczak-Jaegermann, Banach-
Mazur Distance and Finite Dimensional Operator Ideals, Pitman, 1989.
5. THE BAIRE CATEGORY THEOREM AND
THE CLOSED-GRAPH THEOREM

In this chapter we shall present several of the most fundamental and


useful results in classical functional analysis: the Baire category theorem,
the Banach-Steinhaus theorem, the theorem of the condensation of
singularities, the open-mapping theorem and the closed-graph theorem.
All these results are very closely related; indeed, they are practically
variants of each other. Nevertheless, it is rather useful to emphasize the
many facets of the same phenomenon, because even a subtle variation
in the formulation can make quite a difference in an application.
The simplest member of this group of results is due to Baire and con-
cerns metric spaces. Recall that a set Y in a topological space X is
dense in X if Y, the closure of Y, is X. In other words, Y is dense if
YOG ± 0 whenever G is a non-empty open set. We shall give several
equivalent formulations of the Baire category theorem: here is the first.
Theorem 1. Let Gi,G2," be a sequence of dense open subsets of a
complete metric space X. Then G = f)™=1 Gn is dense in X.
Proof. As in a normed space, for x E X and r > 0 we set
Dr{x) = { y 6 l : d(x,y) < r) and Br(x) = {y e X: d(x,y) ^ r}.
Although Dr(x) need not be Br(x), we have Dr(x) C Br(x) C Dr+€(x) for
all 6 > 0. Since Gn is a dense open subset of X, for all x E X and
r > 0 the open ball Dr(x) meets Gn in a non-empty open set so there
are y E Xand s > 0 such that Bs(y) C D^iy) C GnnDr(x). This is the
property we shall exploit.
Let then x E X and r > 0. We have to show that Dr(x) n G + 0 .
Let us construct sequences of points xx,x2,... E X and positive
numbers rlfr2,... as follows. Choose xt E X and 0 < rx < 1 such that

75
76 Chapter 5: The Baire category theorem

Bri(xx) C GiHDr(x). Next choose x2 E X and 0 < r2 < \ such that


Sr2(jc2) C G2nZ)ri(jc1), then choose * 3 E X and 0 < r3 < \ such that
5r3(x3) C G3nDr2(x2), etc. By construction, rn -» 0 and flri(*i) D
Br2(x2) D • • • and so, by the completeness of X, n°° = 1 Brn(xn) = {x0}
for some JC0 E X As JC0 E ZJri(*i) C Dr(jc) and x0EBrJ(xn) C Gn for all
n, we have x0 E Dr(jc) DG. D

Because of its importance and many applications, we give some other


forms of Theorem 1 and thereby explain the name of the result as well.
The first is a slightly weaker but very useful variant.
Theorem V. If a complete metric space is the union of countably many
closed sets then at least one of the closed sets has non-empty interior.
Proof. Let X — U ^+ x Fn, where X is a complete metric space and each
Fn is closed. Setting Gn = X\Fn we find that f|^=1 Gn = 0 so that at
least one of the open sets Gn is not dense in X: Gn j1 X. But
IntFn = X\Gn. D

Strictly speaking, Theorem V says precisely that the set G in Theorem


1 is not empty. However, replacing X by Br(x), we see that GC\Br(x) is
not empty either, and so G is dense. Thus Theorem 1 is an easy conse-
quence of Theorem Y.
A subset Y of a topological space X is said to be nowhere dense in X
if the closure of Y has empty interior: Int Y = 0 . Note that a set Y is
nowhere dense if and only if its closure, F, is nowhere dense. A subset
Z C X is meagre in X or of the first category in X if it is the union of
countably many nowhere-dense sets. Clearly, the union of countably
many meagre sets is meagre. A subset of X is of the second category in
X if it is not meagre in X. Thus U is of the second category in X if,
whenever F1,F2,... are closed sets and U C U°°=1 Fn-> then ^n^Pn i2 0
for some n.
Let us use this terminology to give another reformulation of the Baire
category theorem.
Theorem 1". The complement of a meagre subset of a complete metric
space is dense. In particular, a complete metric space is of the second
category.
Also, in a complete metric space the complement of a set of the first
category (a meagre set) is of the second category.
Chapter 5: The Baire category theorem 11

Proof. Suppose Y = U^ = 1 Yn C X with X complete and each Yn


nowhere dense in X. Then Yn is nowhere dense in X so Gn = X\Yn is
a dense open set and therefore G = 0 " = 1 Gn is dense. Clearly
G c z\y.
To see the third assertion, note that the union of two meagre sets is
meagre. •
The two parts of the last variant are trivially equivalent since if A is a
set of the second category in a topological space X and B C X is meagre
then A\B is again of the second category.
The intuitive meaning of a set of the second category is that it is a
large set. Thus, saying that 'the set of points having a certain property
(e.g. where a given function is continuous) is of the second category' is
rather similar to saying that 'almost all points have this property', i.e.
the set of points not having this property has measure zero. Of course,
the second statement makes sense only if we have a measure on the
space, while being of the second category is an intrinsic property of a
metric space. For example, given a collection % of convex bodies in Rn,
loosely speaking one may say that 'almost all sets in % have property P\
meaning that in the natural (Hausdorff) metric on % the set of convex
bodies having P is of the second category.
It is perhaps worth pointing out that the two notions are only similar but
not comparable, even if we restrict our attention to [0, 1] C tR. Indeed,
[0, | ] C [0, 1] does not have full measure, yet is of the second category, and
one can construct a subset of [0,1] with measure 1 that is of the first category
(cf. Exercise 23).
In 1897 Osgood proved the following pretty result about continuous
functions. Let /i,/2,--- •' [0,1] —> IR be continuous functions such that
for each t E [0,1] the sequence (/„(*))" is bounded. Then there is an
interval [a,b] C [0,1] (a < b) on which the sequence /i,/2>--- is uni-
formly bounded.
Let us see that the Baire category theorem implies the following
extension of this result.
Theorem 2. (Principle of uniform boundedness) Let U be a set of the
second category in a metric space X and let !F be a family of continuous
functions / : X —> R such that {f{u): f E 3F} is bounded for every
u E U. Then the elements of 3P are uniformly bounded in some ball
BMo) (*o ^ X;r> 0), i.e. |/(JC)| ^ n holds for some n and all / E 9
and x E Br(x0). In particular, the conclusion holds if X is complete and
\f{x): / E 9^} is bounded for every x E X.
78 Chapter 5: The Baire category theorem

Proof. For n ^ 1 let


Fn = {x E X: \f(x)\ ^ n for all / E 9} = f\ f-\[-n,n}).

Then Fn, as the intersection of a family of closed sets, is closed. By our


assumption,

UC U Fn,
n=l
so by the definition of a set of the second category, IntFn ^ 0 for some
n.
The second assertion is immediate from Theorem 1". D

The next result is only a little more than a reformulation of Theorem


2 in the setting of linear maps.
Theorem 3. (Banach-Steinhaus theorem) Let X and Y be normed
spaces, let U be a set of the second category in X and let 9 C 26(Jf, Y)
be a family of bounded linear operators such that
sup{||7w||: T E 9^} < oo
for all u E U. Then ||r|| ^ N for some AT and all T E 9. In particular,
the conclusion holds if X is complete and sup{||7jc||: T E 9} < <*> for all
JCEX

Proof. For T E 9? the function A"-* R, given by JC •-* ||7x||, is continu-


ous. Hence, by Theorem 2, there is a ball Br(x0) (JC0 E I ' ; r > 0) such
that || 7*|| ^ n for some n and all T E & and JC E Br(x0). As in the
proof of Theorem 2.2, this implies that ||r|| ^ N = n/r for all T E 9.
Indeed, let T E 9 and JC E 2?CY). Then jco + rjt and xo-rx are in
2?r(jc0) and so

||Ii|| = jr\\T{x0 + rx-{x0-rx})\\ ^ * = N

and therefore ||7i| ^ A^ for all T E 9. D

The last result has the following immediate consequence concerning


the condensation of singularities.
Theorem 4. Let X and Y be normed spaces, with X complete, and let
Tnm E ®CY, Y) (/i, m = 1,2...) be such that
GSllr^H =oo for all n.
Chapter 5: The Baire category theorem 79

Then there is a set U C X of the second category in X such that for


« £ [ / w e have
)^\\Tnm{u)\\ =oo for all n.

Proof. For a fixed n, let Vn C X be the set of vectors v such that


fimm^oo||7;m(*;)|| < oo. Then, by Theorem 3, Vn is of the first category.
Hence so is V = U^ = 1 Vn and thus by Theorem 1" the set U = X\V is
of the second category in X. •

The next major result, the open-mapping theorem, is a consequence


of the Baire category theorem and the important lemma below. Recall
that for a normed space X we write D(X) for the open unit ball and
B{X) for the closed unit ball; furthermore, we use the notation
Dr{X) = rD(X) and Br{X) = rB(X).
Lemma 5. Suppose X and Y are normed spaces, X is complete,
T E »(JT, Y) and r(Dr(JT)) D £>,(Y). Then T(Z) r W) D DS{Y).
Proof. We may and shall assume that r = s = 1. Let .4 =
D(Y)C\T(D(X)); then ,4 = B(Y). Given z E D(Y), choose S such that
||Z||<1-5<1 and set y = z/(l-S). We shall show that
y E T(D(X))/(1-S) and so z E r(Dffl).
Let us define a sequence (^)o C Y. Set y0 = 0 and then choose suc-
cessively yi,y2,... E Y such that y n - ^ n - i E S"" 1 ^ and ||j>n-.y|| < 8n.
Having chosen ^ o ^ - ^ n - i ' w e c a n ^n<^ a suitable yn since
^ E D(<yn_1,8/I"1) and the set 8n~lA is dense in D5«-i(Y).
By the definition of A, there exists a sequence (jcn)~ C X such that
7i w = yn-yn-i and ||xj| < Sn~\ Putting

we have

W * £ |k|| < (l-S)- 1 and Tx= i {yn-yn.x) = y.


n=\ n=\
Consequently y E T(D(X))/(1-S) and so z E T(D(X)). As z was an
arbitrary point of D(Y), this implies that D(Y) C rDCY), as claimed. D
Theorem 6. (Open-mapping theorem) Let X and Y be Banach spaces
and let T E <3&(X, Y) be a surjection: r W = Y. Then T is an open
map, i.e. if U C X is open then so is TU.
Chapter 5: The Baire category theorem

Proof. Set G = T(D{X)). Since T is a linear map, it suffices to show


that G contains a neighbourhood of 0 in Y.
Note that T(Dr(X)) = rG and the closure of rG is rG. Since
Y = T(X) = U " = 1 HG, by the Baire category theorem we see that
Int nG ± 0 for some n and so IntG ^ 0 . The set G is convex and
symmetric about 0 so Int G ^ 0 implies that G D Dr(Y) for some
r > 0. Indeed, if Z)(jco,r) C G then D(-jc o ,r) C G and so the convexity
of G implies that £>(0,r) = Dr{Y) C G. But then, by Lemma 5,
G = T(D(X)) D Dr(Y). D

The final two results are easy but very useful consequences of the
open-mapping theorem.
Theorem 7. (Inverse-mapping theorem) If T is a one-to-one bounded
linear operator from a Banach space X onto a Banach space Y then T~l
is also bounded.
Proof. Since T is a bijection, the inverse T~l exists and belongs to
2(Y,X). By Theorem 6, for some r > 0 we have T{D{X)) D Dr{Y).
Thus \\T~l\\ ^ l/r. •

Theorem 8. (Closed-graph theorem) Let ^f and Y be Banach spaces and


T E L(A", Y). Then T is bounded iff its graph,
l\T) = {(JC, Tx):xEX}CXxY
is closed in Jfxy in the product topology.
Proof. One of the implications is trivial: the graph of a continuous map
into a Hausdorff space is closed, so if T is bounded, F(T) is closed.
Suppose then that F(T) is closed. Let Z be the direct sum of X and
Y endowed with the norm ||(*,}>)|| = IWI + fMI- This norm induces the
product topology on XxY and so, by assumption, F(T) is a closed sub-
set of Z. In fact, Z is easily seen to be complete, so F(T) is a closed
subspace of the Banach space Z and therefore it is itself a Banach
space. The linear map U: F(T) —> X, given by U((x,y)) = x, is a
norm-decreasing bijection and so, by Theorem 7, the inverse operator
U~l is bounded. But then ||r|| ^ \\U~l\\ since
llrxHW + llrxHll^-^HIIf/^IIW. •
It is worth pointing out that the closed-graph theorem often makes it
much easier to prove that an operator is continuous. When is a linear
operator T: X^> Y continuous? If xx,x2,... EL X and xn -> x0 imply
Chapter 5: The Baire category theorem 81

that Txn -* Tx0. Now if X and Y are Banach spaces, the closed-graph
theorem tells us that in order to prove the continuity of a linear opera-
tor T: X —• y, it suffices to show that if xn —> x0 and Txn -> y0 then
Txn -> 7JC0, i.e. 7JC0 = >>o - a considerable gain!

Exercises

1. Let / : R + - > R b e a continuous function such that \\mn^^f{nx) =


0 for all x > 0. Show that lim^oo/fr) = 0.
2 + . Let / : R - » R be infinitely differentiate. Suppose for every
x EU there is a kx such that /(*}(JC) = 0 for all k^ kx. Prove that
/ is a polynomial.
3. Let K = [0,1], X = CU(K) and, for w ^ 1, set Fn =
f(t+h)-f(t)
\feX:3tEK such that
h
n V h with t+h
R
Prove that Fn is closed and nowhere dense in C (K). Deduce that
4
the set of continuous nowhere-differentiable real-valued functions
on [0,1] is dense in CR(K).
4. Prove that if a vector space is a Banach space with respect to two
norms then the topologies induced by the norms are either
equivalent or incomparable (i.e. neither is stronger than the other).
5. Let V be a vector space with algebraic basis e ^ ^ , . . . (so
dim V = oo and every v E V is a unique linear combination of the
et) and let || • |f be a norm on V. Show that (F,||||) is incomplete.
6. Let X be a normed space and S C X. Show that if {/(*): x E 5} is
bounded for every linear functional / E X* then the set 5 is
bounded: ||JC|| ^ K for some K and all x E 5. (Using fancy termi-
nology that will become clear in Chapter 8, a weakly bounded set
is norm bounded.)
7. Deduce from the result in the previous exercise that if two norms
on a vector space V are not equivalent then there is a linear func-
tional / E V which is continuous in one of the norms and discon-
tinuous in the other.
8. Let Z be a closed subspace of 1^(0,2). Suppose for every
/ E Li(0,1) there is an F E X whose restriction to (0,1) is / .
Show that there is a constant c such that our function F can always
be chosen to satisfy ||F|| ^ c||/||.
9. Let 1 ^ p,q ^ °° and let A = (a^)™ be a scalar matrix. Suppose
for every x = (XJ)™ E lp the series 2°°=1 aijxj i s convergent for
82 Chapter 5: The Baire category theorem

every /, and y = (yt)i E lq, where yt = 2°°=1 tf//*y- Show that the
map A: lp-> lq, defined by x *-» y, is a bounded linear map.
10. Let <pw : [0,1] -> R+ (n = 1,2,...) be uniformly bounded continu-
ous functions such that

I <pn(x) dx^c
Jo
for some c > 0. Suppose cn ^ 0, (« = 1,2,...) and
00

for every JC E [0,1]. Prove that

n \
11. For two sequences of scalars x = (xt)i and y = (yt)^ set

Let 1 < p , ^ < oo be conjugate indices and let x{n) = (jc/n))°°=1 E /p


(n = 1,2,...). Show that (x(n\y) -> 0 for every y E / f l iff (x (n) )"
is norm bounded (i.e. \\xin)\\ ^ K for some AT and all n) and
^.(w) _» o for every /. (This is a characterization of sequences in lp
tending weakly to 0; cf. Exercise 6.)
12. Let x(n) = (x/"))r=i E lx (n = 1,2,...). Show that (xin),y) -> 0 for
every y E c0 iff U(w))i° is bounded and jc/n) —> 0 for every i. (Simi-
larly to Exercise 11, this is a characterization of sequences in lx
tending weakly to 0.)
13. Show that for 1 ^ p < oo the Lp-norm

on C[0,1] is dominated by the uniform norm ||/|| =


and deduce that C[0,1] is incomplete in the norm || • || p .
14. Let P be a projection on a Banach space X, i.e. let P be a linear
map of X into itself such that P2 = P. Show that X is the direct
sum of the subspaces KerP = P'^O) and ImP = PX if and only if
P is bounded.
15. Let X and Y be normed spaces and T E ^ ( ^ , y). Show that the
graph of T is closed iff whenever xn —> 0 and 7jtn —>.y then y = 0.
Chapter 5: The Baire category theorem 83

As we shall see in the following exercises, the results of this chapter


are very useful in proving the fundamental properties of Schauder bases.
Recall from Chapter 2 that a sequence (xn)i is a Schauder basis or sim-
ply a basis of a Banach space X if every x E X has a unique representa-
tion in the form 2^11 A,-*,-, w ^h the series convergent in norm.

16. Let (xn)i be a basis of a Banach space X Define Pn E ^(Z) by

= 2 A,-*,-,
\i=i

and for x E X set |||*||| = supn||PwJc||. Prove that ||| • ||| is a norm on
X and X that is complete in this norm as well.
17. Let X, (xn)™ and (Pn)i° be as in the previous exercise. Prove that
Pn E <&(X) and supJ|P n || < ». [The number s u p J P j is called
the 6&sw constant of 005°; c^- Exercise 4.13.]
18. Let JCJ , JC2 , be non-zero vectors in a Banach space X, with
lin(jcw)5° = X Prove that (xn)i is a basis of X if and only if there
is a constant K such that

2 A,-*,- 2
1=1

for all 1 ^ m < n and scalars \1,..., \n.


19. Define a sequence of functions ^ n : [0,1] —> [—1,1] as follows.
Set xi(t) - 1 and for n ^ 2 define A: ^ 0 by 2* < « ^ 2^ +1 . Set
/ = n-2k-l and define

-1 i f / + i ^ 2kx < / + 1 ,
0 otherwise.
The sequence (xn(t))7 *s called the Haar system. Prove that the
Haar system is a basis in Lp(0,1) for 1 =^ p < ».
20. Define a sequence of continuous functions <pn : [0,1] —> [0,1] as
follows. Set <pi(t) = 1 and for n ^ 2 put

=I
where \k ls the A:th Haar function, defined in the previous exer-
cise. The sequence ((Pnit))™ is called the Schauder system. Prove
that the Schauder system is a basis in CR[0, 1] with the uniform
norm. Show also that the basis constant is 1.
84 Chapter 5: The Baire category theorem

21. Let (xn)™ be a sequence in a normed space such that


00

> I # I V I •^ 00

for every f E X*. Show that there is a constant M such that

£ |/(* n )| *s AfH/ll
for every/ e X*.
22. Use the Baire category theorem to deduce the result in Exercise
4.15: the linear span {not the closed linear span!) of an infinite
sequence of linearly independent vectors in a normed space cannot
be complete.
23 + . Construct a set S c [0, 1] of measure 1 that is of the first category in
[0, 1].

Notes

Most of the results in this chapter can be found in Stefan Banach's clas-
sic, cited in Chapter 2. The original paper of R. Baire is Sur la conver-
gence de variables reelles, Annali Mat. Pura e Appl., 3 (1899), 1-122;
Osgood's theorem is in W. F. Osgood, Non-uniform convergence and
the integration of series term by term, Amer. J. Math., 19 (1897),
155-90. The Banach-Steinhaus result was proved in S. Banach and H.
Steinhaus, Sur le principe de la condensation de singularites, Fundamenta
Math., 9 (1927), 51-7.
6. CONTINUOUS FUNCTIONS ON COMPACT
SPACES AND THE
STONE-WEIERSTRASS THEOREM

In this chapter we shall study one of the most important classical


Banach spaces, the space C(K) of continuous complex-valued functions
on a compact Hausdorff space K, with the uniform (supremum) norm
||/|| = sup \f(x) I =max|/(*)|.
xSK XE.K
In fact, as we shall see in Chapter 8, every Banach space is a closed
subspace of C(K) for some compact Hausdorff space K, although this
does not really help in the study of a general Banach space.
First we shall show that C(K) is large enough, namely that it contains
sufficiently many functions; for example, every continuous function on a
closed subset of K can be extended to a continuous function on the
whole of K without increasing its norm. This result, which is a special
case of the Tietze-Urysohn extension theorem, is thus reminiscent of
the Hahn-Banach theorem: it ensures that the space C(K) is large, just
as the Hahn-Banach theorem ensured that the dual space of a normed
linear space was large. It is worth emphasizing that the existence of a
good stock of continuous functions on a topological space X cannot be
taken for granted, not even when X seems to be very pleasant. For
example, it can happen that X is a countable Hausdorff space and still
every continuous function on X is constant (see Exercise 19).
In the second part of the chapter we shall show that C(K) is not too
large in the sense that it is the norm closure (i.e. the closure in the
topology induced by the norm) of a good many 'small' subspaces of
functions.
We shall start with a standard result in analytic topology which is
likely to be familiar to many readers.

85
86 Chapter 6: Continuous functions on compact spaces

A topological space is said to be normal if every pair of disjoint


closed sets can be separated by disjoint open sets: if A and B are dis-
joint closed sets then there are disjoint open sets U and V such that
A C U and B C V. Equivalently, a space is normal if for all Ao C £/0,
where Ao is closed and Uo is open, one can find a closed set Ax and an
open set Ux such that Ao C Ux C Ax C Uo.
Lemma 1.
(a) In a Hausdorff space any two disjoint compact sets can be
separated by open sets.
(b) Every compact Hausdorff space is normal.
Proof, (a) Let A and B be non-empty disjoint compact sets in a Haus-
dorff space. For a E A and b E B, let U(a,b) and V{a,b) be disjoint
open neighbourhoods of a and b. Let us fix a point a E A. Clearly,
B C
Vb<EB v(a>b) a n d so> a s B i s compact, B C U" =1 Ha A ) for
some bi,...,bn E 5. Set

U(a) = 0 £/(«,*,-) and V(a) = \J V(aybt).

Then U(a) and V(fl) are disjoint open sets, U(a) is a neighbourhood of a
and V(a) contains B.
Now note that Ufl e ^ U(a) is a n open cover of A and so, as A is
compact, >1 C \Jm=1 U{at) for some a 2 ,..., am E A. Let
m m
U= U t/(«/) and V = fl K««).
1=1 /=i

Then U and V are disjoint open sets, A C U and 5 C K


(b) In a compact space every closed set is compact, so the assertion
follows from (a). •

If A and B are disjoint closed sets in a metric space X then there is a


continuous function / : X-* [0,1] such that / is 0 on A and 1 on B;
indeed,

/ W
[ 1 *E£
will do. The next important result tells us that such an / exists not only
on metric spaces but also on normal spaces.
Theorem 2. (Urysohn's lemma) Let A and B be disjoint closed subsets
of a normal space X. Then there is a continuous function / : X —> [0,1]
such that A C f-\0) and B C / ^ ( l ) .
Chapter 6: Continuous functions on compact spaces 87

Proof. Let q0 = 0, ql - 1 and let <72>93>--- ^e an enumeration of the


rationals strictly between 0 and 1. Set Uo = 0 , Ao = A, U\ = X\B and
Ai = X. Let us construct a sequence of closed sets A0,Ai,... and a
sequence of open sets f/0, U^,... such that £/,• C At and if ^ < qj then
AidUj. Suppose we have found A0,...,An-i and £/ 0 ,..., Un-\.
What shall we choose for An and Unl Let ^ be the maximal qt
(i < n) which is less than qn and let qt be the minimal qt (i < n) which
is greater than qn. Then Ak C f// and so, as X is normal, one can find
a closed set An and an open set Un such that Ak <Z Un <Z An C. Ut.
Having constructed A0,Al,... and UQyU1,..., it is easy to define a
suitable continuous function / : for x G X set

/(*) = inf{qn : x G An}.


Clearly 0 ^ f(x) ^ 1 and / is 0 on A and 1 on B. All we have to check
is that / is continuous. Suppose then that 0 ^ qk < f(x0). Then
x0 $i Ak and so X\Ak is a neighbourhood of x 0 ; for x G X\Ak we have
/(x) ^ qk. Similarly, if/(x0) < ql ^ 1 then xEUt. So L/z is a neighbourhood
of xQ and for x G Ut we have /(x) ^ ^/. D

We are ready to present one of the main results of this section. By


Urysohn's lemma, if A and B are disjoint closed (possibly empty) sub-
sets of a normal space X then there is a continuous function g =
g{A,B)\ X^> [-1,1] such that g\A = - 1 and g\B = 1. Equivalently,
if C is a closed subset of X and /i is a continuous function from C into
the two-point set {-1,1} then h has an extension to a continuous func-
tion g : X-* [-1,1]. The Tietze-Urysohn extension theorem below
claims that every continuous function/: C —• [—1,1] has such an exten-
sion. This result is easily proved by using Urysohn's lemma to provide
us with continuous functions on X whose restrictions to C are better and
better approximations of /.
Indeed, setting Ao = {x G C: f(x) ^ 0} and Bo = {x G C: f(x) ^ £},
there is a continuous function Go: X -^ [0, j] such that G0|^40 = 0 and
G0\BQ = \. Then / ' = f-G0\C maps C into [-1,^] - a very good
start indeed. Next, set
A,={xGC: f'(x) *£ - 1 } , fi^^eC: /'(*) ^ 0}
and take in a continuous function if1 :C->[—£, 0] such that Hl\Al = —\
and H^i?! = 0. Then Fl = Go + Hj is a good approximation of/ on C,
namely/-- FJC maps C into [ — i, •§] and \\FX\\ ^ \, Continuing in this
way, we get continuous functions F2, F3> ... with ||Fn|| < 2~n and
88 Chapter 6: Continuous functions on compact spaces

f ~ S" = 1 /v | C mapping C into [-2~n92"n]. Thus F = Y°=1 % will have


the required properties.
For the sake of completeness, in the proof below we shall present a
streamlined version of this approach.
Theorem 3. (Tietze-Urysohn extension theorem) Let C be a closed
subset of a normal space X and let / : C—• [—1,1] be a continuous
function. Then / has a continuous extension to the whole space X, i.e.
there is a continuous function F: X —> [—1,1] such that F\C = f.
Proof. We shall construct F as the uniform limit of continuous functions
on the whole of X whose restrictions to C are better and better approxi-
mations of / . To be precise, set f0 = / ,
A0 = {xEC: fo(x) ^ - J } and Bo = {x tE C : fo(x) & J } .

By Urysohn's lemma there is a continuous function F o : X—• [—5,5]


such that F0\A0 = — 5 and F0\B0 = 5. Set fx = / 0 - F 0 | C and note that
fi is a continuous map of C into [ - § , § ] . In general, having con-
structed a continuous function

define
An = {xEC: fn(x) < ~mn} and Bn = {x G C
By Urysohn's lemma there is a continuous function

such that

Fn\An = ~mn and


Set

and note that fn + l is a continuous map C into [-(§)' 1 + 1 ,(§)' l + 1 ] .


Let / o , / 1 , . • • and F0,F1,... be the sequences of functions constructed
in this way. Then |Fn(jc)| ^ 5(5)" for all i £ l and AZ ^ 0 and so
F(x) = S ^ = o Fn(x) ^s continuous, being a uniform limit of continuous
functions, and

\Hx)\ <$ i (i)n = 1.


n=0
Chapter 6: Continuous functions on compact spaces 89

Furthermore, for x E C and n ^ O w e have

fix) - 2 hi-
k=0
and so f(x) = F(x). D

It may be worth noting that it is trivial to ensure that \F(x)\ ^ 1: if


F(x) is any continuous extension of / then
F(x) = min{l,max{-l,F(jc)}}
will do.
From now on, let AT be a compact Hausdorff space and let us consider
the space C(K) of all continuous (complex-valued) functions on K. This
is a Banach space in the uniform norm ||/|| = sup{|/(*)|: x E K}. The
results above tell us that C(K) is rather rich in functions. Namely, C(K)
separates the points of X: for any two distinct points x and y there is a
function / E C(K) such that f(x) ± f(y). Also, if A and B are disjoint
closed subsets of K then for some / E C(K) we have f(a) = 0 and
f(b) = 1 for all aEA and b E B. Finally, if A C K is closed and
/ E C(A) then / = F\A for some F E C(K). To see the last statement,
simply write / as a sum of its real part and its complex part.
The space C(K) is closed under uniform limits, and this is another
source of functions that can be guaranteed to belong to C(K). This
leads us to consider relatively compact subsets of C(K), that is subsets in
which every sequence has a subsequence convergent to some element of
C{K).
Recall that a topological space is compact if every open cover has a
finite subcover, it is countably compact if every countable open cover
has a finite subcover and it is sequentially compact if every sequence has
a convergent subsequence. Given a metric space X and N C X, the set
N is an e-net if for every x E X there is a point y E N with d(x,y) < e.
The e-nets one tends to consider are finite. A metric space is totally
bounded or relatively compact if it contains a finite e-net for every
e > 0, or, equivalently, if every sequence has a Cauchy subsequence.
For a metric space, the properties of being compact, countably compact
and sequentially compact coincide; furthermore, a metric space is com-
pact iff it is complete and relatively compact. In particular, a subset of
a complete metric space is relatively compact iff its closure is compact; a
closed subset of a complete metric space is compact iff it is relatively
compact.
90 Chapter 6: Continuous functions on compact spaces

In order to characterize the relatively compact subsets of the complete


metric space C(K), let us introduce some more terminology. A set
S C C(K) is uniformly bounded if it is bounded in the supremum norm,
i.e. if Il/H ^ N for some N and all / E S. We say that 5 is equicontinu-
ous at a point x E K if for every e > 0 there is a neighbourhood Ux of x
such that if y E Ux and / E S then \f(x) —f(y) | < e. Furthermore, 5 is
equicontinuous if it is equicontinuous at every point.
Theorem 4. (Arzela-Ascoli theorem) For a compact space ^ , a set
5 C C(A') is relatively compact if and only if it is uniformly bounded
and equicontinuous.
Proof, (a) Suppose that S is a totally bounded subset of C(K). Let
x E K and e > 0. The set 5 contains a finite e-net: there is a finite set
{/i,...,/„} C S such that if / E 5 then | | / - / | | < e. Then, trivially,
Il/H < 6 + max l5S/ss J|/|| and so S is uniformly bounded. Furthermore, as
fi is continuous, for every x EL K there is a neighbourhood Ut of JC such
that \fi(x)-fi(y)\ < e for y E £/,-. Then t/ = f!"=1 ^ is a neighbour-
hood of x showing that S is equicontinuous at x. Indeed, for / E S let /
(1 ^ i ^ n) be such that | | / - / | | < e. Then for y E (/ we have

I/to -7001 ^ I/to -//to I + I/to -/(r) I + \fi(y) -f(y) I < 3c.
(b) Now suppose that S is uniformly bounded and equicontinuous.
Given e > 0, for x E K let Ux be an open neighbourhood of x such that
I / t o - / W I < 6 i f / 6 5 and y E [ / r Then K - ( J ^ ^ ; l e t
UXi,...,UXn be a finite subcover of K. Since 51 is uniformly bounded,
the set R = {(/(*i),...,/(*„)): / E 5} is a bounded subset of /£, the
vector space C 1 endowed with the supremum norm. (As all norms on a
finite-dimensional vector space are equivalent, we could have chosen
any norm on C".) A bounded set in l£ is totally bounded because a
bounded closed set in l£ is compact; therefore R is totally bounded so
there are functions f x , . . . , fm E S such that the set

is an e-net in R.
We claim that the functions / i , . . . , / m form a 3e-net in S. Indeed,
given / E S, there is an / (1 ^ / ^ m) such that \f(xj) - / ( x y ) | < e for
all j (l^j^n). Let now x G X be an arbitrary point. Then xG Ux.
for some y (1 ^ / ^ n). Consequently,
|/to - / t o | ^ |/to -/(*y) | + \f(xj) -/•(*,) | + |/(*,) - / t o | < 36. •
Chapter 6: Continuous functions on compact spaces 91

In conclusion, let us see that CU(K), the space of continuous real-


valued functions on K, is closed under monotone pointwise limits. This
is Dini's theorem.
Theorem 5. Let ^ be a compact topological space and let
/ 0 , / 1 ? / 2 , . . . ECU(K). Suppose that for every x E K the sequence
(/w(jt))" is monotone increasing and tends to fo(x). Then (fn)i con-
verges uniformly to / 0 , i.e. ||/ n -/ 0 || -* 0.
Proof. Let e > 0. For every x E K there is a natural number nx such
that fn(x) > fo(x) - 6 . As fo~fnx is a non-negative continuous function,
x has an open neighbourhood Ux such that if y E Ux then

Note that if n ^ nx and y E Ux then

Since U Ux is certainly a cover of A' and K is compact, # is covered


by a finite number of these sets: K = \Jk=l Ux. for some points
Xi,..., xk. The last inequality implies that if n ^ n0 = max{nXi,..., nXk}
then l / o W - Z n W l < e for ally, and so | | / 0 - / J | ^ 6 . D

It is worth noting that all three conditions are needed in Theorem 5,


namely that K is compact, the sequence (/n(jc))" of continuous functions
is pointwise increasing and the pointwise limit, f0 is a continuous func-
tion (see Exercise 13).
Now let us turn to one of the most fundamental results in functional
analysis, the Stone-Weierstrass theorem, telling us that the sapce C(K)
is not too large: it is the uniform closure of a good many pleasant sub-
spaces, more precisely, it is the uniform closure of a good many subalge-
bras. Recall that a topological space is said to be locally compact if
every point has a compact neighbourhood. For a locally compact Haus-
dorff space L, let
C(L) = {/ E C L : / is continuous and bounded},
C0(L) = {/ E C(L): the set {x : \f(x) | ^ e} is compact for every e > 0}
and
CC(L) = {f E C(L): supp/ is compact},
where supp/, the support of a (not necessarily continuous) function / , is
the closure of the set {x : f(x) ^ 0}.
92 Chapter 6: Continuous functions on compact spaces

Note that if L is compact then C(L) is just the set of all continuous
functions; similarly, CC(L) is the set of all continuous functions with
compact support and C0(L) is the set of all continuous function f on L
for which {x E L: \f(x) | ^ e} is compact for every e > 0.
If / and g are functions on the same space, define A/, / + g and /g by
point wise operations. With these operations, C(L) is a commutative
algebra. This algebra has a unit, the identically 1 function. By
definition, CC(L) C C0(L) C C(L) and CC(L) and Q(L) are subalgebras
of C(L). We know that the vector space C(L) is a Banach space with
the uniform (supremum) norm:
H/ll =sup{|/(*)|:*eL}.
Theorem 6. Let L be a locally compact Hausdorff space. The function
|| • || is a norm on C(L) and the space C(L) is complete in this norm;
C0(L) is a closed subspace of C(L) and CC(L) is a dense subspace of
C0(L). Furthermore, for/,g E C(L) one has

n&n« II/II II*II (i)


and 1, the identically 1 function, has norm 1.
Proof. The only assertion needing proof is that CC(L) is dense in C0(L).
Given / E Q(L) and e > 0, set Ko = {x E L: \f(x)\ ^ c}. Then Ko is
compact. For x £ ^ 0 let Ux be an open neighbourhood of x with
compact closure Ux. Then ^ 0 C \JX(EK Ux a n ^ s o ^o ^ U n = 1 ^ f° r
some Xi,...,xn E AT0. Setting ^ = \Jn_x Ux., we find that Kx is a
compact set such that Ko C Int Kx.
By Urysohn's lemma (Theorem 2) there is a continuous function
h\ Kx-> [0,1] that is 1 on Ko and 0 on A^UntKx. Extend h to a func-
tion g on L by setting it to be 0 outside Kx:
h(x) ifxEKl9
gW
{ 0 if xEL\Kv
T h e n g E C C (L) a n d so g f E C C (L). Clearly 11^-/11 ^ 6 . •

An algebra A which is also a normed space and whose norm satisfies


(1) is called a normed algebra. If the algebra has an identity e and the
norm of e is 1 then it is called a unital normed algebra. If the norm is
complete then the algebra is a Banach algebra. We saw in Chapter 2
that 2SUQ, the set of bounded linear operators on a normed space X, is
a unital normed algebra and if X is complete then ^(X) is a unital
Chapter 6: Continuous functions on compact spaces 93

Banach algebra. The trivial part of Theorem 6 shows that C{L) is a


commutative unital Banach algebra, CQ(L) is a commutative Banach
algebra and CC(L) is a commutative normed algebra.
The spaces CU(L), C^(L) and CR(L) are defined analogously; they
consist of the appropriate continuous real-valued functions. These
spaces are not only real normed algebras but they are also lattices under
the natural lattice operations/vg and/Ag.
Given real-valued functions / and g on a set 5, define new functions
/ v g and f/\g on 5 by setting, for x E 5,
(/vg)(*) = maxtf(jt),g(jc)} and (f/\g)(x) = min{/(jt),g(*)}.
We call / v g the join off and g and f/\g the meet of f and g. The join
and the meet are the lattice operations.
Of course, the join of two functions is just their maximum and the
meet is just their minimum. It is clear that CR(L), C*(L) and CR(L)
are all closed under the lattice operations.
For a set A of bounded functions on a set 5, the uniform closure A of
A is the closure of A in the uniform topology on 3^(S), the space of
bounded functions on 5 with the uniform norm (Example 2.1 (ii)). Thus
a function f on S belongs to A if for every e > 0 there exists a g E A
such that \f(x) -g(x) | < e for all x E 5.
In this chapter we shall study the uniformly closed subalgebras of
C(L), C0(L), CR(L) and CR(L); so we are particularly interested in uni-
form approximations by elements from various subsets of these algebras.
As the next result shows, the lattice operations are very useful when we
look for such approximations.
Theorem 7. Let K be a compact Hausdorff space and let A be a subset
of CH(K) which is closed under the lattice operations. Then A, the uni-
form closure of A, is precisely the set of those continuous (real-valued)
functions on K that can be arbitrarily approximated at every pair of
points by functions from A.
Remark. Note that A is not assumed to be a subalgebra, not even a
subspace, it is merely a set closed under the lattice operations, i.e. it is a
sublattice of CU(K).

Proof. It is obvious that the uniform closure A is at most as large as


claimed.
Suppose then that / E CR(K) can be approximated by elements of A
at every pair of points. We have to show that / E A.
94 Chapter 6: Continuous functions on compact spaces

Let e > 0. The existence of approximations of / means precisely that


for x, yE K there exists a function/ xy E A such that
I / t o ~fx,y(x) \<e and \f{y) -fx,y(y) \ < e.
Put UXjy = {z E K: fXjy{z) < f{z)+e}. Then UXty is an open set
containing x and y. Fixing x, we find that K = U eK ^x,y> a n d s o a
finite number of the sets UXty also cover K, say K = UJLi ^*,yj- ®y
assumption, the minimum of the corresponding functions A, > v **>/*, ;v*'
namely fx = fx,yi^fx,y2A''' A
fx,yn> a l s o belongs to A. Clearly
A to > / t o " e and fx(y) < f{y) + 6 for all y E K.
Let V, = {y EK:fx(y) > f(y)-e}. Then Vx is open and K =
U Vx so that, as AT is compact, AT = U l i V*, f° r some
xl9 . . . , x m E K Let fe=fxvfX2v-wfXm be the maximum of the
corresponding functions. By assumption, f€ E A, and
f(x) -e < f€(x) < / t o +6 for all x E K.
As there is such an f€ for every e > 0, / E A. •

Lemma 8. A uniformly closed subalgebra A of 3^(S), the space of


bounded real-valued functions on a set S is also closed under the lattice
operations.
Proof. Since (/vg)W = max{f(x),g(x)} = M/to+gto + l/to"gto|}
and f/\g = f+g-fvg, it suffices to show that | / | E A if / E A. Furth-
ermore, as for n ^ 1 we have |/i/| = n |/|, and / E A holds iff nf E A,
it suffices to show that if / E A and ||/|| = swpx<Es\f(*)\ < 1 then
\f\GA.
Let then / e A, \\f\\ < 1 and e > 0. Let 2 £ = 0 ck(t- \)k be the Tay-
lor series for (t+e2)1/2 about t = \. As this series converges uniformly
on [0,1], there is a polynomial
n
P(t) =1
Jo k'(t~h )*
such that
e2)1/2|
\P(t)-(t+ <
e for all 0 « t« 1.
Hence
\P(s2)-(s- 2 + e2)l/2|
<e for all - U J
Chapter 6: Continuous functions on compact spaces 95

and so
|P(s2)-|s|| <2e for all - 1 ^ s ^ 1.
2
Set Q{s) = P(s )—P(0). Then Q is a real polynomial with constant
term 0, and so / E A implies that Q{f) E A. Furthermore, for
- 1 =^ s ^ 1 we have
\Q{s)-\s\\ ^ \P{s2)-\s\\ + \P{0)\ <4e.
Hence
|G(/)W-|/W|| <4£ for all* EK
and so
\\Q(f)- I/I 11*4*,
showing that | / | can be approximated by the function Q(f) E A within
4e. Since A is uniformly closed, | / | E A •

We are ready to prove the first form of the main result of this section.
We say that a set A of functions on a set S separates the points of S if
for all x,y E S (x ^ y) there is a function / E A with /(JC) ^ /(}>)•
Theorem 9. (Stone-Weierstrass theorem) Let K be a compact space
and let A be an algebra of real-valued continuous functions on K that
separates the points of K. Then A, the uniform closure of A, is either
CR(K) or the algebra of all continuous real-valued functions vanishing at
a single point *«, E K.
Proof, (a) Suppose first that there is no point Jto, E K such that
f(Xoo) = 0 for all / E ^4. It is easily checked that then for all x ^ y
(x,y E K) there is a function / E ,4 such that f(x) ± f(y), /(JC) ^ 0 and
/(>0 ^ 0. Then any function on K, say g E UK, can be approximated at
x and >> by elements of A. In fact, we can do rather better than only
approximate: since (f(x),f(y)) and (f2(x),f2(y)) are linearly indepen-
dent vectors in U2, some linear combination of them is (g(x),g(y)). So
for some linear combination h = sf+tf2 of / and f2 we have h(x) = g(x)
and h(y) = g(y). By Theorem 7 and Lemma 8 we have A = CU(K).
(b) Suppose now that /(xoo) = 0 for some Xoo E K and all / E A. We
have to show that if g E CR(K) and gOO = 0 then g E A.
Let us adjoin the constant functions to A, i.e. let B C CR(AT) be the
algebra of functions of the form f+c where / E A and c £ i . Given
e > 0, by part (a) there exists a function h = f+c E B, with / E A and
96 Chapter 6: Continuous functions on compact spaces

cEU, such that ||/+c-g|| < e. As #(*«,) = fix*) = 0 we have \c\ < e
and so \\f-g\\ < 2e. Hence g E A. D

Let us see now what Theorem 9 tells us about subalgebras of C(L).


For a locally compact Hausdorff space L, the one-point compactification
ofL is the topological space on L^ = HJ{X«>} where x^ <£L and a set S C L^
is open iff (a) Sf]L is open in L and (b) if x^ E S then L/S is compact. It is easily
checked that L^ is a compact Hausdorff space inducing the original topology
on L. Set
£?(£«) = ^ E CR(Loo) : /(*„) = 0}.
Then C^iL^) is a closed subalgebra of C^L^). Furthermore, if L is a
locally compact Hausdorff space then /»->/|L is an isometric isomor-
phism between C^(Loo) and C0R(L). This is why the elements of C®(L)
are said to be the continuous functions on L vanishing at oo.
These remarks imply the following reformulation of the Stone-
Weierstrass theorem for real functions. We shall say that a set A of
functions on a set S separates the points of S strongly if for all x,y E 5
(x i1 y) there are functions, f,g E A such that f(x) ^ f(y), fix) ^ 0 and
g(y) f 0. Note that by Theorem 9 if a subalgebra A of CR(AT) separates
the points of K strongly then A = CU(K).
Theorem 9'. Let L be a locally compact Hausdorff space and let A be a
subalgebra of C0R(L) strongly separating the points of L. Then ^4 is
dense in C^iL), i.e. A = C^(L). D

It is obvious that in Theorem 9' we cannot replace C^(L) by CU(L)


because QR(L) itself is a uniformly closed subalgebra of CR(L). Furth-
ermore, we cannot replace C0R(L) by CoiL) either. This can be seen
from the following example. Let A = {z E C: \z\ ^ 1} be the closed
unit disc in C and let A consist of those continuous functions on A that
are analytic in the interior of A. Then A is a closed subalgebra of
Co(4) = C(A), it strongly separates the points but does not contain the
function /(z) = z E Co(4).
The example above is, in fact, not an ad hoc example but one that goes
to the heart of the matter: what we lack is complex conjugation. This
leads us to a form of the Stone-Weierstrass theorem for complex func-
tions, which is an easy consequence of the second variant, Theorem 9'.
Theorem 10. (Stone-Weierstrass theorem for complex functions) Let L
be a locally compact Hausdorff space. Suppose A is a complex
Chapter 6: Continuous functions on compact spaces 97

subalgebra of Q(L) which is closed under complex conjugation and


strongly separates the points of L. Then A = C0(L).
Proof. Let Au be the set of real-valued functions in A. Then Au is a real
subalgebra of A and it is also a subalgebra of C(f(L). As AR contains
( / + / ) / 2 and ( / - / ) / 2 i for every / E A, it satisfies the conditions of
Theorem 9'. Therefore, C0R C A and so
iC0R(L) C A and C0(L) = C?(L) + iQR(L) c l •

Let us note some immediate consequences of the Stone-Weierstrass


theorem; the first is the original theorem due to Weierstrass.
Corollary 11. Every continuous real-valued function on a bounded
closed subset of Un can be uniformly approximated by polynomials. •

Corollary 12. Every continuous complex-valued function on the circle


T = {z E C: \z\ = 1} can be uniformly approximated by trigonometric
polynomials 2£=_ rt ckzk. •

Corollary 13. Let K and L be compact Hausdorff spaces and endow


KxL with the product topology. Then every continuous function
/ : KxL —> C can be uniformly approximated by functions of the form
Ti=l gitoWy), where gi E C(K) and ht E C(L) (1 *s i *s n). •

The Stone-Weierstrass theorem is very useful in showing that various


subspaces of function spaces are dense. As an example, let us look at
the real space L p [0,1], where 1 ^ p < <». The subspace of L p [0,1]
consisting of continuous functions is dense. Since every continuous real-
valued function on [0,1] can be uniformly approximated by polynomi-
als, and the uniform norm dominates the Lp-norm, the space of polyno-
mials is also dense in L p [0,1].

Exercises

1. Show that a metric space is compact iff it is countably compact iff


it is sequentially compact.
2. Show that a metric space is compact if and only if it is totally
bounded and complete.
3. Show that a subset of a complete metric space is totally bounded if
and only if its closure is compact.
98 Chapter 6: Continuous functions on compact spaces

4. A topological space X is completely regular if for every closed set


A C X and point b not in A there is a continuous function
/ : X-+ [0,1] such that f\A = 0 and/(ft) = 1. Prove that the half-
open interval topologies of the real line are completely regular.
5. Let A be a subset of a normal topological space X and let
f:A—> [0,1] be a continuous function. Does it follow that / has
a continuous extension to the whole of XI And what is the answer
if X is a compact Hausdorff space?
6. Let CR(X) be the normed space of bounded continuous real-
valued functions on a topological space X, with the uniform
(supremum) norm. Show that CR(X) is a Banach space.
7. Let V be a subspace of CU(X) such that whenever A and B are
non-empty disjoint closed subsets of X and a, b E R then there is a
function / E V such that f(X) C [a,6], /|A = a and f\B = b.
Prove that V is dense in CU(X).
8. Let K be a compact metric space and S C C(K). Show that S is
equicontinuous iff for every e > 0 there is a 8 > 0 such that
\f(x)-f(y)\<€ whenever d(x,y) < 8. Is this true for every
metric space Kl
9. Prove the following complex version of the Tietze-Urysohn exten-
sion theorem. Let A be a closed subset of a normal topological
space X and let / : A —> C be a continuous function such that
II/IU = supA:e^l/Wl = 1- Show that there is a continuous func-
tion F: X^> C such that ||F||^ = supxBX\F(x)\ = 1.
10. Let L be a locally compact Hausdorff space. Show that a set
5 C C0(L) is totally bounded iff it is bounded, equicontinuous and
for every e > 0 there is a compact set A^ C L such that |/(x)| < e
for all * G L \ # and / E S.
11. Let £ r = { z E C : |z| < r}, /) = {zEC : |z| < 1} and let F be a
set of functions analytic on D such that |/(z)| ^ Afr if |z| = r < 1.
Show that for r < 1 the set F | £ r = {/|5 r : / E F} is a relatively
compact subset of C{Br). Deduce that every sequence (fn)i C F
contains a subsequence (f^)™ which is uniformly convergent on
every Br (r < 1) to a function / analytic on D.
12. Let U be an open subset of C and let / j , / ^ , . . . : £/-» C be uni-
formly bounded analytic functions. Prove that there is a subse-
quence (fnj™ which is uniformly convergent on every compact
subset of U.
13. Show that all three conditions are needed in Dini's theorem, i.e.
that the conclusion ||/ n -/o|| —• 0 in Theorem 5 need not hold if
Chapter 6: Continuous functions on compact spaces 99

only two of the following three conditions hold: (i) K is compact;


(ii) the functions are continuous; (iii) (fn(x))°i is monotonic.
14. Let G be an open subset of U2 and let / : G —> IR be continuous.
Use the Arzela-Ascoli theorem to prove Peano's theorem that for
each point (xo,yo) E G, at least one solution of
y'ix) = f(x,y)
passes through (xo,yo). [HINT: Let V be a closed neighbourhood
of (xo,yo) and set K = sup{|/(jc,y)| : (x,y) E V}. Choose a <
x0 < b such that

: x E [a,b],x <K\CU.
x-x0
The aim is to show that our differential equation has a solution
through (xo,yo), defined on [a,b]. Find such a solution by consid-
ering piecewise linear approximations, say with division points

15. Let K be a compact metric space. Prove that C(K) is separable.


16. Let AT be a compact Hausdorff space. Suppose C(K) = U " = 1 Q>
where each Cn is an equicontinuous set of functions. What can
you say about Kl
17 + . We say that e > 0 is a Lebesgue number of a cover U G r Uy °f a
metric space if, for every point x, the ball B(x,e) is contained in
some JJy. Show that a metric space (X, d) is compact iff for every
metric equivalent to d every open cover has a (positive) Lebesgue
number.
18. Let K be a compact metric space, L a metric space, and let C(K,L)
to be set of continuous mappings from K to L with the uniform
metric
d(f,g) = supKd(f(x),g(x)) = maxd(f(x),g(x)).
Show that a subset S of C(K,L) is totally bounded if and only if it is
equicontinuous (i.e. for all x E K and € > 0 there is a 8 > 0 such
that if d(x,y) < 8 then d{f{x)J(y)) < e for every / E 5) and the
set {/(JC) : / E 5} is a totally bounded subset of L for every x E K.
+
19 . Construct a countable Hausdorff space on which every continuous
function is constant. The first such example was given by Urysohn
in 1925.
20. Let K be a compact Haudsorff space. What are the maximal
ideals of CU(K)7 And the maximal closed subalgebras?
100 Chapter 6: Continuous functions on compact spaces

21. Let K be a compact Hausdorff space. For a continuous surjection


<p of K onto a compact Hausdorff space K', let <p*: CR(K') -»
C R (/0 be given by <?*(/)(*) = (fip)(x) = f(<p(x)). Show that Imcp*
is a closed unital subalgebra of CU(K) and every closed unital
subalgebra of CR(K) can be obtained in this way. Give a similar
description of the closed subalgebras of CR(K).
22. Let A" be a normal topological space and let Ut,...,Un be open
sets covering X. Prove that there are continuous functions
/ i , . . . , / n : X-+ [0,1] such that 2 £ = 1 /*(*) = 1 for every x and
/k(x) = 0 for x£t/ fc (fc=l,...,n). (The functions fl9...,fn are
said to form a partition of unity subordinate to the cover

Notes

The relevant results of Urysohn and Tietze are in P. Urysohn, Uber die
Mdchtigkeit der zusammenhdngenden Mengen, Math. Ann., 94 (1925),
262-95 and H. Tietze, Uber Funktionen, die auf einer abgeschlossenen
Menge stetig sind, J. fur die reine und angewandte Mathematik, 145
(1915), 9-14. Heinrich Tietze proved a special case of the extension
theorem, while Paul Urysohn (who was Russian but tended to publish in
German) proved Theorems 2 and 3 in the generality we stated them.
Various special cases of the extension theorem were proved by a good
many people, including Lebesgue, Brouwer, de la Vallee Poussin, Bohr
and Hausdorff. For the Arzela-Ascoli theorem see C. Arzela, Sulle
serie di funzioni (parte prima), Memorie Accad. Sci. Bologna, 8 (1900),
131-86. Of course, there are several excellent books on general topol-
ogy; J. L. Kelley, General Topology, Van Nostrand, Princeton, New
Jersey, 1963, xiv and 298 pp., is particularly recommended.
The original version of Theorem 9 is in K. T. Weierstrass, Uber die
analytische Darstellbarkeit sogenannter willkurlicher Funktionen reeler
Argumente, S.-B. Deutsch Akad. Wiss. Berlin, Kl. Math. Phys. Tech.,
1885, 633-39 and 789-805. The Stone-Weierstrass theorem itself is in
M. H. Stone, Applications of the theory of Boolean rings to general
topology, Trans. Amer. Math. Soc, 41 (1937), 375-81; an exposition of
the field can be found in M. H. Stone, The generalized Weierstrass
approximation theorem, Math. Mag., 21 (1948), 167-84.
7. THE CONTRACTION-MAPPING THEOREM

This chapter is more or less an interlude in our study of normed spaces.


Given a set 5 and a function / : S -> 5, a point x E S satisfying f(x) = x
is said to be a fixed point of / . There is a large and very important
body of 'fixed-point theorems' in analysis, that is, results claiming that
every function satisfying certain conditions has a fixed point. Theorems
of this kind often enable us to solve equations satisfying raiher weak
conditions. The aim of this section is to present the most fundamental
of these fixed-point theorems and some of its great many applications.
In Chapter 15 we shall return to the topic to prove some more sophisti-
cated results.
A map f:X-*Xofa, metric space into itself is said to be a contrac-
tion if
d(f(x),f(y)) ^ kd(x,y) (1)
for some k < 1 and all x,y E X. One also calls / a contraction with
constant k. It is immediate from the definition that every contraction
map is continuous; in fact, it is uniformly continuous. Although we
shall not make much use of this terminology, a function / between
metric spaces is said to satisfy the Lipschitz condition with constant k if
(1) holds for all x and y. Thus a map / from a metric space into itself is
a contraction map if it satisfies a Lipschitz condition with constant less
than 1.
The result below is usually referred to as BanacKs fixed-point theorem
or the contraction-mapping theorem.
Theorem 1. Let / : X—> X be a contraction of a (non-empty) complete
metric space. Then / has a unique fixed point.

101
102 Chapter 7: The contraction-mapping theorem

Proof. Suppose that k < 1 and / satisfies (1) for all x,y E X. Pick a
point x0 E X and set xr = f(x0), X2 = /(*i), and so on: for n ^ 1 set
*« = /(**-i)- Writing d for d(xo,xi), we find that
d(^,jcn+1)^rd. (2)
Indeed, (2) is trivially true for n = 1; assuming that it holds for n, we
have
d(xn+1,xn+2) = d(f(xn),f(xn+1)) *£ kd(xn,xn+l) *£ kn+1d.
The triangle inequality and (2) imply that if n < m then
d(xn,xm) ^ d(x
m—n — l
2 kn+id

k
" d.
\-k
Hence (JCW)Q is a Cauchy sequence and so it converges to some point
x EX. Since / is continuous, the sequence (/(*„) )o converges to f(x).
But ( / 0 0 ) o = COT* converges to x and so f(x) = x. The uniqueness
of the fixed point is even simpler: if f(x) = x and f(y) = y then
d(x,y) = d{f{x)J{y)) ^ kd(x,y), and so d(x,y) = 0, i.e. JC = y. •

When looking for a fixed point of a contraction map / , it is often use-


ful to remember that the fixed point is the limit of every sequence (XH)Q
consisting of the iterates of a point JC0, i.e. defined by picking a point x0
and setting xn = /(*„-!) for n ^ 1. Let us note the following slight
extension of Theorem 1.
Theorem 2. Let X be a complete metric space and let / : X —> X be such
that / " is a contraction map for some n ^ 1. Then / has a unique fixed
point.
Proof. We know from Theorem 1 that fn has a unique fixed point x,
say. Since
r(/w) = /(/"(*)) = /to,
/(JC) is also a fixed point of / and so f(x) = x. As a fixed point of / i s
also a fixed point of /", the map / does have a unique fixed point. •

The very simple results above and their proofs have a large number of
applications. In the rest of this chapter we shall be concerned with two
Chapter 7: The contraction-mapping theorem 103

directions: first we shall study maps between Banach spaces and then we
shall present some applications to integral equations.
The proof of Theorem 1, sometimes called the method of successive
approximations, is very important in the study of maps between Banach
spaces. Here is a standard example concerning functions of two vari-
ables which are Lipschitz in their second variable. Recall that for a
normed space X we write Dr(X) = {x E X: \\x\\ < r} for the open ball
of centre 0 and radius r.
Theorem 3. Let X and Y be Banach spaces and let
F:Dr(X)xDs(Y)-+Y
be a continuous map such that
\\F(x,y1)-F(x,y2)\\^k\\y1-y2\\ (3)
and
\\F(x,0)\\<s(l-k) (4)
for all x E Dr(X) and yl9y2 E DS(Y), where r,s > 0 and 0 < k < 1.
Then there is a unique map y : Dr{X) -» DS{Y) such that
y(x) = F(x,y(x)) (5)
for every x E Dr{X), and this map y is continuous.
Proof. For x £ Dr(X), define yo(x) = 0, yt(x) = F(x,yo(x)) = F(JC,0)
and d=\\y1(x)-y0(x)\\ = \\y1(x)l Note that, by (4), d<s(l-k).
Furthermore, if n 3= 1 and yn(x) £ ^ ( y ) then set
yn+1(x) = F(x ,yn(x)).
We claim that yn(x) is defined for every n, and that
n
b>n+1(x)-yn(x)\\^k> d<k" s(l-k) (6)
and
n d
<S
i=0 ~~ \-k: '
As in the proof of Theorem 1, this follows by induction on n. Indeed,
(6) clearly holds for n = 0; assuming that yo(x)9yi(x)9...9yn(x) E DS(Y)
and (6) holds for n — 1, inequality (3) implies that
bn+i(x)-yn(x)\\ = ll^x.^W)
104 Chapter 7: The contraction-mapping theorem

In turn, this implies that (7) holds for n:

lbi.+iWll^lboWII+ 2 Ito+ito-y/toll

^ 2 kld

and so, in particular, yn+i(x) E DS{X).


As F is a continuous function, each yn : Dr(X) —> DS(Y) is continu-
ous. Inequalities (6) and (7) imply that the sequence (yn)o is uniformly
convergent to a continuous function y from Dr(X) to DS(Y).
Then (5) holds trivially:
y(x) = \imyn(x) = lim F(x9yn-1(x)) = F(x,y(x)).

Furthermore, >>(*) is unique, since if yo(x) is a solution of (5) then


\\y(x)-yo(x)\\ = ll^,^))-^^^))!!^^!!^)-^)!!,
= yo(x). D

The next result claims that a suitable map close to the identity map is,
in fact, a homeomorphism.
Theorem 4. Let I b e a Banach space and let e : DS{X) —> X be a con-
traction with constant A: < 1 such that

where s > 0 and r = \s{\ - k). Define / : DS{X) -> Z by


/(*) =
Then there is an open neighbourhood U of 0 such that the restriction of
/ t o Uis a homeomorphism of C/onto Dr(X).
Proof. Define F : Dr(X)xDs(X) -> * by

Then F satisfies the conditions of Theorem 3. Indeed, (3) holds by


assumption; furthermore, if x E Dr(X) then

and so (4) also holds. Hence, by Theorem 3, there is a unique function


Chapter 7: The contraction-mapping theorem 105

g: Dr(X) -> DS(X) such that


g(x) = F(x,g(x)) = x-e(g(x)),
i.e.
x = g(x)+e(g(x))=f(g(x))
for all JC E £>rU0. Set [/ = g(Dr(X)). To complete the proof, we have
to check that U is an open neighbourhood of 0 and f:U-+ Dr(X) is a
homeomorphism. Clearly,/is a one-to-one map of £/ onto Dr(X). Since # is
unique, the continuity of/implies that U =j^l(Dr{X)) is open. As g is also
continuous,/is indeed a homeomorphism. Finally,/(0) = €(0)E Dr(X) and so
OEM £/. •

From Theorem 4 it is a small step to the inverse-function theorem for


Banach spaces.
In defining differentiable maps, it is convenient to use the 'little oh'
notation. Let X and Y be Banach spaces. Given a function a : D —> Y,
where D C X, we write
a(h) = o(h)
if for every e > 0 there is a 8 > 0 such that

whenever h E D and ||/i|| < 5. Thus 'a(/t) tends to 0 faster than h\


Let U be an open subset of X and let / : U -> Y. We call / differenti-
able at x0 E [/ if there is a linear operator T E 26(Z, Y) such that

The operator T is the derivative of / at JC0: in notation, /'(* 0 ) = 7\


Thus / is differentiable at x0 with derivative T E ^(A^, Y) iff for every
e > 0 there is a 8 > 0 such that
\\f(xo + h)-f(xo)-Th\\^e\\h\\
whenever \\h\\ < S.
Note that a linear operator T E 26(Z, Y), * •-> Tx, is differentiable at
every point, and T'(x) = T for every x E X. We shall need the follow-
ing simple analogue of the mean-value theorem.
Lemma 5. Let X and Y be Banach spaces, U C X an open convex set
and / : U —> Y a differentiable function, with
106 Chapter 7: The contraction-mapping theorem

11/'toII ^*
for all x E U. Then
\\f(x)-f(y)\\^k\\x-y\\
for all x,y E U.
Proof. If the assertion is false then one can find two sequences
(xn)oAyn)o C U such that x0 f y0, and for n ^ 0 either xn+1 = xn and
OTX and and
yn + l = %(Xn+yn)> n +1 = &Xn+yn) ^ + 1 = yn>
\\f(xn)-f{yn)\\^k'\\xn-yn\\
for some k' > k and all n ^ 0. But then there is a z0 E £/ such that
xn-* Zo>yn~* zo and

as the segments [xn,yn] are nested and


Hence
k'\\xn-yn\\^\\f(xn)-f(yn)\\

^\\f'(zo)(xn-z0)\\+o(xn-z0)
+ \\f'(zo)(zo-yn)\\+o(z0-yn)
^ k\\xn-z0\\+k\\z0-yn\\+o(xn-yn)
= k\\xn-yn\\+o(xn-yn),
which is a contradiction, as ||* n -,y /l || ^ 0 and ||jc n -^|| -> 0. D

Like Theorem 4, the following theorem ensures that, under suitable


circumstances, a function / is a homeomorphism in some neighbour-
hood.
Theorem 6. Let X and Y be Banach spaces, let U be an open neigh-
bourhood of a point x0 E X and let / : U —> Y be such that f'(x) exists
and is continuous in U. Suppose

Then / is a homeomorphism of an open neighbourhood Uo of x0 onto


an open neighbourhood Vo of f(x0), f~l is continuously differentiable
on VQ and
Chapter 7: The contraction-mapping theorem 107

(rvoo = (rir\y))Yl
for y e Vo.
Proof. We may assume that JC0 = 0 and f(x0) = 0. Furthermore, setting
^o = /'(0) anc* replacing / by T^ofiX-* X, we may assume that
Y = X and /'(0) is the identity operator /.
Set e(jc) = f(x) —x. Then 6 : £/—> Y is continuously differentiable and
e'(0) = 0. Hence there is an open ball DS(X) C U such that
Ik'to II * i
for x E DjCY). Then, by Lemma 5,

for all x,y E DjCJf). Hence, by Theorem 4, /(*) = Jt + e(jc) is a homeo-


morphism of some open neighbourhood Uo of 0 onto Vo = Ds/4(X).
The rest is plain sailing. For y0 E Vo set JC0 = fHyo) E Uo and
T = /'(JC 0 ). Furthermore, for >>0 + A: E Fo set xo + h = f~\y^k). As
k -> 0, we have A -» 0, and so A: = Th + o(h) and

Consequently,
f~\yo + k) = r\f(xo + h)) = xo + /z =
proving that (/"VOto) = T~l = (f'(f~1(yo))) . The continuity of
C/" 1 )' follows from the fact that the map T •-» T~l is continuous. •

The contraction-mapping theorem has numerous applications to dif-


ferential and integral equations. Here we shall confine ourselves to
three rather simple examples.
Theorem 7. Let K(s, t) be a continuous real function on the unit square
[0,1] 2 , and let v{s) be a continuous real function on [0,1]. Then there
is a unique continuous real function y(s) on [0,1] such that

y(s) = v(s)+f K{s,t)y{t) dt.


Jo
Proof. We shall consider various operators on the Banach space C[0,1]
with the uniform (supremum) norm. The linear integral operator
L: C[0,1] -> C[0,1], defined by

(Lx)(s) = I* K(s,t)x(t) dt
108 Chapter 7: The contraction-mapping theorem

is called the Volterra integral operator with kernel K(s,f). In fact, we


could assume that K(s,t) is defined only on the closed triangle
0 ^ t ^ s ^ 1, since for s < t we do not use K(s,t). Write Ki(s,t) =
K(s, t) and for n ^ 1 set

Kn+I(s,t) = K(s, u)Kn(u,t)du.


Jt

Here and in the rest of the proof we take an integral

r
I f(u) du

to be 0 if t > s. Alternatively, we may take each Kn(s,t) to be 0 out-


side the closed triangle 0 ^ f ^ s ^ 1. It is easily shown by induction
on n that Ln is precisely the Volterra integral operator with kernel Kn.
Indeed, if this is true for n then

(Ln+1x)(s) = Jf K(s,u) JtU Kn(u,t)x(t) dtdu


o o

= f f1 K(s,u)Kn(u,t)x(t)dtdu
Jo Jo

= \ \ K{s,u)Kn{u,t)dux{t)dt
Jo Jt

= \ Kn+1(s,t)x(t) dt.
Jo
The function K(s, t) is continuous on the closed unit square and so it is
bounded, say

for all 0 ^ s, t ^ 1. Then, again by induction on n, we see that

lBr/ ,, Mn\s-t\n~l
lKMl
^ 0.-1)!
for all n ^ 1. Indeed, if this holds for n then

—7TT du

Mn+1\s-t\n
Chapter 7: The contraction-mapping theorem 109

Hence if n is sufficiently large, say n ^ 4M, then


\Kn(s,t)\ ^1
for all 0 ^ s, t ^ 1. Therefore

f Kn(sMx(t)\dt ^ \\x\\ f \Kn(s,t)\ d*


o
and so
|L-|| ^ h
In particular, Ln is a contraction.
It is time to return to our equation. Define a continuous (affine)
operator T: C[0,1] —> C[0,1] by 7* = v + Lx. The theorem claims
that T has precisely one fixed point. But
jn-l \

\i=0 /

and so Tn is a contraction. Hence, by Theorem 2, T has a unique fixed


point. •

The next application concerns a non-linear variant of the Volterra


integral operator.
Theorem 8. Let K(s,t) and w(s,t) be continuous real functions on the
unit square [0,1] 2 , and let v(s) be a continuous real function on [0,1].
Suppose that
\w(s9t1)-w(s9t2)\ ^A^ki-^l
for all 0 ^ tl9 t2, s ^ 1. Then there is a unique continuous real function
y(s) on [0,1] such that

y(s) = v(s)+ f K(s,t)w(t,y(t))dt.

Proof. Define L: C[0,1] -* C[0,1] by

(Lx)(s) = f K(s,t)w(t,x(t))dt.

The theorem claims that the map T: C[0,1] -> C[0,1], defined by
7JC = v + Lx,
has a unique fixed point.
110 Chapter 7: The contraction-mapping theorem

To show this, we shall turn to trickery. For a > 0, we introduce a


new norm || • ||a on C[0,1]:

I- f
" Jo
Then |||| a is indeed a norm on C[0,1], and it is equivalent to the Lx norm. Set
Xa = (C[0, 1], IhllJ and let Xa be the completion of Xa. Clearly, Xa is the
vector space L^O, 1) with the norm ||||fl, and L extends to a map L: Xa -> X a
given by the formula for L. Furthermore, with M = max{|X(s,r)|: 0 ^ 5, t ^ 1}
we have, for x,yEXa = (Lx(0, 1), |||| a ),

\\Lx -Ly\\a= \ K(s,t){w(t,x(t))-w(t,y(t))} dt ds


Jo o

e""|JC(O -y(t) \ dt ds

= MN \

a
This shows that for a > MN the map
L:Xa-+Xa
is a contraction, and so is T = v + T. It is easily checked that f maps Xfl into
Xa = (C[0,1], ||• || J, so its uniquefixedpoint belongs to C[0,1], and is also the
unique fixed point of T. •
In our final example we have to do even less work, since the kernel
satisfies a Lipschitz-type condition, just like the function w(s,t) in
Theorem 8.
Theorem 9. Let K(s,t,u) be a continuous function on 0 ^ s,t ^ 1,
u 25 0, such that
\K(x,t,u1)-K(s,t,u2)\ ^N{
where N(s,f) is a continuous function satisfying
f N(s,t)dt^k<l,
Jn
Chapter 7: The contraction-mapping theorem 111

for every 0 ^ s ^ 1. Then for every v E C[0,1] there is a unique func-


tion y E C[0,1] such that

y(s) = v(s)+ ( K(s,t,y(t))dt.


Jo
Proof. Define L: C[0,1] -> C[0,1] by
rl
(Lx)(s) = K(s,t,x(t))dt,
JQ

so that a solution of our integral equation is precisely a fixed point of


the map T: C[0,1] -> C[0,1], given by 7JC = t; + Lx. For
*,)> E C[0,l] we have

f
^ Jf N(s,t)\x(t)-y(t)\dt
o
*k\\x-y\l
Hence \\Lx — Ly\\ ^ A:||JC—y\\, implying that L is a contraction and there-
fore so is T. Consequently, by Theorem 1, T has a unique fixed point.

As we remarked earlier, in chapter 15 we shall prove several other
fixed-point theorems. But for the time being we return to the study of
normed spaces.

Exercises

1. Show that a contraction of an incomplete metric space into itself


need not have a fixed point.
2. Let A and B be disjoint subsets of a metric space, with A compact
and B closed. Show that
d(A,B) = inf{d(x,y): x E A,y E B) > 0.
Show also that this need not be true if we assume only that A and
B are closed.
3. Show that if / is a map of a complete metric space X into itself
such that d(f(x),f(y)) < d(x,y) for all x,y E X (x + y) then /
112 Chapter 7: The contraction-mapping theorem

need not have a fixed point. Show however that if X is compact


then / has a unique fixed point.
4. Let / : Z ^ Z b e a map from a complete metric space into itself.
Suppose that for every € > 0 there is a 3 > e such that if
e ^ d(x,y) < 8 then d(f(x),f(y)) < e. Prove that / has a unique
fixed point.
5. Let / be a map of a compact metric space X into itself. Suppose
that for every e > 0 there is a 8 = 8(e) > 0 such that if
d(x,f(x)) < 8 then f(B(x,e)) C B(x,e). Let x0 E X and define
(xn)o as in the proof of Theorem 1: xn = f(xn-i) for n ^ 1. Show
that if d(xn,xn+i) —> 0 as n —> °° then the sequence (jtn)o con-
verges to a fixed point of / .
6. Let / be a map of a complete metric space X into itself such that
d{f{x)J{y)) ^ <p(d(x,y))
for all x,y E X, where cp : R —* R+ is a monotone increasing
+

function such that limn^>00 <pn(f) = 0 for every t > 0. Deduce from
the result in the previous exercise that / has a unique fixed point
a, and lim^oo fn(x) = a for every JC E X
7. Let / be a continuous map of a (non-empty) compact Hausdorff
space into itself. Show that there is a (non-empty) compact subset
such that f(A) = A.
8. Show that on the space c0 the norm function x = (jt,-)" •-*
||JC|| = sup|jr,-l is differentiable at x iff |JC,-| has a unique maximum
(i.e. there is an index j such that |JC;| > |JC,-| if / ^ / ) .
9. Show that on the space lx the norm function

* = to)r-> Ml = 2 k l
1=1

is not differentiable at any point.


In the next five exercises, X, Y and Z are Banach spaces.

10. Let f,g: X - * Y be differentiable at JC0 E X and let a and /3 be


constants. Show that af+Pg is differentiable at x0 and
(af+pg)'(xo) =af'(xo)+pg'(xo).
11. Let / : [/ —» V be a homeomorphism of an open set U C X onto an
open set V C Y. Suppose that / is differentiable at x0 E U and
))'1 e a(Y,-Y). Show that / - 1 : F ^ [/ is differentiable at
Chapter 7: The contraction-mapping theorem 113

and

12. Let f / C Z a n d VC Y be open sets, a n d / : £/^> y and g : V ^ > Z


continuous maps with f(U) C V. Suppose that / is differentiable
at a point x0 E U and g is differentiable at y0 = /(JC0) E V. Show
that the map g°f: £/ -» Z is differentiable at JC0 and

13. Let U be a connected open subset of X and, for n = 1,2,..., let


fn : [/ —• y be differentiable at every point of U. Suppose that
every x0 E U has a neighbourhood JV(*o) o n which S^ = 1 H/nWll i s
uniformly convergent. Show that if 2° C =1 fn(x) ^s convergent for
some xEU then it is convergent for all xEU, the map

/M = £ /„(*)
n=l
is differentiable at every xEU and

/'(*) = ni= \ fn(x).


14. Let U d X bt an open convex set and /:£/—> y differentiable.
Show that for all jc,y, z E [/ we have
\\f(x)-f(y)-f'(z)(x-y)\\ ^ \\x-y\\ sup J/'«-/'(*)II,
s(E[x,y]
where, as usual, [x,y] is the closed segment {tx + (\ — t)y : 0 ^

Notes

The contraction-mapping theorem is from S. Banach, Sur les operations


dans les ensembles abstraits et leur applications aux equations integrates,
Fundamenta Mathematica, 3 (1922), 133-81. Of course, it can be found
in most books on general topology and basic functional analysis. A rich
source of applications to differential and integral equations is D. H.
Griffel, Applied Functional Analysis, Ellis Harwood, Chichester, 1985,
390 pp.
8. WEAK TOPOLOGIES AND DUALITY

When studying a normed space or various sets of linear operators on a


space, it is often useful to consider topologies other than the norm
topology we have introduced so far. Two of the most important of
these topologies are the weak topology and the weak-star topology.
Before we define these topologies and prove Alaoglu's theorem, the
fundamental result concerning the weak-star topology, we shall review
some additional concepts in point set topology and present a proof of
Tychonov's theorem.
Given a topological space (X,T), a collection or of subsets of X is said
to be a sub-basis for r if a C r and every member of r is a union of
finite intersections of sets from a. Equivalently, a is a sub-basis for r if
a C T and for all U E r and x E U there are sets S1,..., Sn such that
x E n" = 1 Sf C U. More formally, a is a sub-basis for T if

(1)
F(E& J
where cr(<a>) is the set of all finite subsets of a-. In most cases we may
omit the second term on the right, {0,X}, especially if the intersection
of an empty collection of subsets of X is taken to be X. Not only do
sub-bases tend to be more economical than bases, but they also enable
us to use an arbitrary family of sets to define a topology. Indeed, if
a C d*{X) is any collection of subsets of X then the collection r defined
by (1) is a topology on X.
The possibility of using an arbitrary set system to define a topology
enables us to use a set of functions into topological spaces to define a
topology on a set. Let A" be a set and for each y E F let fy be a map
of X into a topological space (Xy9ry). Then there is a unique weakest

114
Chapter 8: Weak topologies and duality 115

topology T o n l such that every fy : (X,T) -> (Xy,Ty) is continuous.


Indeed,taking
<r = {fy\Uy) :UyCXy is open in ry (y E D},
the topology r is given by (1). Thus U C X is open in r iff for every
x E U there are y x , . . . , yn E F and Ux E r7l , . . . , ( / „ E r^, such that

1=1

The topology r defined in this way is called the weak topology generated
by 9, and is denoted by a(X, 9), where 9 = {fy : y E F\.
A map / from a topological space (X',T') into (X,a(X, 3F)) is continu-
ous iff fy°f: (X',r') -> (Xy,Ty) is continuous for every y E F. This so-
called universal property characterizes the weak topology.
The product topology is one of the prime examples of a weak topol-
ogy. Let (Xy,Ty) be a topological space for each yEF and let
X = n E r Xy be the Cartesian product of the sets Xy (y E F). Thus
X is the collection of all the functions x: F -> (J e r Xy such that
x(y) E Xy. One usually writes JCT for x(y) and sets x = ( x y ) y e r ; also,
x y is the y-component of x. Let py\ X^> Xy be the projection onto
A^; thus /?y(x) = xy. The product topology on ^ is the weak topology
generated by the projections py (y E F). To spell it out: a set U C X is
open in the product topology iff for every point x = (xy)ysr there are
indices yi,...,%, and open sets Uyi,...,Uyn (£/y. E ry.) such that
x y E £7y (i = 1 , . . . , n) and

E
n Py^Wy) = {y = (yy) £X'-yyi uyt a = 1,...,«)} c t/.

Thus to guarantee that a point y = (y r ) belongs to a fixed neighbour-


hood of x = (Xy), it suffices to demand that for finitely many indices
Ti > 72 > • • • > 7n > e a c h ^y, belongs to a certain neighbourhood of xy .
Let us turn to the main topic of this section, the weak and weak-star
topologies of normed spaces. Given a normed space X with dual X*,
the weak topology a(X, X*) is the weak topology on X generated by the
bounded linear functional on X. Thus U C X is open in the weak
topology (tv-open) iff for every x E U there are bounded linear func-
tionals / x , . . . , / „ and positive reals e x , . . . , en, such that
{y E X: |/-(x)-/;O0| < et (i = 1,...,»)} C [/.
By replacing /} by fje^ we may take every ef to be 1.
116 Chapter 8: Weak topologies and duality

The weak-star topology a(X*,X) on the dual X* of a normed space X


is the weak topology generated by the elements x E X** for x E X, i.e.
by the elements of X acting on X* as bounded linear functionals (see
Theorem 3.10). In more detail: a set G C X* is open in the weak-star
topology (w*-open) iff for every g E G there are points x1,...,xn EX
and positive reals €1,...,en, such that
{fex*: !/(*,-)-*(*,.)I <€,.(! = I,...,*)}CG.

As before, we may take every e, to be 1.


The Hahn-Banach theorem implies that if Y is a subspace of a
normed space X then the restriction to Y of the weak topology on X is
precisely the weak topology on Y9 i.e.
*(X9X*)\Y=v(Y,Y*). (2)
In view of the proof of Tychonov's theorem we shall give, it seems to
be appropriate to mention Tukey's lemma, which is equivalent to Zorn's
lemma (and so to the axiom of choice). A system 9 of subsets of a set
S is said to be of finite character if a subset A of 5 belongs to-9 if and
only if every finite subset of A belongs to 9.
Lemma 1. (Tukey's lemma) Let 9 be a set system of finite character
and let F E 9. Then 9 has a maximal element containing F.
Proof. Let SF0 = {A E 9: F C A}. Order the elements of 90 by inclu-
sion: A =^ B iff A C B. Then ^ is a partial order on 90. Also, if
% C 9^ is a chain in 8F0 then Z> = \JCG% C belongs to SFQ s i n c e all
finite subsets of D belong to 9, and F C D. Clearly, D E 90 is an
upper bound for %. Hence, by Zorn's lemma, 90 has a maximal ele-
ment. •

A system 9 of subsets of a set has the finite-intersection property if


fl" = 1 Ft? 0 for all F1,...,Fn E 9. It is easily seen that a topological
space X is compact iff the intersection of all subsets of a system 9 of
closed sets which has the finite-intersection property is non-empty.
Indeed, ^ = {X\F: F E 9} is a collection of open sets. Since no finite
collection of sets in ^ covers X, the entire system ^ also fails to cover
X, so there is a point x E X which belongs to no member of (S. Hence

A trivial reformulation of this characterization of compactness is the


following; a topological space is compact iff whenever si is a system of
Chapter 8: Weak topologies and duality 111

subsets with the finite-intersection property then f}AG^ A ^ 0. This is


the characterization we shall use in the proof below.
Theorem 2. (Tychonov's theorem) The product of a collection of com-
pact sets is compact.
Proof. Let {Xy : y E / ] be a collection of compact spaces and let
X = n e r Xy be endowed with the product topology. We have to
show that X is compact, i.e. that if si is a system of subsets of X with
the finite-intersection property then HAGs* A ^ 0.
Let SF be the collection of all set systems on X with the finite-
intersection property. Then 9 is of finite character and so, by Tukey's
lemma, there is a maximal set system SS with the finite-intersection pro-
perty containing si. Since
n BC n i,
we may assume that our set system si is itself is maximal for the finite-
intersection property.
The maximality of si implies that the intersection of any two sets in si
belongs to si, as does the intersection of finitely many sets in si. Conse-
quently, if B C X meets every set in si then B E si. In particular, if
AEsi and ACBCX then B E si.
So far we have made no use of the fact that X is a product space, let
alone a product of compact spaces. Let us do so now. For each y £ f ,
the system {py(A): A E si} of subsets of Xy has the finite-intersection
property. As Xy is compact, for some point xy E Xy we have
xy E PI Py(A).

This means precisely that for every neighbourhood Uy of xy, the set
Py\Uy) meets every element of si. Hence p~l{Uy) E si and so

whenever {yt,..., yn} C f and Uy. is a neighbourhood of xy..


Let x = (*: r ) re r E x- W e c l a i m t h at x E HAGsl A. Indeed, for
every neighbourhood U of x there are indices yx,..., yn E F and neigh-
bourhoods Uy. ofxt (i = l,...,n) such that

n P-\uy) c u.
i l
118 Chapter 8: Weak topologies and duality

Consequently, (3) implies that U E si, i.e. U intersects every element of


si. Hence every neighbourhood of x intersects every A E si, and so
x E A for every A E si. This proves our claim and completes the proof
of our theorem. •

From here it is a small step to Alaoglu's theorem, the fundamental


result concerning the weak-star topology; in view of the fact that both
the weak-star topology and the product topology are weak topologies,
this is hardly surprising.
Theorem 3. (Alaoglu's theorem) The unit ball B(X*) of the dual of a
normed space X is compact in the weak-star topology.
Proof. For x E X, set Dx = {£ E R: |£| ^ ||JC||} if X is a real space and
Dx = {£ E C: \i;\ *z \\x\\} if X is a complex space. Furthermore, let D
be the product [] _ v Dx endowed with the product topology. Since
Dx, a closed ball of the scalar field, is compact, Tychonov's theorem
implies that D is a compact space.
Let us write B* for B(X*) endowed with the weak-star topology, and
define a map cp : B* -> D by setting, for/ E B*,
<p(f) = (/(*)
The definitions of the product topology and the weak-star topology
imply that <p is a continuous map of B* into D, and it is clear that cp is
one-to-one. Furthermore, again by the definitions of the product topol-
ogy and the weak-star topology, <p~x is a continuous map from <p(B*)
onto B*. Thus B* is homeomorphic to the subset (p(B*) of D. There-
fore, to complete the proof, it suffices to show that <p(B*) is a closed
subset of the compact space D.
Let then £ = (£x)xex E D be in the closure of <p(B*). Define a scalar
function / on X by setting f(x) = gx for x EX. We claim that / is a
linear functional, i.e. f E X'. To see this, let x,y E X and let a and )3
be scalars. Since £ E <p(B*), for every n G N there is a continuous
linear functional fn E B* such that

l/W-/«W| + l/<3')-/«O')l + l/(ax+i8y)-/(,(«x+^)| < i . (4)


As fn(ax+fiy) = otfn(x)+/3fn(y), inequality (4) implies that
/(ox + jSy) =a/(*)+/3/(3>),
i.e. / E JT. Since |/(*)| = | £ j ^ ||x|| for every x E X, f is not only a
Chapter 8: Weak topologies and duality 119

linear functional, but it is also continuous and ||/|| ^ 1, i.e. / E B*.


Finally, by the definition of / , we have £ = (p(f) E <p(B*)9 completing
the proof. •

As an immediate consequence of Alaoglu's theorem, we can see that


a subspace of the Banach space of continuous functions on a compact
Hausdorff space is the most general form of a normed space.
Theorem 4. Every normed space X is isometrically isomorphic to a sub-
space of C(K), the Banach space of continuous functions on K, where K
is B(X*) endowed with the weak-star topology. In particular, every
normed space is isometrically isomorphic to a subspace of C(K) for
some compact Hausdorff space K.
Proof. Let X and K be as in the theorem; thus AT is a compact Haus-
dorff space, namely B(X*) endowed with the weak-star topology. For
x E X let fx be the restriction of x E X** (defined by x(x*) = x*(x) for
x* E X*) to K = B(X*). By the definition of the weak-star topology,
fx E C(K); indeed, the weak-star topology on X* is precisely the weak-
est topology in which every function x is continuous.
The map x »-> fx of A' into C{K) is clearly linear. Furthermore, by
the Hahn-Banach theorem,

ll/xll =
= sup{|jt%t)|:jc*E

and so the map X —> C{K) is a linear isometry. •

It is customary to write w-topology for the weak topology and w*-


topology for the weak-star topology. Similarly, one may speak of w-
open sets, w*-closures, w-neighbourhoods, and so on, all taken in the
appropriate spaces. Thus a H>-open subset of X* is a a(X*,X**)-open
subset, and the w*-closure of a subset of X** is the closure of the set in
the <7(Ar**,Ar*)-topology. As before, we consider l a s a subspace of
X**, the isometric embedding being given by JC^-^JC, where (x,x*) =
<**,*> for all** EX*.
In order to avoid some inessential complications, for the rest of the
section we shall consider only real spaces.
The weak topology is weaker than the norm topology, so every
w-closed set is also norm closed. The following result claims that for
convex sets the converse is also true.
120 Chapter 8: Weak topologies and duality

Theorem 5. A convex set C of a normed space X is closed iff it is w~


closed.
Proof. Suppose that C is closed and x0 ^ C. Then d — d(x0, C) > 0.
Set D = {x E X: d{x$ ,x) < d\. By the separation theorem (Theorem
3.13) there is a bounded linear functional x* E X* such that
s = sup(;t*,;t) ^ inf (x*,x).
*ec X<ED

Then
U= {xEX: (x*,x) >s}
is a w-neighbourhood of x0 and UH C = 0 . This shows that C is also
w-closed. As we remarked earlier, the converse is trivial. •

Theorem 6. B(X**) is the >v*-closure of B(X) in X**.


Proof. It is easily checked that the closed unit ball B(X**) is w*-closed.
Let B be the w*-closure of B(X). Since B(X**) is w*-closed and
B(X) C B(X**), we have B C B(X**). Furthermore, as the w*-closure
of a convex set is convex, B is a convex set which is closed in the norm
topology of X**.
Suppose that, contrary to the assertion of the theorem, B ± B(X**).
Then there is a point XQ* E B(X**)\B SO that, by the separation
theorem (Theorem 3.13) there is a bounded linear functional xfi** on
X** such that

for all b E B. Let JCQ be the restriction of xfi** to the subspace X of


X**. Then

for all x E B(X) and so ||JCO|| ^ 1, contradicting

The results above have some beautiful consequences concerning


reflexivity. Note that the w-topology on a normed space X is precisely
the restriction to X of the w*-topology on X**.
Theorem 7. A normed space X is reflexive iff B(X) is w-compact.
Proof. By Theorems 3 and 6, B(X) is a w*-dense subset of the w*-
compact set B(X**). Since a subset of a compact Hausdorff space is
Chapter 8: Weak topologies and duality 121

compact iff it is closed, B(X) is a w*-compact subset of X** iff


B{X) = B(X**). But this means precisely that B{X) is w-compact iff
X = X**. •

The next result can also be deduced from the Hahn-Banach theorem;
now we are well equipped to give a very straightforward proof.
Corollary 8. A closed subspace of a reflexive space is reflexive.
Proof. Let 7 be a closed subspace of a reflexive space X. Since
B(Y) = B(X)DY, the norm-closed convex set B(Y) C X is a(X,X*)-
compact. But, as we remarked in (2), the restriction of cr(X,X*) to Y is
precisely a(Y, Y*). Hence B(Y) is w-compact and so, by Theorem 7, Y
is reflexive. •

Theorem 9. Let A" be a Banach space. The following assertions are


equivalent:
(i) X is reflexive;
(ii) a(X*,X) = a(X*,X**), i.e. on X* the w-topology and the w*-
topology coincide;
(iii) X* is reflexive.
Proof. The implication (i) => (ii) is trivial (and so is (i) => (iii)). Sup-
pose that (ii) holds. Since B(X*) is w*-compact, it is also w-compact.
Hence, by Theorem 7, A'* is reflexive. Thus (ii) => (iii).
Finally, suppose that (iii) holds. As X is a Banach space, the ball
B(X) is norm closed in X**. Therefore, by Theorem 5, B{X) is w-
closed, i.e. e r O ^ A ^ ^ - c l o s e d . Since X*** = X*, this means that
B(X) is w*-closed in X**. But, by Theorem 6, the w*-closure of B(X)
is B(X**), so B(X) = B(X**). Hence (iii) => (i). •

If X if reflexive and / E X* then the function |/|, being w*-


continuous on B(X), attains its supremum there: for some x0 = B(X)
we have
Il/H =sup{\f(x)\:xEB(X)} = \f(xo)\.
Equivalently, there is an x0 E B(X) such that ||/|| = f(x0). (Of course,
if / ^ 0 then x0 is not only in the closed unit ball but also in the closed
unit sphere S(X).) It is a rather deep theorem of R. C. James that this
property characterizes reflexive spaces: a Banach space X is reflexive iff
every bounded linear functional attains its norm on B{X).
122 Chapter 8: Weak topologies and duality

Our next aim is to show that every Banach space comes rather close
to having this property: the functionals which attain their suprema on
the unit ball are dense in the dual space. (This property used to be
called subreflexivity. As the result claims that every Banach space is
subreflexive, the term has gone out of use.) In fact, we shall prove
somewhat more.
Given a Banach space X, define
n(X) = {(*,/): x E S(X), f E S(X*)9 f(x) = 1}.
Then II(X) C XxX* and, denoting the projection of XxX* onto X by
px and onto X* by px*, we have px(II(X)) = S(X). Furthermore,
Px*(IJ(X)) is precisely the set of functionals in S(X*) which attain their
norms on S(X).
For 8 > 0 let us write II8(X) for the set of pairs (x,f) E S(X)xS(X*)
such that / is almost 1 at x:

We shall show that not only is px*(n(X)) dense in S(X*), which is pre-
cisely the subreflexivity of X, but also every point (x,f) of IIS(X) is
close to some point (y,g) of II(X):

\\x-y\\ + \\f-g\\<<p(fi), (5)


where <p(8) -» 0 as 8 -> 0.
The proof of this theorem is based on the following lemma, whose
proof is left to the reader (see Exercise 13).
Lemma 10. Let X be a real normed space and let f,g E S(X*) be such
that

whenever x E B(X) and f(x) = 0, where 0 < e < \. Then either


||/-g||*26or||/+*||*2e. •

Theorem 11. (Bishop-Phelps-Bollobas theorem) Let X be a Banach


space and let JC0 E S(X) and / 0 E S(X*) be such that

where 0 < e < 1. Then there exist xx E 5(^) and /i E 5(Z*) such that
1, II/0-/1II * 6 and ||*o-*ill < 6.
Chapter 8: Weak topologies and duality 123

Proof. In the notation introduced before Lemma 10, the theorem claims
that for every (JCO,/O) E TI8{X) there is an (*i,/i) e II(X) such that
ll/o~/ill ^ € anc* ||*o~*ill < ^ ; in particular, (5) holds with <p(§) =
45 1/2 , say.
As linear functionals are determined by their real parts and the map
/ »-> R e / is an isometry, we may assume that X is a real space.
Let

and define a partial order ^ on B = B(X) by setting x ^ y if


\\x-y\\^kfo(y-x).
This is indeed a partial order, and a simple application of Zorn's lemma
(see Exercise 16) shows that the set Bo = {x E B: JC0 ^ *} has a maxi-
mal element, say xt. Since x0 ^ xx,

Il-Ko"*!!' ^ kfo(xi—xo) ^ k8

Set

c-2-

and let D be the convex hull of B U C:


D = co(BUD) = {ty + il-t)z:y E BUC, z E BUC,0 ^ t ^ 1}
[-t)z : y E B, z E C,0^ t^l}.
Then £> is a centrally symmetric convex set containing B and C.
We claim that jq £ IntZ). Indeed, otherwise there a r e O < $ < f < l ,
y\E B and ^ e C such that

Since fo(xo) > 0 and JC ^ y implies that /0(JC) ^ fo(y), we see that

/o(*o) ^ /o(^i) = sfoiyi) < foiyi)


and so
fo(yi~*i) = (l-s)foiyi) > il-s)foixo) > 0. (6)
Furthermore,
124 Chapter 8: Weak topologies and duality

y1-x1 = (l
and so

lbi-*ill ^ (ls) + (t-s)l < (l-5)(l + ^). (7)

Inequalities (6) and (7) give

\\y1-x1\\ ^ ( l - s ) ( l + - ) = 777^(l-5)/o(*b) * kfo(yi-Xl),

showing that xx ^ yx. But, as xx is maximal, we have xx = yl9 contrad-


icting (6), say. This proves that JCX £ IntD, as claimed.
By the separation theorem (Theorem 3.13) there is a linear functional
/i E X* such that/jUi) = 1 and/xOt) ^ 1 for all x E D. Then H/J = 1
and /i(jc) ^ 1 for all x E C, and so fx(x) ^ ^6 for all x G fin £eKer/ 0 .
Hence, by Lemma 10,
ll/o-/ill<« or H/0 + /1II * €.
But

and so we have ||/o-/il| ^ €. D

In Chapter 12 we shall give some applications of Theorem 11 to


numerical ranges. Our last aim in this chapter is to prove another
major result of functional analysis, the Krein-Milman theorem.
Although in its proper generality this result has nothing to do with dual
spaces and weak topologies, we present it here since it has some natural
applications concerning dual spaces.
Our proof of the Krein-Milman theorem is based on another
equivalent form of the axiom of choice, namely on the well-ordering
principle, enabling us to apply transfinite induction.
An ordered set is a set with a linear (total) order. An ordered set
(5, ^) is said to be well ordered if every (non-empty) subset T of 5 has a
smallest element, i.e. an element t0E T such that t0 ^ t for all t E T.
Lemma 12. (Well-ordering principle) Every set can be well ordered.
Proof. A subset B of an ordered set {A, ^) is said to be an initial seg-
ment if a E A, b E B and a ^ b imply that a E B.
Let S be a set and let P = {(Sy, ^y) : y £ / ] be the set of all well
ordered sets such that Sy is a subset of 5. For (Sy,^y) and (S 5 , ^ 5 ) in
Chapter 8: Weak topologies and duality 125

Pset

if Sy is an initial segment of Ss and on Sy the two orderings coincide.


Clearly (P, ^) is a partially ordered set.
It is easily seen that in (F, *s) every chain has an upper bound.
Indeed, let {(Sy, ^y): y £ r 0 } be a chain. Set 5' = U y G r o $y and for
a, b E S' set a ^f b if a, b E Sy and a ^y b for some y E T o . Then ^ '
is an order on 5'. Furthermore, if T C S' and f' E T then f' E Sy for
some y E r , and the smallest element of 7"nS r in (Sy,^y) is also a
smallest element of T in (5, ^ ' ) -
Since every chain has an upper bound, by Zorn's lemma from
Chapter 3, (P, ^) has a maximal element (S o ,^o)- ^-° c o m p l e t e the
proof, all we have to show is that So = 5. Suppose that this is not so,
say sx E S\S0. Extend the order ^ 0 t o a n order ^x on Sx = SoUfrx} by
setting s ^i Si for all s E 5 0 . Then (S1,^1) is a well ordered set and
(So, ^o) < (^i> ^i)> contradicting the maximality of (5 0 , ^ 0 ) . D

Let K be a convex subset of a vector space. A point in K is said to


be an extreme point of K if it is not in the interior of any line segment
entirely in K. Thus a E K is an extreme point of K if whenever
b,c E K, 0 < t < I and a = tb + (l-t)c, we have a = b = c. The set
of extreme points of K is denoted by Ext K.
Recall from chapter 3 that the convex hull of a subset 5 of a vector
space V is

co5= 2 ^ : ^ 6 S , ^ 0 ( / = 1,...,/!), £ *i = 1(« = 1,2,...)


Ui 1=1

If 5 is in a normed space then co 5, the closed convex hull of 5, is the


closure of coS; by Theorem 3.14, coS is the intersection of all closed
half-spaces containing 5.
In its full generality, the Krein-Milman theorem concerns locally con-
vex topological vector spaces. As we do not study these spaces, we
shall be satisfied with a somewhat artificial form of this result. Given a
vector space V with dual V, for F C V and S C V the F-closed convex
hull of 5 is

co F 5 = 0 {x E V: f(x) ^ sup/(>>)}.
/(EF yes
For the rest of the chapter, we say that 5 is F-convex if co F 5 = 5. By
126 Chapter 8: Weak topologies and duality

Theorem 3.14, if A" is a normed space then coS = c o ^ S for all 5 C X;


a set X is convex and closed (in the norm topology) iff it is X*-convex.
Theorem 13. (Krein-Milman theorem) Let r be a Hausdorff topology
on a vector space V and let FdV. Suppose that F separates the
points of V (i.e. if f{v) = 0 for all / E F then v = 0) and each / E F is
r-continuous. Let AT be a non-empty T-compact F-convex subset of V.
Then Ext K ? 0 and K = c6FExtK.
Proof. Let {fy : y E 1} be a well ordering of F. Then there is a func-
tion 5 : F —> R, y »-* 5 y , such that for all y E T we have
sy = max{/r(*) :xEK, f8(x) = 55 for 8<y}. (8)
This is immediate from the fact that F (i.e. F) is well ordered and if ss
is determined for all S E T preceding y EL F then

d<y

is a non-empty compact set on which the continuous function fy attains


its supremum, sy.
The existence of the function s implies that there is a point a E K
such that fy(a) = sy for all y E F. This point a is an extreme point of
K. Indeed, if a = tb + (l-t)c for some 0 < t < 1 and b,c E K, then
(8) implies that fy(a) = / r (6) = / r (c) for all y e f . As F separates the
points of V, we have a = b = c. Thus E x t # ^ 0 .
Set # ' = coFExt K. Then K' C K since # is F-convex. To complete
the proof, we have to show that, in fact, K' — K. Suppose that this is
not so. Then there is a functional /o E F such that
max fo(x) < max fo(x). (9)
Jt€E/£ ^JK

Choosing a well ordering of F in which /o is the first element, the point


a produced by the procedure above is such that fo(a) = maxxEKf0(x).
But a E K\ contradicting (9). •

Corollary 14. Let K be a compact convex subset of a Banach space.


Then/: = co Ext K. •

Corollary 15. Let X be a normed space. Then B(X*) = c6ExtB(X*).


Proof. By Alaoglu's theorem, B(X*) is w*-compact. Furthermore, it is
A'-convex since if / 0 E X* and / 0 £ B(X*) then fo(xo) > 1 for some
JC0 E 5(Z) and f(x0) ^ 1 for all x E B(X). D
Chapter 8: Weak topologies and duality 127

The last result implies that B(X*) has 'many" extreme points:
sufficiently many to guarantee a 'large" closed convex hull, namely
B(X*). Occasionally this fact is sufficient to ensure that a given space is
not a dual space. For example, it is easily seen that ExtB(c0) = 0 ; so
c0 is not a dual space. Indeed, let x = (xn)™ E B(c0). Then \xk\ < \
for some k and so, with yn = zn = xn for n ^ k, yk = xk + \ and
*k = **~2> w e h a v e }> = 0^)1% z = (*»)" E s ( c o) and AT = § 0 + z), so
that x £ Ext5(c 0 ). Thus Ext 2?(c0) is indeed empty.
The Krein-Milman theorem is one of the fundamental results of func-
tional analysis, leading to deep theorems in operator theory and abstract
harmonic analysis, like the Gelfand-Naimark theorem concerning
irreducible *-representations of C*-algebras, and the Gelfand-Raikov
theorem about unitary representations of locally compact groups. But
these results will not be presented in this book. We shall return to com-
pact convex sets in Chapters 13-16, when we study compact operators
and fixed-point theorems. However, our next aim is to study the 'nicest'
Banach spaces, namely Hilbert spaces.

Exercises

1. Prove identity (2), i.e. show that if Y is a subspace of a normed


space X then the weak topology on Y is precisely the topology
induced on Y by the weak topology on X.
2. Let Y be a closed subspace of a reflexive space X, and let x0 E X.
Show that the function x »-> ||JC—JCO|| is weakly lower semicontinu-
ous, i.e. it is lower semicontinuous in the w-topology on X.
Deduce that the distance of JC0 from Y is attained, i.e. there is a
point y 0 E 7 such that
Ik-yoli = d(xo, Y) = inf{||*b-;y||: y e Y).
3. Show that if Y is a closed subspace of a reflexive space X then
X/Y is reflexive.
4. Show that

d(x,y)= i 2-»\xn-yn\
n= \

defines a metric on /«,, and the restriction of this metric to #(/«,)


induces the weak-star topology on #(/a>).
5. Let X be an infinite-dimensional normed space. Show that the w-
closure of S{X) = {xEX: \\x\\ = 1} is B{X) = {x G X: \\x\\ ^ 1}.
128 Chapter 8: Weak topologies and duality

6. Let (xn) be a sequence in lp (1 < p < oo) weakly converging to


x0 E lp (thus xn -> x 0 in the w-topology, i.e. (xn,x*) -» (xo,x*)
for every x* E l£ = lq). Prove that if | | x j | -> ||xo|| then
lk-*oll-*o.
7. Show that a sequence (*„) C lp (1 < /? < oo) converges weakly to 0
iff (xn) is bounded (i.e. ||JCJ| ^ K for some K and all n) and
lim^oo jc/n) = 0 for every i, where xn = (x[n), JC2(W) , . . . ) •
8. Let xn = (*i(fl),*2(lf)»---) € / ! ( » = 1,2,...). Show that (xn,y) -> 0
for every y £ c 0 iff (x,,)" is bounded and limn_>oo JC/W) = 0 for every i.
9. Formulate and prove assertions analogous to the previous two
exercises for the spaces Lp(0,1) and Lq(0,1).
10. Let X be an infinite-dimensional normed space. Show that the
norm function ||JC|| (x E X) is not weakly continuous at any point
of JT.
11. Let (en)i be the standard basis of l2 and let
A = {em + m e n : l ^ m < n } .
Prove that 0 is in the w -closure of A but no sequence in A con-
verges weakly to 0.
12. Let X be a Banach space, / E X* and M = Ker/. Prove that for
every x E X there is a point of M nearest to x iff / attains its
norm on B(X).
13. Prove Lemma 10: if X is a real normed space and f,g E S(X*) are
such that (Kerf)nB(X) C g'^-e^] for some 0 < e < \, then
\\f-g\\ ^ 2e or ||/+g|| ^ 2e. [HINT: If f ^ ±g then there are
fl,frel such that f(a) = g(6) = 1 and /(£) = g(a) = 0. Set M =
lin{a,b} and iV = Ker/flKerg, and denote by D the projection of
B(X) into M parallel to N. Let || • ||' be the norm on M with unit
ball D, and note that with f0 = f\M and g0 = g\M we have
ll«/o + /3goll' = l|a/+/3sll for all a,/3 E R. This shows that it
suffices to prove the lemma in the case when X = M, i.e. when X
is 2-dimensional. Use a simple geometric argument to complete
the proof.]
14. Show that the assertion of Lemma 10 is best possible: for
0 < e < \, there are functional f,g E S(X*) such that
(KtTf)nB(X) C g - ^ - ^ e ] , ||/+g|| ^ 1 and | | / - g | | = 2e.
15. Show that Zorn's lemma indeed implies the existence of a maximal
element xx ^ x0 in the proof of Theorem 11.
16. Check that Theorem 11 can be strengthened a little: if
2
| / 0 ( J C 0 ) - 1 | < je , where 0 < € < \ then we may demand
Chapter 8: Weak topologies and duality 129

ll/o~/ill ^ € a n d ll*o~*ill < e + e 2 . (This variant of the result is


essentially best possible.)
17. Let K C Un be a compact convex set. Show that if n = 2 then
ExtK is compact, but this need not be the case if n ^ 3.
18. Show that B(lt) = coExtB(l t ).
19. What are the extreme points of CR[0,1]?
20. Let L be a locally compact Hausdorff space, which is not compact.
Show that the closed convex set B(C^(L)) has no extreme points.
21. Let X be an infinite-dimensional Banach space whose unit ball has
only finitely many extreme points. Show that X is not a dual
space, i.e. there is no normed space Y such that X = Y*.
22. Show that Cu[0,1] is not a dual space.

Notes

Tychonov's theorem is from A. Tychonoff, Uber ein Funktionenraum,


Mathematische Annalen, 111 (1935), 762-66; somewhat earlier, an
incomplete version of the result appeared in A. Tychonoff, Uber die topo-
logische Erweiterung von Rdumen, Mathematische Annalen, 102 (1929),
544-61. (The modern transliteration of the name Tychonov' is closer to
the Russian original than the one used in German before the second world
war. Similarly, the old-fashioned usage of Markoff, Tschebischeff, Egor-
off, etc., is due to German influence.) A proof of the thoerem can be
found in almost any good book on general topology, for example, in J. L.
Kelley, General Topology, Van Nostrand, Princeton, N.J., 1955, xiv +
298 pp.
Theorem 3 is from L. Alaoglu, Weak topologies of normed linear spaces,
Annals of Math., 41 (1940), 252-67; in fact, the result was announced two
years earlier in Leonidas Alaoglu, Weak convergence of linear functional,
Bull. American Mathematical Society, 44 (1938), p. 196.
The theorem of James mentioned after Theorem 9 is from R. C. James,
Characterizations of reflexivity, Studia Math., 23 (1964), 205-16. The ori-
ginal form of Theorem 11, stating that for a Banach space X the set of
functionals attaining their norms on S(X) is dense in X*, is from E. Bishop
and R. R. Phelps, A proof that every Banach space is subreflexive, Bull.
Amer. Math. Soc, 67 (1961), 97-8; the form we gave it in is from B.
Bollobas, An extension to the theorem of Bishop and Phelps, Bull. London
Math. Soc., 2 (1970), 181-2. Our presentation followed that given in F. F.
Bonsall and J. Duncan, Numerical Ranges II, London Math. Soc. Lecture
Note Series, vol. 10, Cambridge University Press, 1973, vii + 179 pp.
9. EUCLIDEAN SPACES AND HUBERT SPACES

Having studied general normed spaces and Banach spaces, in the next
two chapters we shall look at the 'nicest' examples of these spaces,
namely Euclidean spaces and Hilbert spaces. These spaces are the
natural generalizations of the n-dimensional Euclidean space I".
A hermitian form on a complex vector space V is a map / : Vx V -* C
such that for all x,y,z E V and A,/A E C we have
(i) /(AJC + ^ , 2 ) =\f(x,z)+fjLf(y,z)
(ii) f(y,x) = f(x,y) foralljc^EF.
Similarly, a hermitian form on a real vector space V is a map
/ : VxV-+ U satisfying (i) and (ii). In that case (ii) says simply that / is
symmetric, so a real hermitian form is just a symmetric bilinear form. If
/ is a hermitian form on a complex vector space V then g = R e / is a
real hermitian form on VR, the space V considered as a real vector space.
Conversely, for every real hermitian form g on VR there is precisely one
complex hermitian form / on V with g = R e / (see Exercise 1).
Let / be a hermitian form on a real or complex vector space V. Note
that f(x,x) is real for all x E V and f(Xx,Xx) = \X\2f(x,x) for all x E V
and scalar A. We call f positive if, in addition to (i) and (ii), it satisfies
(iii) f(x,x) ^ 0 foralljcE V.
Two vectors x and y are said to be orthogonal with respect to f if
f(x>y) = 0; orthogonality is sometimes expressed by writing xly. A
vector orthogonal to itself is said to be isotropic. We call / degenerate if
some non-zero vector is orthogonal to the entire space; otherwise the
form is non-degenerate.

130
Chapter 9: Euclidean spaces and Hilbert spaces 131

Theorem 1. Let / be a positive hermitian form on V. Then for all


j c j E F w e have
\f(x,y)\2 ^ f(x,x)f(y,y) (1)
and
f{x+y,x+y)l/1 ^ f(x,x)1/2 + f(y,y)1/2. (2)
Furthermore, / is non-degenerate iff 0 is the only isotropic vector with
respect to / .
Proof. In proving (1), we may replace x by kx if |A| = 1 so we may
assume that f(x,y) is real. Then for all £ E R we have
0 ^ f(tx+y,tx+y) = t2f(x,x)+2tf(x,y)+f(y,y)
so (1) follows.
To see (2), note that, by (1),
f(x+y,x+y) *z f(x,x) + 2\f(x,y)\+f(y,y)
^ fb,x)+2[f(x,x)f(y,y)]1'2+f(y,y)

Finally, if x is isotropic then (1) implies that x is orthogonal to every


vector in V. D

Inequality (1) is, of course, just the Cauchy-Schwarz inequality; ine-


quality (2) is a variant of Minkowski's inequality for p = 2.
In view of Theorem 1, a positive hermitian form is non-degenerate iff
it is a positive-definite hermitian form, i.e. a hermitian form / such that
if x + Othen/(x,jt) > 0.
An inner-product space is a vector space V together with a positive
definite hermitian form on V. This positive-definite hermitian form is
said to be the inner product or scalar product on V and we shall write it
as (•, •) (and, occasionally, as (•,•))• Thus (•, •) is such that for all
x9y,z E V and for all scalars A and /i we have
(i) (A* + /iy,z) = \(x,z)+fL(y,z);
(ii) (y,x) = (x,y);
(iii) (x,x) 5* 0, with equality iff x = 0.
More often than not, it will not matter whether our inner-product
space is a complex or a real inner-product space. As the complex case
tends to look a little more complicated, usually we shall work with
132 Chapter 9: Euclidean spaces and Hilbert spaces

complex spaces. By Theorem 1, if (•, •) is an inner product on a vector


space V then ||*|| = (x,x)1/2 is a norm on V. A normed space is said to
be a Euclidean space or a pre-Hilbert space if its norm can be derived
from an inner product. A complete Euclidean space is called a Hilbert
space. Clearly every subspace of a Euclidean space is a Euclidean space
and every closed subspace of a Hilbert space is a Hilbert space.
The Cauchy-Schwarz inequality states that \{x,y)\ ^ ||*|||M|; in par-
ticular, the inner product is jointly continuous in the induced norm.
The inner product defining a norm can easily be recovered from the
norm. Indeed, we have the following polarization identities:
4(x,y) = \\x+y\\2-\\x-y\\2 + i\\x + iy\\2-i\\x-iy\\2 (3)
if the space is complex, and
2(x,y) = ||*+;H|2-||;t||2-|M|2 = *{||*+>'||2-||*->'||2} (4)
in the real case. Therefore in a Euclidean space we may, and often
shall, use the inner product defining the norm. For this reason, we use
the terms 'Euclidean space' and 'inner-product space' interchangeably,
and we may talk of orthogonal vectors in a Euclidean space.
The complex polarization identity (3) has the following simple exten-
sion. If T is a linear operator (not necessarily bounded) on a complex
Euclidean space then
4(Tx,y) = (T(x+y),x + y)-(T(x-y),x-y)
+ i(T{x + iy),x + iy)-i(T(x-iy),x-iy). (3')
This implies the following result.
Theorem 2. Let £ be a complex Euclidean space and let T E !£(E) be
such that (Tx,x) = 0 for all x E E. Then T = 0.
Proof. By (3') we have (Tx,y) = 0 for all x,y E E. In particular,
||7x||2 = (Tx, Tx) = 0 for all x E E, so T = 0. •

It is worth pointing out that Theorem 2 cannot be extended to real


Euclidean spaces (see Exercise 2).
In a Euclidean space, the theorem of Pythagoras holds and so does
the parallelogram law.
Theorem 3. Let E be a Euclidean space. If xi,...,xn are pairwise
orthogonal vectors then
Chapter 9: Euclidean spaces and Hilbert spaces 133

2 xj = 2 (5)

Furthermore, if x,y E E then


(6)
Proof. Both (5), the Pythagorean theorem, and (6), the parallelogram
law, are immediate upon expanding the sides as sums of products. To
spell it out,
2
n n
I 2J X( y 2J Xi\ 2J \X( 9X1) * 2a \Xi 5 Xi)

and
:+y)-(x-y,x-y) = 2||x| D

In fact, the parallelogram law characterizes Euclidean spaces (see


Exercise 3). This shows, in particular, that a normed space is Euclidean
iff all its two-dimensional subspaces are Euclidean. Furthermore, a
complex normed space is Euclidean iff it is Euclidean when considered
as a real normed space.
The last assertion is easily justified without the above characterization
of Euclidean spaces. Indeed, if E is an Euclidean space then Rt(x,y) is
a real inner product on the underlying real vector space of E and it
defines the same norm on E. Conversely, if V is a complex vector space
and (•, •) is a real inner product on V (i.e. (•, •) is an inner product on
VR) such that the induced norm || • || satisfies ||AJC|| = |A| \\x\\ for all x E V
and A G C then
(x,y) = (x9y)-i(ix,y) = (x,y) + i(x,iy)
is a complex inner product on V defining the original norm | | | | and
satisfying Re(x,y) = (x,y).
Examples 4. (i) Clearly, (x,y) = 2" = 1 xJi is a n inner product on /2"
and so l£ is an n-dimensional Hilbert space.
(ii) Also, (x,y) = 2°°=1 xi^i *s a n inner product on l2 and so l2 is a
separable infinite-dimensional Hilbert space.
(iii) Let E be the vector space of all eventually zero sequences of
complex numbers (i.e. x = (Xj)™ belongs to E if xt = 0 whenever i is
134 Chapter 9: Euclidean spaces and Hilbert spaces

sufficiently large), with inner product (x,y) = 2°°=1 xi?i- Th e n £ is a


dense subspace of /2; it is an incomplete Euclidean space,
(iv) It is immediate that
fb —
(f,g) = f(*)g(t) dt
'a

is an inner product on the vector space C([a,b]); the norm defined is


the /2-norm
/ fb \l/2

(J j
and not the uniform norm. This is an incomplete Euclidean space (see
Exercise 14).
(v) Also,

(/>*)= f /(')g«d;
•*a

is the inner product on

L2(0,1) = 1/ G C[o> 1]: / i s measurable and J \f(t)\2 dt < °°L

defining the L2-norm

With this norm L2(0,1) is a Hilbert space.


(vi) Let V = Cn, let A = (atj) be an nxn complex matrix and, for
x = (*,.)? and y = (y^i in V define
n
f(x,y) = 2 «i>^i>7-

Then / is a hermitian form if a;i = afy for all /,/". Also, every hermitian
form on V can be obtained in this way. •

Theorem 5. The completion of a Euclidean space is a Hilbert space.


Proof. The assertion is that the inner product (•, •) on a Euclidean
space E can be extended to the completion E of E. Let (xn) and (yn)
be Cauchy sequences in E and let x and y be their equivalence classes
in E. Define
Chapter 9: Euclidean spaces and Hilbert spaces 135

(xj) = lim(xn,yn).
n—Kx>

Since
\(xn9yn)-(xm9ym)\ ^ \(xn-xm,yn)\ + \(xm,yn-ym)\

the limit exists. It is easily seen to depend only on x and y and not on
the particular representatives (*„), (yn). Furthermore, (•, •) is an inner
product on E, extending E and defining precisely the norm on the com-
pletion E. •

Let £ be a Euclidean space. For x E E, write x1 for the set of vec-


tors orthogonal to x:
x1 = {y<EE:x±y} = {yEE:(x,y) = 0}.
Clearly x ± is a closed subspace of E and therefore so is
S1 = {y E E : x Ly for all x E S) = Ox1

for every subset S of E. If F is the closed linear span of S, i.e.


F = HnS, then S1 = F1 and ( S 1 ) 1 D F.
Call two subspaces Fx and F2 orthogonal if JCX 1JC2 for all xx E Fx and
x2 E F2- If ^ is a subspace of £ then F and F1 are orthogonal sub-
spaces and F f l F 1 = {0} since if x E F f l F x then x Ix and so x = 0. If
JC E F and y E F 1 then ||x+}>||2 = IM| 2 +|M| 2 , so the projections
x+y »-> JC and jc+y •-• y are both bounded. Thus F+FL is the orthogo-
nal direct sum of F and F x . In general, F+FL need not be the entire
space, not even when F is closed (see Exercise 6). However, when F is
complete, this is the case.
Theorem 6. Let F be a complete subspace of a Euclidean space E.
Then E is the orthogonal direct sum of F and F 1 : every x E E has a
unique representation in the form
x = xx + x2 (*i E F, JC2 E FL).
Furthermore, if y E F and y ^ JCX then
(7)
Proof. Let x E £ and set d = d(*,F). What characterizes xxl By (7),
it has to be the unique vector in F nearest to x, precisely at distance d.
136 Chapter 9: Euclidean spaces and Hilbert spaces

Let then yn E F be such that

\\x-yn\\2<d2+K
By identity (6), the parallelogram law,
\\yn-ym\\2 = 2\\x-yn\\2 + 2\\x-yJ2-\\2x-yn-ym\\2
< 2 2_
^ n m '
and so (y,,)* is a Cauchy sequence. Since F is complete, >>„ -> JCX for
some JCJ E F. Then HJC-JCJ = d.
Let us show that x2 = x-xt is orthogonal to F. Suppose this is not
so. Then F has an element, say y, such that (x2,y) ^ 0. Replacing y
by z = (Jt2,)>))> we have (JC2,Z) E R and (*2,z) > 0. But then for
sufficiently small e > 0 we have
\\x-(x1 + ez)\\2=\\x2-ez\\2
= d2-2(x2,z)e+\\z\\2e2<d2,
contradicting d(x,F) = d.
Inequality (7) is immediate from identity (5), the Pythagorean
theorem. If y E F and x1-y ^ 0 then, as (x1 -y) lx2, we have
\\x-y\\2 = \\xl+x2-y\\2 = I k - y f + N I 2 > ||*2||2 = d\ D
In view of Theorem 6, if F is a complete subspace of a Euclidean space
E then we call F 1 the orthogonal complement of F in E. The map
Pp. E = F+ FL -* E defined by JC = Xi + JC2 •-> JCX is called the orthogonal
projection onto F; we have just shown that there is such a map.
Corollary 7. Let F be a complete subspace of a Euclidean space E.
Then there is a unique operator P F E 28(£) such that

\ for XEF\
Furthermore,
ImPF = F, KerPF = F \ ?} = P f , (I-PF)2 = I-PF
and if *, _y £ £ then
(Pjric.y) = (PFx,PFy) = (*,F f y).
If F + (0) then ||P f || = 1 and if F ? E then | | / - P F | | = 1. D
Chapter 9: Euclidean spaces and Hilbert spaces 137

Corollary 8. Let if be a Hilbert space, let 5 C H and let M be the


closed linear span of 5. Then {SL)L = (M1)1 = M.
Proof. As H = M+ML = (M±)± + M± and M C (M 1 ) 1 , we have
M = (M1)1 = (S1)1. •

Given a Hilbert space / / and a vector x0 E //, the function / : H -> C


defined by /(#) = (jt,jt0) is a bounded linear functional of norm ||JCO||.
Indeed, / is clearly linear, and \f(x)\ = \(x,xo)\ ^ ||*oll 11*11 anc* s o
\\f\\ * \\xo\\. Furthermore, \f(xo)\ = \\xof and so ||/|| ^ \\xo\\. The
existence of an orthogonal decomposition implies that all bounded linear
functionals on H can be obtained in this way.
Theorem 9. (Riesz representation theorem) Let / be a bounded linear
functional on a Hilbert space H. Then there is a unique vector JC0 E H
such that/(;c) = (x,x0) for all x E H. Furthermore, ||/|| = ||jto||.
Proof. We may assume that / ^ 0. Then M = Ker/ is a closed one-
codimensional subspace of H. Consequently M1 is one-dimensional,
say M 1 = {A£i^ A E C}, where xx E H and |JCX || = 1.
Put x0 = f(xi)xi. Every vector x E H has a unique representation in
the form
x = y + \xl9 (y E Af; A E C).
Then

To see the uniqueness of x0, note that if (x,x0) = (X,XQ) = 0 for all
x E H, then
ll*o-*6ll2 = (xo-xo,xo-xo) = 0.
The assertion ||/|| = ||JCO|| was shown immediately before the theorem.

Corollary 10. Let H be a Hilbert space. For y E H let fy E H* be
defined by fy(x) = (x,y)(x E / / ) . Then the map H-+ H* defined by
y *-> fy is an isometric anti-isomorphism between H and //*, i.e. if
y*-*f and z »-* g then ky + iig*-* \f+jlg. If / / is a real Hilbert space
then the map is an isometric isomorphism between H and H*. D
138 Chapter 9: Euclidean spaces and Hilbert spaces

The last result enables us to identify a Hilbert space with its dual; this
identification is always taken for granted. It is important to remember
that the identification is an antf-isomorphism since this will make the
adjoint of a Hilbert-space operator seem a little different from the
adjoint of a Banach-space operator.

Exercises

1. Let / be a hermitian form on a real vector space U. Let


V = U+iU and extend / to a function / : Vx V —> C by setting
= f(x,z) +f(y, w)+if(y,z)-if(x, w).
Show that / is a hermitian form on the complex vector space V
and f\U = / . Show also that / is positive iff / is, and that / is
degenerate iff/is.
2. Give an operator T E $(/22) such that (Tx, x) = 0 for all x E /22
and || r|| = 1.
3. Show that a normed space E is a Euclidean space if and only if the
parallelogram law holds in E, i.e.
\\x+y\\2+\\x-y\\2 = 2\\x\\2
4. Let £ be a Euclidean space. Show that for xx,..., xn E £ we have
2
II

where the summation is over all 2n sequences (e,)" (e, = 1 or - 1 ) .


5. Letw = ( i , i , J , . . . ) E / 2 a n d
F = {x = (xk)°? : (x, u) = 0 and xk = 0 if k is large enough} C /2.
Show that
FL = lin{w} = {Aw : A is a scalar}.
6. Construct a Euclidean space E and a closed subspace F C E
( F ^ £ ) such that F1 = (0). [Note that then, in particular,
E + F+F1.]
7. Let E be a Euclidean space and let P E ^(E) be a norm-1 projec-
tion: P2 = P and ||P|| = 1. Show that P is the orthogonal projec-
tion onto F = ImP, i.e. KerP = FL and E = F+F1.
8. Let A be a non-empty closed convex subset of a Hilbert space H.
Show that the distance from A is always attained: for every b E H
Chapter 9: Euclidean spaces and Hilbert spaces 139

there is a unique point a = a(b) E A such that

||6-fl|| =d(b,A)=inf{\\b-x\\:xEA}.
Show also that the map b »-> a(b) is continuous.
9. Let JC, y and z be points in a Euclidean space. Prove that

10. Let X be the space of continuously differentiable functions on


[0,1], with norm
1 /2
/f V
2 2
11/11 = (J {*/ (*) + 2|/'to| }d*j .
Show that X is a Euclidean space. What is the inner product
inducing the norm?
11. Let F be a closed subspace of a Hilbert space H. Show that the
quotient space H/F is also a Hilbert space.
12. Let F be an arbitrary set and let 12(F) be the vector space of all
functions / : T—> C such that 2 e r 1/(7)12 < °°« In particular, if
/ E / 2 (f) then {y E T : /(y) ^ 0} is countable. Show that 12(F) is a
Hilbert space with then norm

|rl/(r)l:
13. Let {Hy : y G 1} be a family of Hilbert spaces. Let H be the vec-
tor space of functions / : F -*• \J e r Hy such that f(y) G Hy and

2 11/(7) II2
Show that
kV2
/1 - I 2 ll/(r)ll2
111
is a norm on H and with this norm H is a Euclidean space. Is H
necessarily a Hilbert space?
14. Turn C[0,1] into a Euclidean space by setting

f,8) =
(f,8) f
Jo
Show that this space is incomplete.
140 Chapter 9: Euclidean spaces and Hilbert spaces

15. / : [0,1] -> R such that / ' E L2(0,1) and /(0) = 0. Prove that V is
a real Hilbert space with inner product

•r /WW *•
16. A real «xn matrix A is called orthogonal if (Ax,^4y) = (jc,y) for
all x,y E /2n. Prove that A is orthogonal iff it maps orthogonal
vectors into orthogonal vectors and has norm 1.
17. Let P be a bounded linear projection in a Euclidean space E (i.e.
P E 5(£) and P 2 = P). Show that ||P|| = 1 iff P ± 0 and
ImPlKerP.
18. Let A be a non-empty subset of a Hilbert space /f and let
T e ^(//) be such that TH C A and (JC- 7*) l i for every x E H.
Show that T is a bounded linear operator, A is a closed linear sub-
space and T is the orthogonal projection onto A: T = PA.
19. Show that the unit ball of l2 contains an infinite set A such that
||jt->>|| > \[2 for all x,y E A (x ? y). Show also that A must con-
sist of norm-1 vectors.

Notes

The concept of an abstract Hilbert space was introduced by J. von Neu-


mann, Eigenwert Theorie Hermitescher funktional Operatoren, Math.
Ann., 102 (1930), 49-131. Earlier special realizations of a Hilbert space
had been considered by several people. In particular, from 1904 to 1910
David Hilbert published some fundamental papers on integral equations
which led him to consider some 'Hilbert spaces' of functions. It seems
that the name 'Hilbert space' was first used by F. Riesz in his book, Les
systimes d'equations a une infinite d'inconnus, Paris, 1913, for what we
know as /2: "Considerons l'espace Hilbertien; nous y entendons
l'ensemble des systemes (JC^) tels que 2 \xk\2 converge."
10. ORTHONORMAL SYSTEMS

In this chapter we continue the study of Euclidean spaces and Hilbert


spaces. As we shall see, every separable Euclidean space contains the
exact analogue of the canonical basis of /2" = (Cn, || • ||). This means that
we can use 'coordinates', which in this context we call Fourier
coefficients, to identify the points of the space, these coefficients behav-
ing very much like the coordinates in C" or Un. If our space is in fact a
Hilbert space then the space is thus naturally identified with l2, telling us
that all separable Hilbert spaces are isometric.
Let £ be a Euclidean space and let S C E be a set of vectors. If E is
the closed linear span of S, i.e. linS = E, then we call S fundamental or
total. We call S orthogonal if it consists of non-zero pairwise orthogonal
elements. An orthogonal set of unit vectors is said to be an orthonormal
set. An orthogonal set is complete if it is a maximal orthogonal set. A
complete orthonormal set in a Hilbert space is said to be an orthonormal
basis. We shall see, amonst other results, that every orthonormal basis is
a Schauder basis of the best kind, with basis constant 1 (see Exercises
16-18 of Chapter 5), and every Hilbert space contains an orthonormal
basis.
Occasionally, we call a set 5 of vectors a system of vectors, especially
when 5 is not written in the form of a sequence.
Theorem 1. A fundamental orthogonal set in a Euclidean space is com-
plete. In a Hilbert space every complete orthogonal set is fundamental.
Proof, (i) Let 5 be a fundamental set in a Euclidean space E. Then
S1 = (linS) 1 = E1 = (0); so if 5 is orthogonal, then it is complete.
(ii) Let S be a complete orthogonal set in a Hilbert space H. Set
M = finS. Then M1 = (0) so M = (A/-1)1 = H. Thus 5 is a funda-
mental system. •
141
142 Chapter 10: Orthonormal systems

The Gram-Schmidt orthogonalization process enables us to replace a


sequence of linearly independent vectors by an orthonormal sequence.
Theorem 2. Let x1, x2,... be linearly independent vectors in a Euclidean
space E. Then there exists an orthonormal sequence yi,)?2,• • • such that
hn{xl9...,xn} = lmfo,...,>>„} for every n.
Proof. For n = 0,1,... set
Mn = Iin{x1,...,xn} and zn+1 = xn+1-PMnxn+u
where PM is the orthogonal projection onto Mn, so that (x — PMnx)±Mn
for every x E E. Then z n + 1 ^ 0 since xn+1 0 Mn and PMnxn+i e ^ « -
Also, z n + i l M f l . Set^ n + 1 = 2: n+1 /||z w+1 ||. Then the sequencey 1 ,y 2 ,"-
has the required properties. Indeed, by the construction, Mn =
lin{yi,...,y n } and so the ^ are linearly independent. Furthermore,
zn+1±Mn and so )>„+!!Af,, = hn{yl9...,yn}. D

In fact, it is easy to define explicitly the orthogonal projection PMn


appearing in the proof of Theorem 2; namely, set PMQ — 0 and for n ^ 1
define
n
PMHX = 2 (x,yi)yi.

The Gram-Schmidt orthogonalization process enables us to show that


every separable space (i.e. every space containing a countable dense set)
contains a fundamental orthonormal sequence. Also, by Zorn's lemma
every Euclidean space contains a complete orthonormal system.
Theorem 3.
(a) Every separable Euclidean space contains a fundamental orthonor-
mal sequence.
(b) In a Euclidean space, every orthonormal system is contained in a
complete orthonormal system. In a Hilbert space, every orthonor-
mal system is contained in an orthonormal basis.
Proof, (a) Let (xn)i be a dense sequence in a Euclidean space E. Dis-
card xn if it is in Iin{x1,...,xn-1} and apply Theorem 2 to the sequence
obtained,
(b) Let So be an orthonormal system in a Euclidean space E. Set
X = {S : SQ C 5 and 5 is an orthonormal system in E}.
Then X is partially ordered by inclusion and every totally ordered subset
X' of X has an upper bound in X, namely Us^x> 5. Hence, by Zorn's
Chapter 10: Orthonormal systems 143

lemma, X has a maximal element Sx. Thus S1 is a complete orthonor-


mal system containing So. If E is a Hilbert space, then the system 5X is
a fundamental system. •

Examples 4. (i) Let T be the unit circle in the complex plane:


T = { z 6 C : \z\ = 1}. Equivalently, T is the real line modulo 2TT:
T = {e1': t E U} = {e1': 0 ^ t ^ 2TT}.
Let E be the Euclidean space C(T) with inner product

where we have used the second convention: / and g are continuous


functions from [0,2TT] to C, with /(0) = /(2TT) and g(0) = g{2ir). The
functions ?>„(*) = emt (n = 0, ± 1 , ±2,...) form an orthonormal system in
E. By the Stone-Weierstrass theorem, the linear span of these func-
tions is dense in C(T) endowed with the uniform norm. Hence
{<pn : n = 0, ±1,...} is a fundamental, and so complete, system in E.
The completion of E is L2(T).
(ii) Let E be the Euclidean space C [ - l , 1] with inner product

2
The functions l,t,t ,...
-t
(f,8) = I /(')«(')
are linearly independent in E and, by the
Stone-Weierstrass theorem, form a fundamental system in E. Let
(G/i(0)o be the sequence obtained from l,t,t2,... by the Gram-
Schmidt orthogonalization process. Then Qn(t) is a multiple of the nth
Legendre polynomial

where D is the differentiation operator (see Exercise 1).


(iii) I,sinf,cosf,sin2f,cos2f,... is an orthogonal sequence in the Hil-
bert space L 2 (0,TT). Since the subspace of continuous functions is dense
in L 2 (0,TT), by the Stone-Weierstrass theorem this sequence is a funda-
mental orthogonal sequence.
(iv) For n = 0 , 1 , . . . define the nth Rademacher function rn E L2(0,1)
by
1 if \2nt\ is even
r (t)
" • -1 if [2nt\ is odd.
144 Chapter 10: Orthonormal systems

Equivalently, rn(t) = sign sin 2nirt. (Note that we view rn as an element


of L 2 (0,1); so it is, strictly speaking, an equivalence class of functions
differing on sets of measure 0. Thus it is also customary to define rn(t)
to be (-1)* if kl~n < t < (A;+l)2~ n and leave it undefined for t =
k2~n.) Then (rn(f))o is an orthonormal system in L 2 [0,1]; it is easily
seen to be incomplete.
(v) Let -oo ^ a < b ^ oo and let <p E C(a,b) be a positive function.
For / and g in the space Cc(a, b) of continuous functions with compact
support, set

(f>8) = f f(t)8®<p(t) <*'•


*a
Then (•, •) is an inner product on Cc(a,b); denote its completion by
L2((p). The space L2(<p) consists of all measurable functions f(t) on
[a, b] such that

r
Now let a = 0, b = oo and cp(t) = e '. Then the functions 1 , M 2 , . . .
form a fundamental system in L2(<p) and the polynomials obtained from
them by the orthogonalization process are the Laguerre polynomials up
to constant factors, i.e. multiples of e*Dn(e~V*) (n = 0 , 1 , . . . ) (see Exer-
cise 5).
(vi) Consider the space L2((p) constructed in Example (v) but
with a = —oo? b = oo and (p(t) = e~ r . Orthogonalizing the sequence
l,t,t2,... we obtain multiples of the Hermite polynomials, i.e. multiples
of e'2Drt(e~'2) (n = 0 , 1 , . . . ) (see Exercise 6). •

If ((pic)i is an orthonormal basis in an n-dimensional Euclidean space


E, then every vector x E E is a linear combination of the cpk:
x = 2 £ = 1 ck(Pk- Furthermore ck = (x,<pk). Also, if

C
* = 2 *<P* and >> =

then

(x,y)= 2 c*4 and W2= £ £


A: = l A: = l
In other words, the correspondence x ** (ck)1 identifies E with l2.
Chapter 10: Orthonormal systems 145

The main reason why an orthonormal basis (cpk)i in a Hilbert space H


is very useful is that analogous assertions hold concerning the represen-
tations of vectors in H. As we shall see, for every vector x E H there
are unique coefficients (ck)i such that x = 21-1 ck<Pk- This sequence
(ck)i satisfies ||JC|| = 2 r = 1 \ck\ afl d every sequence satisfying
2 r = 1 ICA:|2 < °° arises as a sequence of coefficients. Thus an orthonor-
mal basis enables us to identify H with l2, just as any n-dimensional
Euclidean space can be identified with l2.
Theorem 5. Let (<pk)i be an orthonormal sequence in a Hilbert space
H. Then, for a scalar sequence c = (ck)i, the series S^ = 1 ck<Pk *s c o n -
vergent iff c E l2 (i.e. 2 £ = 1 ICA: 12 < °°)• K the series is convergent then
/ 00 \V2

=||C||2= 2 |C,| 2 . (1)


U /
Proof. Set xn = 2 ^ = 1 CA:<PA:- Then, for 1 ^ n ^ m , by the Pythagorean
theorem, we have
m
\\xn-xmf= 1 \ck\\
k=n + l
Hence (xn)i is convergent iff 2 ^ = 1 \ck\2 is convergent. Relation (1)
holds since, again by the Pythagorean theorem,

Ikll2= i kl 2 .
k=l •
A slightly different formulation of Theorem 5 goes as follows. Let
(<Pk)°i be an orthonormal sequence in a Hilbert space H, and let
c = c
( k)i ^ h- Then there is a vector x E H such that (x,cpk) = c^ for
every k. This is usually called the Riesz-Fischer theorem. It looks par-
ticularly easy (it is particularly easy) because the space is assumed to be
complete. The original form of the theorem concerned L 2 [0,1], where
the completeness, which is far from trivial, had to be proved.
Let (<pk)°i be an orthonormal sequence in a Hilbert space H. For
x E H, set ck — (x,<pk) (k = 1,2,...). We call c1,c2,... the Fourier
coefficients of x with respect to ((Pk)™; the series

2 ckq>k
k=l

is the Fourier series of x with respect to


146 Chapter 10: Orthonormal systems

It is easily seen that if (cpk)i is an orthonormal basis then every vector


is the sum of its Fourier series and so, in particular, it is determined by
its Fourier coefficients. In fact, we have the following slightly more gen-
eral result.
Theorem 6. Let (<pk)i be an orthonormal sequence in a Hilbert space H
and let M = l i n ^ ) * . Then, for all x,y E H, we have

(0 f (x,<pk)<pk = PM(x);
k= \

(H) 2 |(WI2=||PM«I

(iii) 1 (x,<pk)(y,<Pk) = (PM(x),PM(y)) = (x,PM(y)) = ( W j ) .

Also, suppose c = (ck)i E / 2 , i.e. 2 ^ = 1 \ck\2 < °°- Then there is a


unique vector u E M with Fourier coefficients ck = (u,(pk) (k =
1,2,...), namely w = 2 £ = 1 CA:<PA:- Furthermore, v E H has Fourier
coefficients (cfc)5° iff w = PMv.
Proof. Set

Then, for 1 ^ k ^ n, we have (JC—JCn,<pfc) = 0 and so x — xn + (x—xn)


is a representation of x as a sum of orthogonal vectors. Therefore, by
the theorem of Pythagoras,
n
2
||v||2 _| | v ||2 +I | | v _ v ||2 > | | v ||2 _ V l / v /« M
ll^ll ~ ll^nll ll^ Xn\\ ^ ||JCn|| - 2^ \\X 9<Pk)\ •

Letting n -> », we see that 2 ^ = 1 |(*>^*)| 2 < 0 ° so, by Theorem 5,


2 r = 1 (x,<pk)<Pk is convergent, say to JC' E M.
As for every k we have

(x-x',<pk) = {x-\\m^xn,(pk) = ( x , ^ ) - U m U , , , ^ )

= (x,<pk)-(x,<pk) = 0,
JC-JC'is orthogonal to M and so x' = PA/^- This proves (i); further-
more, (ii) follows from (i) and (1).
Chapter 10: Orthonormal systems 147

To see (iii), set


n

Then PMy = lim yn and so

(PMx,y) = (x,PMy) =

= lim 2 (^>

The proof of the second part is equally easy. By Theorem 5, the series
2 ^ = 1 ck<Pk is convergent to some vector u E M. Then

(u,<pk) = lim I 2 c/^p/, <pk) = ck

for every k, and so u does have Fourier coefficients c l 5 c 2 , Also, if


w = PMv then
(«;,^) = (v,PM<pk) = (PMv,<Pk) = (u,<Pk)
and so u and ^ have the same Fourier coefficients. Finally, if v E H
and (v,<pk) = c^ for every k then (v-u,(pk) = 0 for every fc and so
v — u is orthogonal to lin(^)5° = M. Therefore u — PMV, as claimed. •

The following useful corollary is amply contained in the result above.


Corollary 7.
(a) (Bessel's inequality) If {(PkY\ is an orthonormal sequence in a
Euclidean space E and x E E then

(b) (Parseval's identities) If ((Pk)i is a complete orthonormal sequence


in a Hilbert space H and x,y E H then

= Ikll2 and £ (x,<Pk)(y,<Pk) = (x,y). D


148 Chapter 10: Orthonormal systems

Parseval's identities imply that every infinite-dimensional Hilbert


space with an orthonormal basis (i.e. a complete orthonormal sequence)
is isometric to l2.
Theorem 8. Let (<Pk)°\ be an orthonormal basis in a Hilbert space H.
For x E H define
x{k) = (x,<pk) and x = (x(k))°? = ((x,<pk))i.
Then the map H -> l2 given by x •-> x is a linear isometry of H onto l2.

Corollary 9. Every n -dimensional Euclidean space is linearly isometric
to l2 and every separable infinite-dimensional Hilbert space is linearly
isometric to /2. •

Having seen that all separable infinite-dimensional Hilbert spaces are


isometric to /2, what can we say about other Hilbert spaces? As in
Exercise 11 of Chapter 9, given a set F, we denote by 12(F) the vector
space of complex-valued functions f on F with countable support and
such that 2 y e r |/(y)| 2 < °°. Then
/ \1/2
2
11/11= 2 l/(r)l
\yer /
is a norm on 12(F) and with this norm 12(F) is a Hilbert space. Thus
/2(N) is precisely l2 and / 2 ({1,2,..., n}) is l2 .
Theorem 10. Every Hilbert space is isometrically isomorphic to a space

Proof. Let H be a Hilbert space. Then H contains a complete ortho-


normal system, say {cpy : y E / ] . Then, by Bessel's inequality, for every
x EL H and for every countable subset Fo of F, we have

Hence the set Fx = {y : (x,<py) ^ 0} is countable,


*- 2 (x><Py)<Py

is orthogonal to every ^^ and so


Chapter 10: Orthonormal systems 149

Setting x{y) = (jt,<pr), the map if—> I2{f) given by x »-» i is a linear
isometry of / / onto I2(r). •

The fact that every separable (infinite-dimensional) Hilbert space is


isometrically isomorphic to l2 is very important when we are studying
the abstract properties of Hilbert spaces. Nevertheless, in applications
Hilbert spaces often appear as spaces of functions, and then we are
interested in the connections between the Hilbert space structure and
the properties of the functions.
One of the most important Hilbert function spaces is L^T), the com-
pletion of C(T), the space of continuous functions on the circle T,
defined in Example 4(i). In the standard orthonormal basis {tmt: n =
0, ±1, ±2,...}, the A:th Fourier coefftcent ckoffE L2(T) is

and
n

k = -n
is the nth partial sum. Then, by Theorem 8,
Jim \\SJ-f\\ = 0 (2)
for every/ E L2(T), with ||-|| denoting the Hilbert space norm, i.e. the
norm in L2(T). Thus the partial sums converge in mean square to / .
For / E C(T) relation (2) is an immediate consequence of the
Stone-Weierstrass theorem, which tell us that / can be uniformly
approximated by a trigonometric polynomial. (Of course, we used pre-
cisely the Stone-Weierstrass theorem to show that {em': n =
0 , ± l , ± 2 , . . . } is a complete system.) In particular, for every e > 0
there is a trigonometric polynomial/? such that \f(t)-p(t)\ < e for all t.
But if p has degree n (i.e. p(t) = 2£ = _ r t dkelkt) then Snp = p. Since
the projection operator Sn has norm 1, we have
I|5 W /-/|| = \\(Sn-I)f\\ ^ USn-l)p\\+2\\f-p\\ < 2e.
The very easy relation (2) leaves open the question whether the
Fourier series of / E L2(T) tends to / point wise in some sense. For
example, it may not be unreasonable to expect that if / is continuous
then (Snf)(t) tends to f(t) for every t. Sadly, this is not true, as was
150 Chapter 10: Orthonormal systems

first shown by du Bois Reymond in 1876. Nevertheless, Fejer proved in


1900 that there is a simple way to recover a continuous function from
the partial sums of its Fourier series. To be precise, Fejer showed that
the Fourier series of a continuous function if Cesdro summable and the
sum is the function itself: if / is continuous then

2 skf,
A:=0
i.e. the average of the partial sums, tends uniformly to / . Fejer's
theorem was of tremendous importance: it launched the modern theory
of Fourier series.
Although (Snf)(t) need not converge to f(t), the set of points at which
it fails to tend to f(t) cannot be too large. It was asked by du Bois Rey-
mond whether the partial sums Snf of a function / E C(T) tend to /
almost everywhere, and later Lusin conjectured that the answer is in the
affirmative for every / E L 2 (T). For many years this was one of the
most famous conjectures in analysis; finally, it was proved by Carleson
in 1966. The very intricate proof is one of the greatest triumphs of hard
analysis; not surprisingly, it is far beyond the scope of this book.

Exercises

1. Show that the Legendre polynomials

form an orthogonal basis of L 2 ( - l , 1). Deduce that, up to a posi-


tive factor, Pn(t) is the nth term in the sequence obtained from
l,t9t2,... by the Gram-Schmidt orthogonalization process, count-
ing P0(t) = 1 the Oth term, Px(t) = t the first, P2(t) = i ( 3 f 2 - l ) the
second, etc.
2. Show that Pn{\) = 1 and Pn(-\) = (-1)".
3. Deduce from the previous exercises that

nPJ(i)-(2n-l)tPn-y!(i)Hn-\)Pn-2® =0
for all n ^ 2.
4. Show that D ( ( / 2 - l)P'n(t)) is orthogonal to tk for k < n, and
deduce that

) Z { ) & ) { )n{) =0
for every n ^ 1.
Chapter 10: Orthonormal systems 151

5. Check that the Laguerre polynomials e ' D ^ e ' V ) (n = 0 , 1 , . . . ) are


indeed orthogonal polynomials in L2(<p), where cp : [0,<») —» R is
e~', as in Example 4(v).
6. Check the analogous assertion about the Hermite polynomials
e ' D ^ e " ' ) and weight function <p = e~' : (- 0 0 , 0 0 ) —> R, as in
Example 4 (vi).
7. Let L2(<p) ^ e a s defined in Example 4(v), with a positive weight
function cp : (a, b) —> R. Suppose that 1, f, f2,... E L2(<p). Let
Po>Pi>P2>- be the polynomials obtained by orthogonalizing the
sequence l,t,t2,... . (We do not normalize the sequence; so pn(t)
is the unique monic polynomial of degree n orthogonal to every tk
for A: < n.)
(i) Show that

(Pn,tPn-l) = \\Pn\\2'
(ii) Show that for n ^ 2 we have
pn(t) = (t-an)pn_x(!L)-bnpn_2(i)
where
\\Pn-lf
an=
\\Pn-l\\2
8. Let JCX , . . . , JCW be unit vectors in a Euclidean space and let
yi,..>,yn be the sequence obtained from it by Gram-Schmidt
orthogonalization, with

;=i
Show that \\a\ ^ 1, with equality only if y t = ktiXi.
9. For at = (^ y ); = 1 E /2n (i = 1,...,/*) let ^ = A(al9...,an) =
(aij)"J=1 be the nxn complex matrix whose ith row is at. Prove
Hadamard's inequality, stating that

|deM| ^ ft \\atl
i= \
with equality if and only if either some at is 0 or the at are orthog-
onal.
10. Let xi,...,xN be linearly independent vectors in a Euclidean
space, with N = ("21). Show that there are orthonormal vectors
yl9...,yn such that yt = 2 ; e y 4 Ay-xy, where A1,A2,...,An are dis-
joint subsets of { 1 , 2 , . . . , N}.
152 Chapter 10: Orthonormal systems

11. Define the Gram determinant of a sequence Xi,x2,...,xn in a Hil-


bert space as

G(xl9...,xn) = det((xi9Xj))Zj=l9
i.e. as the determinant of the nxn matrix whose entry in the ith
row and /th column is the inner product (xi9xj). Show that
G(xi9...9xn) ^ 0, with equality iff the set {x1,...xn} is linearly
dependent.
Show also that if {x1,..., xn} is a linearly independent set span-
ning a subspace Mn then

. M\- (G(x,xux2,...,xn)\1/2
9Mn) -

12. Prove the following weak form of the Riemann-Lebesgue lemma:


i f / E C(T) then

f(t)e~int At -> 0 as|/i|


2 * Jo
13. The aim of this exercise is to prove Fejer's theorem. F o r / G L2(T)
set

k=O

(i) Show that


r 2ir

t) = — ( f(x)Kn(t-x) dx,

where Kn(s), the Fe/'^r kernel, is defined as

(ii) Prove that if 0 < s < 2TT then

and
Chapter 10: Orthonormal systems 153

(iii) Deduce from (i) and (ii) that Kn{s) ^ 0 for all s E T,
Kn(s) -> 0 uniformly for 0 < 8 ^ s ^ 2ir-8 < 2TT and

(iv) Deduce Fejer's theorem: if / G C(T) then


<rn(f,t)-+f(t)
uniformly in t as n —> °°.
(v) Show also that if / E Loo(T) and / is continuous at t0 then

14 + . Let / / be a Hilbert space and let Sx = {xy : y E F} and 5 2 =


{yr : y E / } be such that (xa,y^) = 8ap (where 8af3 = 0 i f a + p
and 8 a a = 1). Suppose 5X is a fundamental system, i.e. UnS1 =
//. Does it follow that S2 is also fundamental?

Notes

The chapter is about the beginning of abstract Fourier analysis. As we


have hardly scraped the surface of Fourier analysis proper, the reader is
encouraged to consult a book on the topic; T. W. Korner, Fourier
Analysis, Cambridge University Press, 1988, xii + 591 pp., is particu-
larly recommended.
The original Riesz-Fischer theorem was proved in F. Riesz, Sur les
systemes orthogonaux de fonctions, Comptes Rendus, 144 (1907), 615-
19 and 734-36, and E. Fischer, Sur la convergence en moyenne,
Comptes Rendus, 144 (1907), 1022-4. It was a clear case of indepen-
dent discovery, precisely described by Fischer in the introduction of his
paper: "Le 11 mars, M. Riesz a presente a l'Academie une Note sur les
systemes orthogonaux de fonctions (Comptes Rendus, 18 mars 1907).
J'etais arrive au meme resultat et je l'ai demontre dans une conference
fait a la Societe mathematique a Briinn, deja le 5 mars. Ainsi mon
independance est evidente, mais la priorite de la publication revient a
M. Riesz."
Carleson's theorem, mentioned at the end of the chapter, was proved
in L. Carleson, Convergence and growth of partial sums of Fourier
series, Acta Math., 116 (1966), 135-57.
Exercise 9 is from J. Hadamard, Resolution d'une question relative aux
determinants, Bulletin Sci. Math. (2), 17 (1893), 240-348, and Exercise
154 Chapter 10: Orthonormal systems

13 is from L. Fejer, Sur les fonctions bornies et integrables, Comptes


Rendus, 131 (1900), 984-7 and Investigations of Fourier series (in Hun-
garian), Mat. es Fiz. Lapok, 11 (1902), 49-68; see also Untersuchungen
uber Fouriersche Reihen, Math. Annalen, 58 (1904), 51-69.
11. ADJOINT OPERATORS

In the next four sections we give a brief account of the theory of


bounded linear operators on Banach spaces. Our aim is to present
several general concepts and prove some of the fundamental results.
We are mainly interested in the spectral theory, to be treated in three
sections, but before we embark on that, in this chapter we study the
basic properties of the adjoint of an operator.
Throughout this chapter we shall use the product notation for the
value of a linear functional on an element: (x,f) = (f,x) = f(x) for x in
a vector space V and / i n V , the dual of V. In particular, if A' is a
normed space and X* is the dual of X, i.e. the Banach space of all
bounded linear functional on X, then (•, •) is the bilinear form on
XXX* given by (xj) = /(*). Thus

(\x + ny,f) = \(x,f)+ti(y,f) and (x,Af+fig) = A <*,/>+/*<*,£>.


(Note the absence of conjugates: (•, •) is a bilinear form and not a her-
mitian form. This is why it is not confusing to have (x,f) = (/,*).)
Recall that |<JC,/)| ^ ||x||||/|| for all x E Xandf E X*. Furthermore,

||jc|| = sup{|(x,/)| : / E B(X*)} = max{|<*,/>| : / E B(X*)}


and
Il/H = sup{\(x,f)\:x(EB(X)}.
Let X and Y be normed spaces. Recall from Chapter 1 that the
adjoint or dual T* of an operator T E 28(Jf, Y) is the unique map
T* : Y* -+ X* such that

(x,T*g) = (Tx,g) (1)

155
156 Chapter 11: Adjoint operators

for all x E X and g E Y*. Indeed, for g £ 7*, we have a function T*g
on X defined by (T*g)(x) = (Tx,g); it is easily seen that this function is
linear and, in fact, T*g E Y*. In turn, the second dual (T*)* of T is the
unique map T** : X** -» Y** such that

for all <p e r a n d # E Y*.


Let us summarise the basic properties of taking adjoints.
Theorem 1. Let X and Y be normed spaces and let T, Tl9T2E ^(X, Y).
Then
(a) r E SXY*,**) and ||r|| = ||r||;
(b) for scalars Ax and A2, we have (A1T1 + A 2 r 2 )* = A 1 7 7 + A 2 7 A ;
(c) with the natural inclusions X C X** and y c Y * * we have
r**|^ = T, i.e. T**JC = 7* for all x E X;
(d) if Z is a normed space and 5 E SS(Y,Z) then (57)* = T*S*;
(e) if Tis invertible, i.e. T~l E <&(Y,X) then T* is also invertible and
1 1

Proof. Part (a) is Theorem 3.9, part (b) is immediate from the definition
of the adjoint, and (c) follows from Theorem 3.10, giving the natural
embeddings X-* X** and Y-> Y**. To see (d), note that for x E X
and h E Z* we have
(x,T*S*h) = (Tx,S*h) = (STx,h).
Finally, (e) follows since T~lT = Ix, where Ix is the identity operator
on X and so 1^ = {T~lT)* = r ^ r " 1 ) * . Similarly, IY* = ( r " 1 ) * r .
Hence (T*)" 1 = (7" 1 )*. •

Let now H and K be Hilbert spaces, with their inner products written
as (•, •) • We know from the Riesz representation theorem in Chapter 8
that a Hilbert space is naturally isometric with its dual but the isometry
is an anti-isomorphism. Thus x EH can be considered as a functional
on H: it acts on H as multiplication on the right: (-,*). Then for
T E <&(H,K) we have 7* E <&(K,H).
Equivalently, T* is defined by
(x,T*y) = (Tx,y) (2)
for all x EH and y E K. Indeed, for a fixed y E K the function
/ : if—» C defined by/(x) = (Tx,y) is a bounded linear functional on H.
Hence, by the Riesz representation theorem, f(x) = (x,u) for some
Chapter 11: Adjoint operators 157

unique vector u E H. We define T*y to be this vector u. It is immedi-


ate that T* is a linear map from # to H. Since
I/Ml ^ ||7*|||b|| ^ n 71 Ml I'll,
we have

IN =
and so T* is a bounded linear map and ||r*|| ^ ||r||.
Theorem 2. Let H and K be Hilbert spaces and let T, T1,T2E <&(H,K).
Then T* e <3&(K,H), \\T*\\ = \\T\\ and, for all scalars Ax and A2 we have
(3)
(4)
and
T** = T. (5)
Furthermore, if K = H then
2
f (6)
Proof. We known that ||T*|| = | | 7 | and (TiT2)* = TZTf; furthermore
T** = T since H is reflexive. To see (3), note that for JC £ H and
y £ K we have

= A1(T1A;,>')+A2(r2j«;,v)

Of course, this is also clear from Theorem 1 (b) and the fact that the
natural isometry between a Hilbert space and its dual is an anti-
isomorphism.
Finally, let TE<&(H). Then
||r|| 2 = sup||7i|| 2 = sup I (7*, 7*) I

II7*||||71| = ||T||2. D
158 Chapter 11: Adjoint operators

Let us emphasize again that in (3) we have to take the conjugates of


the coefficients, unlike in Theorem l ( b ) , which is the analogous state-
ment for normed spaces. This is because in (3) we take If as a map
from K to H, rather than from K* to H*, as in the normed-space case;
the conjugates appear because H and H* are identified by an anti-
isomorphism. Theorem 2 is of special interest in the case H = K. An
involution on a complex Banach algebra is a map x »-• x* such that (3),
(4) and (5) hold:
=
* Ax**+A2*2 ,
A C*-algebra is a Banach algebra with an involution satisfying (6), i.e.
in which \\xx*\\ = ||JC||2 (and so ||JC||2 = \\x*x\\ = \\x*\\2). Thus Theorem 2
claims that 96(//) is a C*-algebra with involution T*-* T*. Hence every
norm-closed subalgebra of 26(/f) which is closed under taking adjoints
(i.e. every closed *-subalgebra of 26(//)) is also a C* -algebra. What is
remarkable is that the converse of this statement is also true: every C*-
algebra is isomorphic to a closed *-subalgebra of 2S(/f). This is the
celebrated Gelfand-Naimark theorem; although it is beyond the scope
of this book, we shall present some exercises concerning it at the end of
the next chapter. Given a normed space X with dual X*, for a subspace
K C X define the annihilator of K in X* as

K° = {/ E X* : (x,f) = 0 for all x E K}.


Similarly, for a subspace L C X*, the annihilator of L in X (or the
preannihilator of L) is
°L = {x E X: (xj) = 0 for a l l / E L}.
Strictly speaking, K° usually denotes the polar of a set K C X:
K° = {f<EX*: \(x,f)\ ^ 1 for all/ E L}
and °L is the prepolar of a set L C A* (or the polar of L in X):
°L = {xEX: \(xj)\ ^ 1 for a l l / E } .

Of course, if K and L are subspaces, as we have chosen them, or at


least if they are unions of one-dimensional subspaces, then the two
definitions coincide (see Exercise 1).
It is clear that for any sets K C X and L C X* the annihilators K°
and °L are closed; furthermore,
AT0 = (linAT)0 = (fin AT)0 and °L = °(linL) = b (finL).
Chapter 11: Adjoint operators 159

Theorem 3. Let X and Y be normed spaces and let T G 28(Jf, Y). Then
Ker T = °(Im T*) and Ker T* = (Im T)°. •
Proof. Clearly
KerT= {xEX*: Tx = 0}
= {x G * : <7*,g> = 0 for all g G Y*}
= {xEX:(x, T*g) = 0 for all g <E Y*} = °(Im 7*).
Similarly,
Ker T* = {g G Y*: T*g = 0}
= fe G y*: <*, r g > = 0 for all * G * }
= fe G 7* : <rjc,g> = 0 for all x G X} = (Im T)0. •

Note that if L C X* then ° L C Z and L° C X** and so, in general,


we cannot expect °L to be equal to L° under the natural inclusion
X C X**. However, if X if reflexive and so X is identified with X**
then for every set L C I * we have °L = L°. If / / is a Hilbert space
then not only is H a reflexive space but also the dual H* is identified
with H. With this identification, L° = °L = LL for every set L C H.
Hence Theorem 3 has the following immediate consequence.
Corollary 4. Let H and # be Hilbert spaces and let T E <3&(H,K).
Then
Ker T = (Im T*)± and Ker T* = (Im D ± . •

It is worth noting that Im T = °(Ker T*) does not hold in general


since Im T need not be closed. However, if Im T is closed then we do
have Im T = °(Ker T*) (see Exercise 3).
Let H be a Hilbert space. An operator T G 26(//) is called hermitian
or self-adjoint if T* = T. Thus T is self-adjoint iff
(7x,.y) = (x, Ty) for all *,>> G //.
Clearly an operator T G 28(//) is self-adjoint iff (x,y) = (Tx,y) is a her-
mitian form on H. If 5 and r are commuting self-adjoint operators then
ST is also hermitian since
(STx,y) = (Tx,Sy) = (x,TSy) = (x,STy).
160 Chapter 11: Adjoint operators

In particular, if T is self-adjoint then so is Tn for every n ^ 1. Note


also that if T is self-adjoint then by Theorem 2,
||r2|| = ||rr|| = ||7f, . ||r4|| = ||r2||2 =
etc. Therefore \\T2k\\ = ||r||2*. Also, if 1 ^ n ^ 2k then

If T is self-adjoint then (Tx,x) is real for every x E H since


(Tx,x) — (x, Tx). We call a self-adjoint operator T positive if (7JC,JC) ^
0 for every x E H.
Note that, for T E $(//), the operator T*T is a positive self-adjoint
operator. Indeed,
(T*T)* = r*(r*)* = T*T and (T*Tx,x) = (Tx,Tx) = \\Tx\\2 ^ 0.
Replacing T by r*, we see that TT* is also a positive self-adjoint opera-
tor.
Theorem 5. Let H be a complex Hilbert space. Then every operator
T E <&(H) has a representation in the form T = 7\ + ir 2 , where Tx and
T2 are hermitian, and this representation is unique.
Proof. Set
7i = $(T+T*) and T2 = ~h(T-T*).
Then 7\ and T2 are hermitian and T = 7\ + ir 2 . The uniqueness follows
from the fact that if 7\ and T2 are hermitian and 7\ + ir 2 = 0 then

and so T2 = 0 and 7X = 0. •

Examples 6. (i) Let T be the right translation on /2, i.e. let


T((x1,X2,...)) = ((0,xl9x2,...)). Then r* is the left translation:
r((jc!,jc2,...)) = ((x2,x3,...)). cieariy ||r|| = | | r | | = i, r r is the
identity / but TT* ^ I:
TT*((xux2,x3,...)) = ((0,JC 2 ,JC 3 ,...)).
(ii) For <p E C[0,1] define T9 : L2(0,1) -> L2(0,1) as multiplication
Chapter 11: Adjoint operators 161

Then
T; = Tj and | | r j = ||7^|| = mzx{\<p(t)\ : 0 ^ ^ 1}.
Clearly 7^ is a positive hermitian operator iff cp is a non-negative real-
valued function.
(iii) Let M be a closed subspace of a Hilbert space and let PM be the
orthogonal projection onto M. Then PM is a positive self-adjoint opera-
tor and (as every projection) satisfies P ^ = PM. •

It is easily seen that the properties in Example 6 (iii) characterise the


orthogonal projections in a Hilbert space.
Theorem 7. Let / / b e a Hilbert space and let P E 2S(//) be a self-
adjoint projection: P2 = P = P*. Then M = ImP is closed and P is the
orthogonal projection of H onto M: P = PM.
Proof. Since P is a projection, we have x-Px E KerP and x =
{x-Px) + Px for every x E H. So H = Ker P +ImP. By Corollary 4,
we have KerP = (ImP*) 1 = (ImP) 1 and so H is the orthogonal direct
sum of Ker P and Im P. •

In addition to hermitian operators and orthogonal projections, let us


introduce two other important classes of operators. Given a Hilbert
space H, an operator T G 28(#) is said to be normal if TT* = T*T, and
unitary if T is invertible and its inverse is T*. Note that every hermitian
operator is normal, and so is every unitary operator. In the following
results we characterize normal and unitary operators.
Theorem 8. Let H be Hilbert space and T E 38(i/).
(a) Tis normal iff ||7JC|| = ||r*Jt|| for all x E H.
(b) If Tis normal then ||r"|| = \\T\\n for every n & 1 and
K e r T = Kerr* = (ImT) 1 = ( I m T ) 1 .
Proof, (a) Clearly,
|| 7JC||2— II 7^*JC||2 = (Tx,Tx)-(T*x,T*x)
= (T*Tx,x)-(TT*x,x) = ((T*T-TT*)x,x).
From Theorem 9.2 we know that T*T- TT* = 0 iff ((T*T- TT*)x,x) =
0 for every x.
(b) If T is normal then, by (a), we have Ker T = Ker T*. Hence, by
Corollary 4,
162 Chapter 11: Adjoint operators

ImT 1 = KerT* = KerT= (ImT) 1 .


Furthermore, as T*T is hermitian,
\\Tfn = \\T*T\\n = \\(T*T)n\\,
and so
imi2" = wrT)»\\ * i K m i m ^ imnirii
implying | | r | | = || T\\\ D

As a consequence of Theorem 8 one can see that Theorem 7 can be


strengthened: if a projection is normal then it is an orthogoanl projec-
tion (see Exercise 7).
Theorem 9. Let H be a Hilbert space and let U E 2S(//) be such that
Im U = H. Then the following are equivalent:
(a) U is unitary;
(b) U is an isometry: ||£/*|| = ||JC|| for every x E H;
(c) U preserves the inner product: (Ux, Uy) = (x,y) for all x,y E H.
Proof. The polarization identity (3) in Chapter 9 implies that U is an
isometry iff it preserves the inner product. Thus (b) and (c) are equiva-
lent.
If U is unitary then
(Ux9Uy) = (U*Ux,y) = (x,y).
Conversely, if (Ux, Uy) = (x,y) for all x,y E H then (U*Ux,y) = (x,y)
and so U*U = 1. Since Im U = H and U is an isometry, U is invertible.
Therefore U* = U'1. •

Our last aim in this chapter is to show that the converse of Theorem
1 (v) also holds if A" is a Banach space, i.e. T* is invertible iff T is. First
we note a simple condition for invertibility. Call T E 2S(A", Y) bounded
below if || 7*|| ^ e||jt|| for all x E X and some e > 0.
Theorem 10. Let Z b e a Banach space, Y a normed space, and let
1 6 » ( * , Y). Then T'1 E &(Y,X) iff ImT is dense in Y and T is
bounded below.
Proof. The necessity of the two conditions is obvious. Suppose then
that the conditions are satisfied. If T is bounded below then it is injec-
tive, so T'1 E i£(Z,X), where Z = ImT. Since Z is dense in Y, for
Chapter 11: Adjoint operators 163

every y E Y there is a sequence {zk)°\ in Z converging to y. Then by


the second condition, (T~1zk)™ is also convergent, say to x. Hence
Tx = lim^oo T{T~lzk) = l i m * ^ zk = y
and so Y = Z.
This shows that 7 " 1 E ^(y,JT). If ||7i|| ^ e\\x\\ for all x E X then,
clearly, HT"1!! ^ 1/e. D

Theorem 11. Let X be a Banach space, Y a normed space and let


T E » ( * , y). Then T* is invertible iff T is.
Proof. We have seen that if T is invertible then so is T*. Suppose then
that 7* is invertible. Let us check that the two conditions in Theorem
10 are satisfied.
By Theorem 3 we have (Im T)° = Ker T* = (0) and so Im T is dense
in y. To see that T is bounded below, let x E X and let / be a support
functional at x, i.e. let / E X* be such that (*,/> = ||x|| and ||/|| = 1.
Then
Ml = <*,/> =

^ lliiWKrr1/!! ^ ll
Therefore ||7JC|| ^ IKr*)"!""!*!!, completing the proof. D

If Im T is not dense in Y, say TX C Z, with Z a closed subspace, then


Z C Ker/ for some / E Y* (f f 0). Hence (x, T*f) = {TxJ) = 0 for
all x E X and so 71*/ = 0. In particular, if T* is bounded below then
Im T is dense in Y. This gives us yet another condition for invertibility;
let us state it together with Theorems 10 and 11.
Theorem 12. Let X be a Banach space, Y a normed space, and let
T E <&(X, Y). Then the following conditions are equivalent:
(a) Tis invertible;
(b) T* is invertible;
(c) Im T is dense in Y and T is bounded below;
(d) T and T* are both bounded below. •

The question of invertibility brings us to the study of the spectrum of


an operator and the structure of the algebra of bounded linear opera-
tors. But that requries a new chapter.
164 Chapter 11: Adjoint operators

Exercises

1. Given a Banach space X and a set K C X, the annihilator of K in


X* is
Ka = {f<EX*: <*,/> = 0 for all * E AT}
and the annihilator of a set L C A'* m ^ (or the preannihilator of
L)is
a
L = {xGX: <*,/> = 0 for a l l / E L}.
Show that if K and L are subspaces then Ka = K° and aL = °L,
where K° is the polar of K and °L is the prepolar of L. Show also
that
Ka = (UnK)a = QinK)a = (UnK)°
and
a
L = a(linL) = a(finL) = °(li
2. Give examples showing that for a Banach space X and a subspace
LCX*, the sets °L and L° need not be equal under the natural
inclusion XCX**.
3. Let X and Y be normed spaces and T E S8(Ar, Y). Show that
°(Ker 7*) is the closure of Im T.
4. A subspace U of a normed space V is said to be an invariant sub-
space of an operator S E a(V) if S£/ C [/, i.e. 5w E [/ for all
u E U. Let Z b e a normed space and T E 26(Jf). Show that a
closed subspace Y of X is an invariant subspace of T iff Y° is an
invariant subspace of T*.
5. Let X and Y be Banach spaces and T E <$>{X, Y). Prove that Im T
is closed iff Im T* is closed.
6. Let Z b e a Banach space. Show that for T E &(X) the series
2 ~ = 0 ^"/ r t • converges in norm to an element of ^(X), denoted by
exp7\ Show also that (expT)* = expT* and if S E &(AT) com-
mutes with T then

(expSXexpT) = (exp r)(exp5)exp(5+ T).

In the exercises below, H denotes a Hilbert space.


Chapter 11: Adjoint operators 165

7. Let T be a bounded linear operator on a Hilbert space H. Show


that T has an eigenvector iff T* has 1-codimensional closed invari-
ant subspace.
8. Let / / b e a Hilbert space and P E 28(//) a projection: P2 = P.
Show that the following are equivalent:
(a) P is an orthogonal projection;
(b) P is hermitian;
(c) P is normal;
(d) (Px,x) = \\Px\\2 for all x G H.
9. Let U E 26(//) be a unitary operator.
(i) Show that Im( £/-/) = Im(i/*-/) and deduce that
\
(ii) For n ^ 1 set

5n = -
n
Show that £ „ * - • P M * for every x E H, where M = Ker((/-/).
(One expresses this by saying that Sn tends to PM in the'strong
operator topology.)
10. Show that if T E 26(//) is hermitian then expir is unitary.
11. The aim of this exercise is to prove the Fuglede-Putnam theorem.
Suppose that R,S,TE 28(//), with R and T normal and RS = ST.
(i) Show that
(expfl)S = S(expT).
(ii) By considering exp(R*-R)Sexp(T- T*), show that

||(exp/?*)Sexp(-r)ll ^ l|5||.
(iii) F o r / e (26(/f))* and A G C set
F(A) = /(exp(Afl*)Sexp(-Ar*)).
Show that F(A) is an analytic function and |F(A)| ^ ||/||||5|| for
every A G C. Apply Liouville's theorem to deduce that F(A) =
F(0) = f(S) for every A and hence that
exp(A/?*)S = 5exp(Ar*).
(iv) Deduce that R*S = ST*.
166 Chapter 11: Adjoint operators

Notes

The notion of an adjoint operator was first introduced by S. Banach,


Sur les fonctionelles lineaires II, Studia Math., 1 (1929), 223-39. Our
treatment of adjoint operators is standard. The Gelfand-Naimark
theorem was proved in I. M. Gelfand and M. A. Nalmark, On the
embedding of normed rings into the ring of operators in Hilbert space, Mat.
Sbornik N.S., 12 (1943), 197-213; for a thorough treatment of the subject see S.
Sakai, C*-algebras and W*-algebras, Springer- Verlag, New York-Heidel-
berg-Berlin, 1971. For the Fuglede-Putnam theorem in Exercise 11, see
chapter 41 in P. R. Halmos, Introduction to Hilbert space and the theory of
spectral multiplicity, Chelsea, New York, 1951.
12. THE ALGEBRA OF BOUNDED
LINEAR OPERATORS

In this chapter we shall consider complex Banach spaces and complex


unital Banach algebras, as we shall study the spectra of various ele-
ments. Recall that a complex unital Banach algebra is a complex alge-
bra A with an identity e, which is also a Banach space, in which the
algebra structure and the norm are connected by ||e|| = 1 and \\ab\\ ^
||a||||6|| for all a,b E A. If there is an involution x •-> x* in A such that
x** = x, (x+y)* = x*+y*, (Ax)* = Ax*, (xy)* = y*x* and ||JC*X|| =
||JC||2 then A is a C*-algebra. As we noted earlier, if X is a complex
Banach space then 98 (JQ is a complex unital Banach algebra, and if H is
a complex Hilbert space then 28(//) is a C*-algebra.
An element a of a Banach algebra A is invertible (in A) if
ab = ba = e for some b E A; the (unique) element 6 is the inverse of
a, and is denoted by a'1. The spectrum of a E A is
<rA{a) = SpA(a) = {X E C: Xe-ais not invertible in v4},
and the resolvent set of a is 8A(a) = C W ^ a ) . A point of 8A(a) is said to
be a regular point. The function /? : 8(a) —» ^4 given by /?(A) =
(Ae — a ) " 1 is the resolvent of a. The element /?(A) is the resolvent of a at
A or, with a slight abuse of terminology, the resolvent of a.
The prime example of a Banach algebra we are interested in is the
algebra ^(X) of bounded linear operators on a complex Banach space
X; so our algebra elements are operators. In view of this, if T E 26(X)
then we define the spectrum and resolvent set of T without any reference
to»CY):
= {A E C : XI- T is not invertible}
where / is the identity operator on X, and
p(T) = C\a(T).

167
168 Chapter 12: The algebra of bounded linear operators

If A is a complex unital Banach algebra then A can be considered to


be a subalgebra of 2&G4), the algebra of all bounded linear operators
acting on the Banach space A, with the element a corresponding to the
operator La of left multiplication by a (so that a*-+ La, where
La(x) = ax for every x E A). In particular if a E A is invertible then so
is La E 9&G4), with inverse La-K Conversely, if S E ^(A) is the inverse
of La, so that
x = (LaS)x = a{Sx)
for every x E A, then with b = Se we have ab = 1 and so a{ba-e) =
(ab)a-a = 0. Hence ba-e E KerL a and so ba = e. Thus 6 is the
inverse of a. Also, Ae-tf is invertible iff A/-L f l is invertible. Hence

Although the spectrum of an operator T E $h{X) depends only on


how T fits into the algebraic structure of 98 CX), it is of considerable
interest to see how the action of T on X affects invertibility. In particu-
lar, we may distinguish the points of a{T) according to the reasons why
XI— T is not invertible.
What are the obstruction to the invertibility of an operator 5 E 28(JQ?
By the inverse-mapping theorem, S is invertible iff KerS = (0) and
imS = X. Thus if S is not invertible then either KerS ± (0) or
ImS j= X (or both, of course). Of these, the former is, perhaps, the
more basic obstruction.
Accordingly, let us define the point spectrum of T E 28(X) as
ap(T) = { A E C : K e r ( A / - r ) ^ (0)}.
The elements of ap(T) are the eigenvalues of T; for an eigenvalue
A E o"p(r), the non-zero vectors in Ker(A/— T) are called eigenvectors
with eigenvalue A. Furthermore, Ker(A/- T) is the eigenspace of T at A.
Clearly o-p(r) Ccr(T).
If X is finite-dimensional and S E 3i(X) then the two conditions
KerS = (0) and ImS = X coincide. Hence a{T) = ap(T) for every
operator on a finite-dimensional space.
However, if X is infinite-dimensional then we may have KerS = (0)
and ImS ^ X, so the point spectrum need not be the entire spectrum.
More precisely, by Theorem 11.10 , A E a(T) iff either Im(A7- T) is not
dense in X or A / - T is not bounded below: there is no e > 0 such that
||(A/-r)jc|| ^ €||*|| for every x EX. In the former case A is said to
belong to the compression spectrum crcom(T), and in the latter case, A is
Chapter 12: The algebra of bounded linear operators 169

said to belong to the approximate point spectrum of T, denoted by


o"ap(r). In other words,
<TaP(T) = {A E C: there is a sequence (xn) C S(X)
such that (A/- T)xn -> 0}.
Sometimes (xn) is called an approximate eigenvector with eigenvalue A.
Clearly ap(T) C crap(D and
*{T) =crap(T)Uacom(T).
Sometimes the points of the spectrum are classified further: the resi-
dual spectrum is <JX{T) = o-com(T)\crp(T) and the continuous spectrum is
aG(T) = cr(T)\((Tcom(T)U*p(T)). Thus
<T{T) =(rp(T)U(Tc(T)UaT(T),
with the sets on the right being pairwise disjoint.
It is immediate from the definition that the approximate point spec-
trum crap(r) is a closed set; the point spectrum crp(T) need not be
closed.
We shall show that cr(T) is a non-empty closed subset of the disc
{A E C: |A| ^ ||r||}. The latter assertion is an immediate consequence
of the following simple but important result.
Theorem 1. Let T E »(AT) satisfy ||r|| < 1. Then 1 £ cr(T) and
00
(T T\ ~~ 1 ^ irk
\l l ) — /, l ,
A:=0
where the series on the right is absolutely convergent in the Banach
space ^(X).
Proof. Note that

k=0 k=0
k
and so 2 ^ = 0 T is absolutely convergent.
Hence
00

A:=0

and, similarly,
/ oo \
: Tk\(I-T) = I. D
170 Chapter 12: The algebra of bounded linear operators

Theorem 2. Suppose that S,TE ^{X), T is invertible and | | S - r | | <


||T~l\\~l. Then S is invertible and
Hj-t r-m^ \\T-Y\\S-T\\

Proof. By Theorem 1, we have


00

and so

(i)
and
||7^-l||2|| T _ ell
_ T II ^ 2 II Li U-3)J / II ^

Corollary 3. For a Banach space X, let


%X) = {T e »(J0 : T is invertible}.
Then ^CY) is an open subset of 98(AT). Furthermore, ^ W , with the
topology inherited from ^(X) and operator multiplication, is a topologi-
cal group: it is closed under multiplication, the multiplication is jointly
continuous and the map T'*-* T~l is a homeomorphism of ^(X) onto
itself. •

For operators of the form A o / - T and A/— T, identity (1) takes a


somewhat simpler form. Namely, if

then

i?(A) = (A/-7T 1 = 2 (Ao-AHAoZ-r)-"-1 = £ (A0-\)nR(A0)n+1


n=0 n=0
(2)
with the series converging in operator norm.
Theorems 1 and 2 have the following immediate consequence.
Corollary 4. The spectrum cr(T) is a closed subset of the disc
{ A £ C : |A|*s||r||}. If |AI > ||71| then
Chapter 12: The algebra of bounded linear operators 111

R(k) = (XI- T)-1 = A W / - y ) = ^ . (3)

and

H/?(A)H = i
Note that we do not yet know whether the spectrum a(T) can be
empty or not. In proving that it cannot, we shall make use of Banach-
space-valued analytic functions, an example of which we have just seen
in (2).
Given a Banach space X and an open set D C C, a function
F: D -> X is said to be analytic if for every z0 G D there is an
r = r(z0) > 0 such that D(zo,r) = { z G C : \z-zo\ < r} C D and

Hz) = i an(z-z0)n (4)


/i=0

for some a0, ax,... E ^ and all z G £>(z0, r), with the series in (4) being
absolutely convergent. Thus (2) shows that the resolvent R : p(T) —>
39(^0 is an analytic function, with values in the Banach space ^(X).
The standard results concerning analytic functions remain valid in this
more general setting. For example, as we noted in Exercise 10 of
Chapter 11, the analogue of Liouville's theorem holds: a norm-bounded
entire function in constant. To spell it out: if F: C —> X is analytic and
||F(z)|| ^ M for some M < & and all z G C then F{z) is constant.
Indeed, for/ G A'* the function
g(z) = f(F(z))
is a (complex-valued) analytic function and so it is constant: g(z) = g(0)
for all z. Hence /(F(z)-F(O)) = 0 for all f E X* and so, by the
Hahn-Banach theorem, F(z) = F(0) for all z.
Also, the radius of convergence of (4) is liminfn_>oo||flj|~1/'1, just as in
the classical case. Equivalently, the Laurent series

G(z) = i bnz~n (5)


/i=0

has radius of convergence s = Umsupn_Mo||&/il|1//l.


Indeed, if \z\ > s then (l + e)^,,!!17'1 < \z\ for some e > 0 and every
sufficiently large n. Hence ||6nz~w|| < (l + e)~ n if n is sufficiently large,
implying that (5) is absolutely convergent.
172 Chapter 12: The algebra of bounded linear operators

Conversely, if \z\ <s then there is an infinite sequence nx <


n2< ••• such that \\bnk\\ > \z\"k. But then ||fcn;fcz~n*|| > 1 and so (5)
does not converge.
The analytic functions we shall consider will take their values in ^(X)
or, more generally, in a Banach algebra. If Ft: D(-^ A (i = 1 , 2 ) are
analytic functions into a Banach algebra A then F1F2 : D1HD2-^ A is
also an analytic function. Furthermore, if

Fi(z)= I an(z-z0)", F2(z)= i bn(z-z0)n for zED(zo,r),

then
oo n
F1(z)F2(z) = 2 cn(z-z0)n, where cn = 2 «*&«-*•

These observations more than suffice to prove the main results of the
chapter.
Theorem 5. The spectrum of an operator T E ^{X) is not empty.
Proof. Suppose that a(T) = 0 . Since, as we remarked earlier, formula
(2) shows that R(X) is an analytic function on the resolvent, which is
now the entire plane, R(\) is an entire function. Furthermore, (3)
shows that ||/?(A)|| -> 0 as |A| -* °°. Hence, by Liouville's theorem,
R(\) is constant and, in fact, R(\) = 0 for every A. But this is clearly
impossible. •

For emphasis, let us put Corollary 4 and Theorems 2 and 5 together.


Theorem 6. For every complex Banach space X and operator
TE^(X), the spectrum a(T) is a non-empty closed subset of
{A E C: |A| ^ ||r||}. Furthermore, if A E p{T) and
d(X,a(T)) = min{|A-//,| : /x E a(T)} = d
then \\R(\)\\ ^ l/d. D

It is interesting to note that Theorem 5 gives an independent proof of


the fact that every nxn complex matrix has an eigenvalue. Note also
that the analogue of Theorem 5 fails for real spaces, as shown, for
example, by the rotation (ei,e2) l~^> (e2, —e^) in U2. The spectrum is
very useful precisely because it allows techniques of complex analysis to
be brought into operator theory.
Chapter 12: The algebra of bounded linear operators 173

From Theorem 6 it is a short step to show that the approximate point


spectrum is not empty either.
Theorem 7. The approximate point spectrum (rap(T) contains the boun-
dary da(T) of the spectrum.
Proof. Let A E dcr(T). Pick a sequence A 1? A 2 ,... E p(T) tending to A.
Then, by the second part of Theorem 6, ||/?(An)|| —> °°. Therefore there
is a sequence (yn)i C X such that yn —> 0 and ||^(An)>>J| = 1 for every
n. Setting xn = R{\n)yn, we find that ||jcn|| = 1 and
||(A/-DJCJ| ^ \\{knI-T)xn\\ + \\{\-\n)xn\\ = HjJ + lA-Aj - 0 .
Hence (xn)i is an approximate eigenvector with eigenvalue A. •

As an illustration of the concepts and results presented so far, let us


examine the spectrum of the right shift operator 5 on /p (1 ^ p ^ °°)
defined by S(xi,x2,...) = (0,Xi,x2,...). Since ||5|| = 1, the spectrum
a(S) is a closed non-empty subset of the closed disc A = {A E C: |A| ^
1}. Furthermore, ||5JC|| = ||JC|| for every x and so ||(5-A)JC|| ^
( 1 - |A|)||jc||, implying that aap(S) C dA = {A E C: |A| = 1}. By
Theorem 7 we have aap(S) = dA. (Of course, this is easy to check
directly.)
How much of the circle dA belongs to the point spectrum? Suppose
that Sx = AJC ^ 0, where A ^ 0. Then 0 = AJC1? XX = AJC2, X2 = AJC3,
..., implying that xt = 0, x2 = 0, x3 = 0, . . . . Hence ap(S) = 0 .
Therefore the spectrum CT(S) is the closed disc A, the approximate
point spectrum o-ap(5) is the circle dA and the point spectrum is empty.
What is the spectrum of a 'nice' function / of an operator? This is
easy to answer when / is a polynomial; an analogous result holds in a
much more general case, namely when / is an analytic function on an
open set containing the spectrum.
Theorem 9. Let p(t) be a polynomial with complex coefficients. Then
for T E 3h(X), the spectrum oip{T) is precisely
p{°{T)) = {p(A): A E *{T)}.
Proof. We may assume that the leading coefficient of p(t) is 1 and
p(0) = 0. Given Ao 6 C, let

P«-A o = fl (*-
k=l
Then
174 Chapter 12: The algebra of bounded linear operators

P(T)-\OI= n (T-nki).
k=\
This product fails to be invertible iff at least one of the factors, say
T-/AkI, is not invertible, i.e. /ik E cr{T). Since the fik are the zeros of
p(t) — Ao, this happens iff p(/i) = Ao for some /x E <r{T). D

The spectral radius of T E <&(X) is


r(T) = sup{|A| : A E a(T)} = max{|A| : A E a(T)}.
The spectral radius is a simple function of the sequence (||7"n||)50, as
shown by the following result, Gelfand's spectral-radius formula.
Theorem 8. For T E <&(X) we have
r(T) = li

Proof. By Theorems 6 and 7 we have


r(T)n = r(Tn) ^ \\Tn\\. (6)
n
Hence r(T) ^ liminf w _ 00 ||r f.
On the other hand, as p(T) D {A E C: |A| > r(T)}, relation (3) tells
us that the Laurent series
Tn

\n + l

is convergent for \X\ > r(T). Hence, recalling the formula for the
radius of convergence, we find that r{T) ^ limsupn_>oo||7""||1///l. •

It is easily seen that the spectral radius is an upper semicontinuous


function of the operator in the norm topology; in fact,
r(S+T) ^r(S) + \\ T\\
for all S,TE <&(X) (see Exercise 8).
It is worth recalling that all the results above are true for the spectra of
elements of Banach algebras, not only of elements of <3b(X). In the sim-
plest of all Banach algebras, C, every non-zero element is invertible. In
fact, C is the only Banach algebra which is a division algebra.
Theorem 10 (Gelfand-Mazur theorem) Let A be a complex unital
Banach algebra in which every non-zero element is invertible. Then
A = C.
Chapter 12: The algebra of bounded linear operators 175

Proof. Given a E A, let A E cr(a). Then A — a ( = XI—a) is not inverti-


ble and so A — a = 0, i.e. a = A. •

We know from Theorem 11.11 that T E <&(X) is invertible iff


T* E <&(X*) is invertible. Hence A / - T is invertible iff A / - T* is.
Therefore, recalling that for a Hilbert space //, the dual H* is identified
with H by an anti-isomorphism, we get that for TE ^(H), kl-T is
invertible iff (A/-7 1 )* = A / - r* is invertible. Finally, recalling from
Theorem 11.8 (b) that for a normal operator T we have || y**|| = ||r|| n for
n ^ 1, we have the following result.
Theorem 11.
(a) For a Banach space X and an operator T E 2&U0, we have
a(T*) = cr(T).
(b) For a Hilbert space H and an operator TES6(//), we have
<r(7*) = conjo-(r) = {A : A E a(T)}.
(c) If Tis a normal operator on a Hilbert space then r(T) = \\T\\. •

Let us introduce another bounded subset of the complex plane associ-


ated with a linear operator. Given a Banach space X and an operator
T E 26 CY), the {spatial) numerical range of T is
= {(Tx,f) :xEX,f EX*, \\x\\ = \\f\\ = f(x) = 1}.
With the notation used before Lemma 8.10,
V(T)={f(Tx):(xJ)ElI(X)}.
Thus to get a point of the numerical range, we take a point x of the unit
sphere S(X), a support functional f at x, i.e. a point of the unit sphere
S(X*) taking value 1 at or, and evaluate / a t Tx. It is clear that the
numerical range depends on the shape of the unit ball, not only on the
algebra ^(X). If T is an operator on a Hilbert space then V(T) is just
the set of values taken by the hermitian form (Tx,x) on the unit sphere:
V(T) = {(7JC,JC): ||JC|| = 1}. Nevertheless, the numerical range can be
easier to handle than the spectrum and is often more informative. It is
clear that, just as the spectrum, V(T) is contained in the closed disc of
centre 0 and radius ||r||. Even more, the closure of V(T) is sandwiched
between a(T) and this disc. But before we show this, we prove that
V(T*) can be only a little bigger than V(T).
Theorem 12. For a complex Banach space X and an operator
r e »CY), we have
176 Chapter 12: The algebra of bounded linear operators

V(T) C V(D C V(T),


where V(T) is the closure of V(T).
Proof. The first inclusion is easily seen since if x E S(X), / E S(X*)
and (xj) = 1 then x E S(X**), </,£) = 1 and (TxJ) = (x,T*f) =
(T*f,x), where, as earlier, x denotes x considered as an element of X**.
To see the second inclusion, let £ E V(T*) so that there are / E S(X*)
and <p E S(X**) such that </,^> = 1 and f = <r*/,^). Let 0 < e < 1.
Since B(Jf) is w*-dense in B(X**), there is an x E S(X) such that
|</,«p-*>l < W and |<r/,9-i>| < h2-
Then, by Theorem 8.11, there exist g E S(X*) and >> E 5 W such that
(y,g) = 1, ||x-y|| < € and ||/-g|| ^ e. Hence
|<7t*-y),/>| < €||71| and ||<7>,/-g)|| ^ 6||r||.
But this implies that V(T) has a point close to £, namely the point
(Ty,g) £ V(T):
\<Ty,g)-€\ = \{Ty,g)-(rf,<p)\
^ \(Ty,g)-(T*f,x)\ + le2
^\(Ty,g)-(Ty,f)\+e\\T\\ + ±e2

Theorem 13. For a complex Banach space X, the spectrum of T is con-


tained in the closure V(T) of the numerical range V(T).
Proof. Suppose that
d(\,V(T)) = inf{|A-/x| : /x E V(T)} = d > 0.
By Theorem 12 we also have d(A,K(r*)) = d. To prove that a{T) C
F(7), we have to show that \I-T is invertible.
Given x E S(X), pick a support functional / E 5(X*) at x, i.e. a
norm-1 functional with <*,/> = 1. Then (7JC,/) E K(T) and so

Hence

for all xEX.


Chapter 12: The algebra of bounded linear operators 111

Similarly, as d(\, V(T*)) = d, we have

for all / G X*. Thus both A / - T and (A/- T)* are bounded below. But
then, by Theorem 11.12, A / - T is invertible. •

Another aspect of the connection between the spectrum and the


numerical range given in Theorem 13 is that coo-(r), the convex hull of
spectrum, is precisely f) co V(T), where the intersection is taken over all
numerical ranges V(T) with respect to norms on X equivalent to the
given norm. But we shall not give a proof of this result.
The rest of the chapter is about a striking application of the spectral-
radius formula to obtain a remarkable theorem related to material
beyond the main body of this book. The theorem is Johnson's
uniqueness-of-norm theorem, but the beautiful and unexpectedly simple
proof we present is due to Ransford.
Let us start with a classical inequality concerning complex functions,
namely Hadamard's three-circles theorem, stating that if / is analytic in
the annulus Rx < \z\ < R2 then Mr = max| z | = r |/(z)| is a convex func-
tion of logr for R1 < r < R2. Thus if f(z) is analytic in the open disc
\z\ < Ro, where Ro > 1, then
| / ( 1 ) | 2 « max|/(z)| max |/(z)| (7)
\z\=R \z\ = \/R
for all R ( 1 < R < Ro).
To see (7), note simply that g(z) = /(z)/(l/z) is analytic in (an open
set containing) the annulus \/R ^ \z\ ^ R and so attains its maximum
there on the boundary. In particular,
|/(1)| 2 = |g(l)| =£ max|g(z)| ^ | 2 maxj/(z)|.

A similar inequality holds for Banach-space-valued analytic functions,


with the norm replacing the modulus. In particular, if f(z) is analytic in
the open disc \z\ < Ro, with values in a Banach space X then
||/(1)||2 ^ max||/(z)|| max ||/(z)||. (8)
\z\=R \z\ = l/R
for all R ( 1 < R < Ro). Indeed, let f(z) = Y^=o anzn (an E X) and let
cp E S(X*) be a support functional at /(I) = 2 ^ = 0 an '-

<P 2 •
n=0 / n=0 n=0
178 Chapter 12: The algebra of bounded linear operators

Set

g(z) = <p(f(z)) = i cp(an)zn.


n=0
Then, by (7)
||/(1)||2 = \g(l)\2 ^ max |g(z)| max \g{z)\
\z\=R \z\ = l/R
= maxkC/Cz))^ jn*xR\<p(f(z))\

= maxJ ||/(z) || maxlU ||/(z)||,


\z\=R "\z\ = l/R " '

proving (8).
In fact, if we have an analytic function with values in a Banach alge-
bra then in inequality (8) we may replace the norm by the spectral
radius.
Lemma 13. Let f(z) be an analytic function in the open disc |z| < Ro,
with values in a Banach algebra A. Then
r(f(Y)f < maxr(/(z)) max r(f(z)).
\z\-R \z\-l/R
Proof. From the spectral-radius formula, we know that ||/(z) 2 ||2 is
monotone decreasing to r(f(z)), and r(f(z)) is a continuous function of
z for \z\ = R and |z| = l/R. Consequently, by Dini's theorem
(Theorem 6.5), for every e > 0 there is an n such that
max \\f{z)r\\2'n max ||/(z) 2 l 2 ' n ^ max {r(/(z))+ 6} max {r(/(z)) + e}.
]U
\z\=R " \z\ = l/RlU " \z\=Rl KJ }
\z\ = l/R V
'

Applying Theorem 9 and inequality (8), we find that


r(f{\)f ^ \\f(z)2Yn
^ maxH^z)2!2"" max ||/(z)2l2~"
\z\=RlU " \z\ = l/RlU "
{K/()) }| =

As this holds for every e > 0, the result follows. •

The radical Rad B of a complex (unital) Banach algebra B is the inter-


section of all the maximal left ideals of B. The following lemma relates
the radical to the spectral radius.
Lemma 15. If b E B is such that r{b'b) = 0 for all b' E B then
b E Radfi.
Chapter 12: The algebra of bounded linear operators 179

Proof Suppose that b fc RadS, that is b £ L for some maximal left


ideal L. Then Bb + L is a left ideal properly containing L, and so
Bb + L = B. Hence e = b'b + l for some b' E B and I E L, where e is
the identity. But then e-b'b = / E L and so e-b'b is not invertible.
Therefore r(bfb) ^ 1. •

Now we are ready to give Ransford's proof of Johnson's theorem,


which is slightly more than the assertion that if Rad/?={0} then all
Banach-algebra norms on B are equivalent.
Theorem 16. Let A and B be Banach algebras, with RadB = {0}. Then
every surjective homomorphism 6 : A —> B is automatically continuous.
Proof. Suppose that an -» 0 in A and 0(art) —* b in S. By the closed-
graph theorem (Theorem 5.8) it suffices to show that b = 0. Since
Rad5 = {0}, this is the same as showing that b E RadS, and by
Lemma 15 this follows if we show that r{b'b) = 0 for all b' E B.
Let then b' E B. Pick a, a' E A with 0(a) = b and 6{a') = &'. Set
cn = fl'flrt, c = a'a, dn = 6{cn) and d = 6(c) = fe'6. Then cn -> 0 in A
and dn = 0(0,,) -* d in B. Define a linear function pn : C —> A by

and set
*)) = 0(cn)z+[6(c)-e(cn)l
Note that/? n (l) = c and ^(1) = rf.
Since a homomorphism does not increase the spectral radius.

Hence, by Lemma 14, for all n ^ 1 and i? > 1 we have


r ( < / ) 2 ^ (RWcJ + Wc-'
Letting n —> oo5 We find that

and so, letting i^ -> oo, we see that r(d) = r(bfb) = 0, as desired. •

Corollary 17. (Johnson's uniqueness-of-norm theorem) Let B be a com-


plex unital Banach algebra with norm |[-|| 0 and RadB = {0}. Then
every Banach algebra norm on B is equivalent to || • ||0. •
180 Chapter 12: The algebra of bounded linear operators

The algebra %(X) of bounded linear operators on a Banach space X


satisfies the conditions in Corollary 17, so all norms on 2S(X) turning it
into a Banach algebra are equivalent.

Exercises

In the exercises below, X is a complex Banach space and T E ^(X).

1. We call T a left (right) divisor of zero if there is an 5 E ^(X) such


that 5 ^ 0 and TS = 0 (ST = 0). Show that the point spectrum of
Tis
crp(T) = {A E C: A / - T is a left divisor of zero}
and the compression spectrum is
crcom(T) = {A E C: XI — T is a right divisor of zero}.
2. We call T a /e/if (ng/if) topological divisor of zero if there are
TUT2,... 6 « ( J ! 0 such that ||7;|| = 1 and 77; -> 0 (TnT^>0).
Show that the approximate point spectrum of T is
crap(r) = {A E C: XI— T is a left topological divisor of zero}.
3. Show that 71 is a right topological divisor of zero iff T* E ^(X*) is
a left topological divisor of zero.
4. Show that if A E cr{T) then A/— 71 is either a right divisor of zero
or a left topological divisor of zero. Deduce that
°{T) =crcom(T)U(Tap(T).
5. Let X be reflexive. Prove that if T is not invertible and is neither
a left nor a right divisor of zero then it is both a left topological
divisor of zero and a right topological divisor of zero.
6. Show that
^ap(^) = {A E C: A / - T* is not surjective}
and
crap(7^) = {A E C: A / - T is not surjective}.
7. Suppose that S,T E a (AT) commute: ST = TS. Show that
r(S + D ^ KS) + KF) and
Show also that these inequalities need not hold if 5 and T are not
assumed to commute.
Chapter 12: The algebra of bounded linear operators 181

8. Show that for S,TE <&(X) we have

Show also that if r(S) = 0 then r: <3l(X) -> [0,») is continuous at


5.
9. Suppose that r(T) < 1. Show that

IWIo= I \\Tnx\\
is a norm on X, equivalent to the original norm | • ||.
10. Prove that
r(T) = inf{||r||': I * II' is a norm on X, equivalent to | • ||}.
11. Check that the resolvent R(\) of T satisfies the resolvent identity
R(\)-R(ti) = ( M -
for all A,ii G p(T).
12. Check that if S, T G 8(JS0, A G p(ST) and A ^ 0, then A E p(TS)
and

Deduce that a(ST) U{0} = o-(rS)U{0}. Show also that a(ST) =


a(TS) need not hold.
13. Let A' be a non-empty compact subset of C. Show that K is the
spectrum of a Hilbert space operator: there is an operator
5 G 26(//), where H is a Hilbert space, such that cr(5) = A'.
14. Show that if T is a normal operator on a Hilbert space then
r(T) = | r||.
15. For w = (w*)? G /„ define 7^ : l2 -> /2 by rw(x) = ( w ^ ) J = 1 .
Show that Hr^ll = Halloo = sup^lw^l. Show also that eigenvalues
of Tare w1,w2,... and cr{T) = {wn}. What is 7^?
16. For w = (w^)" G /oo define Sw : /2 -> /2 by

Express 115^11 and the spectral radius K^w) in terms of the


sequence w.
17. Let A = {z G C: \z \ ^ 1} and H = L2(A). Let T G »(if) be the
operator of multiplication by z. Show that a(T) = A and r has no
eigenvalue.
182 Chapter 12: The algebra of bounded linear operators

18. Use the Hahn-Banach theorem to show that for T E <3l(X) we


have
sup Re V(T) = sup{c E IR: there is an x = x(c) E X (x f 0),
such that ||(1-re + rT)x\\ ^ ||jc||for allr ^ 0}.
19. An operator T E 9 ( 1 ) is said to be dissipative if sup Re V(T) ^ 0.
Show that if T is dissipative then

for all x E X and r ^ 0.


20. Show that if A E dco V(D and (A/- T)JC = 0 then

for all y E X. (Note that if T is dissipative and Tx = 0 then

for r > 0.) Deduce that if A E dc6V(T) and (\I-T)2z = 0 then


(\I-T)z = 0.
21. An operator T E <&(X) is said to be hermitian if V(T) C IR. (Note
that X is a Banach space; for a Hilbert space this definition coin-
cides with the usual (earlier) definition of a hermitian operator on
a Hilbert space.) Thus T is hermitian iff both \T and -\T are dis-
sipative. Prove that T is hermitian iff

||exp(ir)*|| = Ml
for all r E IR and J t 6 l , where

V 5*
expS = / —
" AC!
A:=0
for S E <&(X).
22. Let / / be a Hilbert space and let 5 C » ( / / ) be such that
V(S) C [0,oo). Show that

(Sxyy) ** (Sx,x)1/2(Sy,y)1/2
and deduce that

for all x,y £ / / . Deduce that if 0 £ V(5) then 0 £ a(S).


Chapter 12: The algebra of bounded linear operators 183

23. Show that if S is a hermitian operator on a Hilbert space then


V(S) = coo-(S).
(Note that the assertions in the last two exercises are easily deduced
from Theorem 11 (c) as well.)
24. Prove that the result in the previous exercise holds for a normal
operator 5 o n a Hilbert space.
25. Let M be a proper ideal of a unital Banach algebra B. Show that
the closure of M is also a proper ideal. Deduce that every maxi-
mal ideal of B is closed.
26. Let M be a maximal ideal of a complex unital Banach algebra B.
Show that B/M is also a complex unital Banach algebra.

The aim of the next five exercises is to prove the commutative


Gelfand-Naimark theorem. In these exercises A is a commutative com-
plex unital Banach algebra.

27. Show that if M is a maximal ideal of A then A/M is isometrically


isomorphic to C.
28. Let h : A —• C be a non-zero homomorphism. Show that \\h\\ = 1.
29. Let i t be the set of maximal ideals of A. By Exercise 27, i t may
be identified with the set of non-zero homomorphisms h : A —> C.
By Exercise 28, this is a subset of B(A*). Give i t the relative
weak* topology (i.e. the topology induced by the weak* topology
on A*). Endowed with this topology, we call M the maximal ideal
space of A. Prove that i t is a compact Hausdorff space.
30. For x E A the Gelfand transform of x is the function x : i t —> C
defined by x(h) = h(x), where h E i t is considered as a homomor-
phism h : A —> C. Show that the spectrum a(x) is the range of x.
31. Show that if A is a commutative unital C*-algebra with maximal
ideal space i t then the Gelfand transformation x »-> x maps ^4
isometrically and isomorphically onto C(it), the commutative C*-
algebra of continuous functions on the compact Hausdorff space
it.
32. Let lx(l) be the algebra of all (doubly infinite) complex sequences
x = (*„)!! oo such that

IWIi= 2 k,l<«,
n = — oo

with convolution product ;c*y = z, where


184 Chapter 12: The algebra of bounded linear operators

00

Show that /X(Z) is a commutative Banach algebra. Show also that


the maximal ideal space of /X(Z) can be identified with the unit cir-
cle T = {z G C: \z | = 1}, where z : lx(Z) -* C is defined by

n = —oo

33. Deduce from the result in the previous exercises that if

m = i xnent * o
for all t, and 2°° = _ 00 k« I < °°> then

7^= £ ^e1"', where £ b J < » .


34 + . Let if be a Hilbert space and let T e 2S(i/). Show that for every
e > 0 there is an invertible operator 5 E S6(/f) such that

Notes

The original reference to Gelfand's spectral-radius formula is I. M. Gel-


fand, Normierte Ringe, Mat. Sbornik N. S., 9 (51) (1941), 3-24; this is
also one of the references to the Gelfand-Mazur theorem; the other is
S. Mazur, Sur les anneaux lineaires, C. R. Acad. Sci. Paris, 207 (1938),
1025-27.
Theorem 12 is from B. Bollobas, An extension to a theorem of Bishop
and Phelps, Bull. London Math. Soc, 2 (1970), 181-2, and Theorem 13
is from J. P. Williams, Spectra of products and numerical ranges, J.
Math. Anal, and Appl., 17 (1967), 214-20. A good account of numeri-
cal ranges can be found in F. F. Bonsall and J. Duncan, Numerical
Ranges II, London Mathematical Society Lecture Note Series, vol. 10,
Cambridge University Press, 1973, vii + 179 pp.
Johnson's theorem is from B. E. Johnson, The uniqueness of the
(complete) norm topology, Bull. Amer. Math. Soc, 73 (1967), 537-9;
its simple proof is from T. J. Ransford, A short proof of Johnson's
uniqueness-of-norm theorem, Bull. London Math. Soc, 21 (1989),
487-8.
Chapter 12: The algebra of bounded linear operators 185

The commutative Gelfand-Naimark theorem is taken from I. M. Gel-


fand and M. A. Naimark, On the embedding ofnormed rings into the ring of
operators in Hilbert space, Mat. Sbornik, 12 (1943), 197-213; another classical
reference is M. A. Naimark, Normed Rings, Revised English edition; trans-
lated from the Russian by Leo F. Boron, Groningen: Noordhoff, 1964.
13. COMPACT OPERATORS ON BANACH SPACES

For the sake of simplicity we shall assume that all the spaces appearing
in this chapter are complex Banach spaces. Our aim is to study a class
of operators closely resembling the operators on finite-dimensional
spaces; we shall show that these operators are somewhat similar to the
nxn complex matrices.
Vaguely speaking, the operators we shall look at are 'small' in the
sense that they map the unit ball into a 'small' set. To be precise, an
operator T E ^(X, Y) is compact if TBX, the image of the unit ball Bx
under T, is a relatively compact (i.e. totally bounded) subset of Y.
Thus T is compact if and only if TBX is compact. Equivalently, T is
compact if and only if for every bounded sequence (xn) C X the
sequence (Txn) has a convergent subsequence.
We shall write 2S0(^C Y) for the set of compact operators from X into
y. Analogously to 9(Z), we write %(X) for %(X,X).
Examples 1. (i) Every finite rank operator T E 26(X, Y) is compact, i.e.
if dimlm T = dim TX < oo then T E %(X, Y). Indeed, set Z = Im T.
Since Z is finite-dimensional, Bz is compact and so TBX is a subset of
the compact set ||r||fi z .
We shall denote by 2SQO(X Y) the set of (bounded) finite rank opera-
tors from X to 7.
(ii) Every bounded linear functional / E Z* is a compact operator
from X to C.
(iii) Let / be the closed unit interval [0,1] and let X be the Banach
space C(/) of continuous functions with the supremum norm. Let
K(x,y) E C(/x/), i.e. let A' be a continuous function on the closed unit
square Ixl. F o r / E C(I) define a function Tf E C(I) by

186
Chapter 13: Compact operators on Banach spaces 187

(Tf)(x) =
•fo K(x,y)f(y) dy.
J

Then T E 2S(Z) and it is easily seen that, in fact, T E <&0(X). Indeed,


by the Arzela-Ascoli theorem (Theorem 6.4) we have only to check
that TBX is uniformly bounded and equicontinuous. If |^(jc,y)| ^ N
for every (x,y) E Ixl then

\(Tf)(x)\ ^N t \f(y)\ dy ^ N\\f\\,


Jo
and so TBX is uniformly bounded by N. Also, for e > 0, there is a
8 > 0 such that if |jti-* 2 | < 8 then \K(xl9y)-K(x2,y)\ < e. There-
fore if/ E Bx and l^—JC2| < 8 then

f
Jo

f
J
o
-1
\K(xl9y)-K(x29y)\dy<e.

Thus TBX is indeed equicontinuous and so T is compact, as claimed.


(iv) Let H be a Hilbert space with orthonormal bases (e,-)" and
For every operator 71 E 26(/f) we have

£ II^II2= £ £ K^,^)I 2 = £ £ l<e,-,re/.)|2= £


1= 1 1= 1 ;= 1 i= l y= l ;= 1

and so
2
£ \\Tff=
/= 1
tf= £ ££|<r/;.,ey>| = £ I I ^ H ^ £
/= 1 7= 1 7= 1 i= l

This shows that

is independent of the orthonormal basis (e,-)".


An operator 71 E 2S(//) is said to be a Hilbert-Schmidt operator if its
Hilbert-Schmidt norm, ||r|| HS , is finite. One often writes 2S2(#) for the
set of Hilbert-Schmidt operators on H, and ||r|| 2 for ||r|| HS .
As the notation ||r|| 2 indicates, the Hilbert-Schmidt norm is the l2-
norm of the sequence formed by the entries of the matrix representation
of T. Indeed, set atj = {Tej9et) so that T is given by the matrix
A = {atj) in the sense that
188 Chapter 13: Compact operators on Banach spaces

l'") - £ (I, <">*)"•


Then
M2
1 \aij\
V =l /=1
Putting it another way, with

7= 1

we have at E H for every /,


/ oo \l/2
2
imiHs= 2 Ikll and r* = 2 <*,«/>*,-.
V=i / *=i
In particular,

l|7*||2 = £ |(x,a,.)| 2 ^£ W2||a,||2 = IW

It is easily seen that every Hilbert-Schmidt operator is compact.


Indeed, with the notation as above, put

= \x= £ X& E H: \Xi\ ^ k | | , i = 1,2,... .

Since 2°°=1 Ikill2 < °°? the set A is a compact subset of H (see Exercise
1). Since TBH is a subset of A, it is relatively compact. •

The class of compact operators is an example of a closed operator


ideal. An operator ideal 9ft is a function that assigns to every pair X, Y
of Banach spaces a subset <3l(X, Y) of <&(X, Y) such that if T E <3i(X, Y),
5 E » ( y , Z ) and i? E 9(W,-Y) then ST E »(Jir,Z) and Ti? E 2ft(W, 7).
Theorem 2.
(a) <&0(X, Y) is a closed subspace of a (AT, Y).
(b) If I G S80(^, Y), 5 E »(Y,Z) and 7? E 26(W,Z) then
5 7 E 28O(*, Z) and TO E » 0 (W, Y).
Proof, (a) Let us show first that ^0(X, Y) is a subspace of 2S(^, Y), i.e.
if 5, T E 2S0Ur, Y) and A,/A E C then A5 + ^ r E 2S0(^, Y). Let (xn) be
Chapter 13: Compact operators on Banach spaces 189

a bounded sequence in X. As S is compact, (xn) has a subsequence, say


(xnj, such that (Sxnk) is convergent. The operator T is also compact,
and so (xnk) has a subsequence, say (xmk), such that (Txmk) is conver-
gent. But then ((AS + /jiT)xmk) is also a convergent sequence.
Now let us show that %(X, Y) is a closed subset of &(X,Y). Sup-
pose Tn E %(X, Y) and Tn-* TG SB(AT, Y). We have to show that
7 ^ is totally bounded. Given e > 0 let n be such that | | r n - r | | < e.
As Tn is compact, there are x1,...,xmEBx such that { 7 ^ : 1 = ^
i ^ m} is an e-net in TnBx. Thus if JC E 5 ^ then there exists an xt such
that ||7^JC— T^JC/H < 6. Then

showing that {7*, : 1 ^ i ^ m} is a 3t-net in TBX.


(b) Note that a bounded linear operator maps a bounded sequence
into a bounded sequence and a convergent sequence into a convergent
sequence. •

Since finite rank operators are compact, by Theorem 2 (a) every limit
of finite rank operators is compact. The problem of whether every com-
pact operator can be obtained in this way was, for many years, one of
the best-known problems in functional analysis. After about 40 years
the approximation problem was solved in the negative by Per Enflo in
1973, who constructed a separable reflexive Banach space X for which
is not the closure of S&oot^O- Since %(X) is the closure of
whenever X has a (Schauder) basis (see Exercise 7), Enflo's
example also showed that not every separable reflexive Banach space
has a basis, solving another long-standing question. As so often in
mathematics, the counterexample turned out to be the start of the story:
it opened up whole new fields of research on approximation and basis
problems. But we cannot go into that in this book.
The operator ideal of compact operators is closed under taking
adjoints as well.
Theorem 3. An operator T E 2ft(Jf, Y) is compact if and only if
r E®(Y*,X*) is compact.
Proof, (i) Suppose first that T E a (A", Y) is compact, i.e. the set
K = TBX is compact. For a functional / E Y* let Rf be the restriction
of / to K. Clearly Rf E C(K) and, in fact, the map R : Y* -> C(K) is a
bounded linear map, where, as usual, C(K) is taken with the supremum
norm. Let <P = RBY*. Note that for/ E Y* we have
190 Chapter 13: Compact operators on Banach spaces

\\T*f\\ =sup{(x,rf):xeBx}
= sup{(Tx,f):xGBx}
= sup{(y,f):y<E TBX}
= suV{(y,f):y<EK}=\\Rf\\.
This shows that T*BY* is isometric to <P C C(K), with the isometry
given by T*f »-> Rf.
Consequently T*BY* is totally bounded if and only if <P is. By the
Arzela-Ascoli theorem, <P is totally bounded if and only if it is uniformly
bounded and equicontinuous. Both conditions are easily checked.
If fEBY* then ||r*/|| ^ \\T*\\ = ||r||, and so <P is uniformly
bounded. Also, if / E BY* and y,y' E K then
\(Rf)(y)-(Rf)(y')\ = \f(y-y')\ « | b - / | | ,
and so <P is equicontinuous.
(ii) Now suppose that T* E 28(Y*,X*) is compact. By part (i), the
map T** E <3&(X**,Y**) is compact, i.e. T**Bx** is relatively compact.
But under the natural embeddings X C X** and y C Y** we have
TBX = T**BX C T**Bx**, and so TBX is also relatively compact. •

Theorems 2 and 3 state that the compact operators form a closed


operator ideal that is also closed under taking adjoints. In particular,
for a Banach space X, the set 28oUO is a closed ideal of the Banach
algebra <&(X). If H is a Hilbert space then <&0(H) is a closed ideal of H
which is also closed under taking adjoints: it is a closed *-ideal.
The main aim of the chapter is to present the spectral theory of com-
pact operators, due to Frigyes (Friedrich or Frederic) Riesz. Recall that
the spectrum of an operator T E 28 (X) is
a(T) = {A E C: T-kl does not have a bounded inverse},
where / is the identity operator on X. We proved in the previous
chapter that for every T E ^(X) the spectrum of T is a non-empty
closed subset of C, contained in the disc {z E C: \z\ ^ ||7i|}. As we
shall see, for a compact operator T E 95(X), the spectrum of T resem-
bles the spectrum of an operator on a finite-dimensional space, i.e. the
spectrum of an nXn matrix. To be precise, if X is infinite-dimensional
and T E 2&0U0 then a(T) is a countable set whose only accumulation
point is 0 and if A E a(T) (A ^ 0) then A is an eigenvalue of T with
finitely many linearly independent eigenvectors.
Chapter 13: Compact operators on Banach spaces 191

Theorem 4. Let T E %(X) and a > 0. Then T has only finitely many
linearly independent eigenvectors with eigenvalues having modulus at
least a.
Proof. Suppose xx, x2,... is an infinite sequence of linearly independent
eigenvectors such that Txt = A,-*/ j= 0 and |Af-1 ^ a for every i. Set
Xn = lin{xi,...,xn}. By Theorem 4.8 (b), there exists a sequence
{yJt C Xsuch t h a t ^ G Xn and rf^,^^) = |bJ| = 1.
Put zn = yn/Xn and note that ||zj| ^ I/a, Tzn E Xn and
Tzn-yn E Xn_1. Indeed, the first two assertions are obvious and if
yn = 2 * = 1 c*** then

^ 1 I

Hence if n > m then

Consequently the bounded sequence (z^)^ does not contain a subse-


quence (2^)1° such that Tznk is convergent, contradicting the compact-
ness of T. D

Our next aim is to show that if T is a compact operator and A ^ 0 is


not an eigenvalue of T then A £ cr(r), i.e. XI- T is invertible. Equiva-
lently, cr(r)U{0} = o-p(r)U{0}, i.e. with the possible exception of 0,
every point of the spectrum is in the point spectrum. In proving this we
may and shall assume that A = 1; so our aim is to prove that S = I—T
is invertible. We need two lemmas, both of which are proved in a more
general form than necessary.
Lemma 5. Let T E %(X) and set S = I- T. Then SX is a closed sub-
space of X.
Proof. Set N = KerS; by Theorem 4 we know that N is finite-
dimensional, say with basis {bl9...,bn}. Then there is a closed subspace
MCX such that X is the direct sum of M and N: X = M@N. Indeed,
if we choose / i , . . . , / w E X* such that ft{bj) = 8(j then we may take
M = n n = 1 Ker/J. The projections of x E X into TV and M are
PNM = E" =1 fi(x)bi andpM(x) = x-pN(s).
Let 5 0 be the restriction of 5 to M: So = S\M. Then SX =
SM = S0M and KerS 0 = KerSHM = {0}, and so So is injective. Hence
192 Chapter 13: Compact operators on Banach spaces

to prove that SX = S0M is closed, it suffices (and, indeed is necessary)


to prove that 5 0 is bounded below.
Suppose that So is not bounded below, i.e. Soxn -> 0 for some
sequence (xn)i C M (\\xn\\ = 1). Since T is compact, (Txn)i has a con-
vergent subsequence, and so we may assume that (Txn)™ itself is conver-
gent, say Txn —> y. Then
xn = (So+ T)xn = Soxn + Txn -> y,
and so ||y|| = 1. But we also have Soxn —> Sy, and so Soy = 0, contrad-
icting KerS 0 = {0}. •

Lemma 6. Let S,T E 28(AT) be such that 5+ T = / and SX C Y, where


Y is a closed proper subspace of X. Then for every e > 0 there is a
point Jt0 E B^ such that d(7jt0, TY) > 1 - e . •

Proof. By Theorem 4.8(a), there exists an JC0 E B^ such that


d(x0, Y) > 1-e. As 7*o = *o-Sx o , 5jc° G Y a n d 7 Y = ( 7 " 5 ) y c y>
we have
, 7Y) ^ rf(jco-5*o, 7) = dUo, Y) > 1-6. •

Theorem 7. Let T be a compact operator and suppose A ^ 0 is not an


eigenvalue of T. Then A £a(T).
Proof. By replacing T by 7/A, it suffices to prove the result for A = 1.
Let then T E 9&0(AT), 5 = / - T and Ker5 = (0). We have to show that
5 is invertible.
Let us prove that SX = X. Set Yn = SnX (n = 0,1,...), so that
y0 = X D YiD • • • . By Lemma 5 the subspaces Yn are closed. Let us
show first that Yn = Yn+1 for some n. Indeed, otherwise Y0D Yx D
and all the inclusions are strict. Then, by Lemma 6, one can find ele-
ments yn E BYn such that d(Tyn, TYn+1) > \. But then, in particular,
||ry r t -7> m || > \ if n ± m, and so (Tyn) has no convergent subse-
quence, contradicting the compactness of T.
We claim that, in fact, Yo = Yx. Suppose that this is not so. Then
there is an m such that Ym^ ± Ym = Y m+1 . Let u E Ym.1\Ym. Then,
as Su E Ym = y m + 1 = 5Ym, there exists a point v E Ym such that
Su = Sv. But then S(u-£;) = 0 and so 0 ^ u — v E KerS, contradicting
our assumption. Consequently Yt = Yo, i.e. SX = X, as claimed.
The proof is essentially complete. The bounded map 5 : Z—> X is a
1-1 map of the Banach space X onto itself and so, by the inverse-
mapping theorem, S is invertible. •
Chapter 13: Compact operators on Banach spaces 193

Let us restate the information contained in Theorems 4 and 7 about


the spectrum of a compact operator as a single result.
Theorem 8. Let T be a compact operator on an infinite-dimensional
Banach space. Then a(T) = {0, Ax, A 2 ,...}, where the sequence
A 1 ,A 2 ,... (of non-zero complex numbers) is either finite or tends to 0;
furthermore, every A, is an eigenvalue of T, with finite-dimensional
eigenspace. •

With some more work, we can get more detailed information about
the structure of compact operators. If T is compact and S = I-T then
ImS is a finite-codimensional closed subspace of X; even more, for a
suitable n ^ 1, X is the direct sum of KerS" and ImS". We prove this,
and a little more, in the following theorem.
Theorem 9. Let J b e a Banach space, T E %(X) and S = I-T. Set
Nk = KerS* and Mk = ImS* (k = 0,1,...), where S° = /. Then (Nk)%
is an increasing nested sequence of finite-dimensional subspaces and
(Mfc)o is a decreasing nested sequence of finite-codimensional subspaces.
There is a smallest n ^ 0 such that Nn = Nm for all m ^ n. Further-
more, Mn = Mm for all m ^ n, X is the direct sum of Mn and Nn, and
S\Mn is an automorphism of Mn.
Proof. By expanding Sk = (I- T)k, we see that Sk = I-Tk, where Tk is
compact. Hence, by Lemma 5, Mk is a closed subspace of X. Clearly
No C Nx C • • • and Mo D Mt D • • •; furthermore, we know that each
Nk is finite-dimensional.
As in the proof of Theorem 7, Lemma 6 implies that there is a smal-
lest n such that Nn = Nn+1, and there is also a smallest m such that
Mm = M m + 1 . Then Nn = Nn. for all n' ^ n and Mm = Mm, for all
m' ^ m.
Let us turn to the main assertions of the theorem. We prove first that
NnDMn = {0}. Let y E NnHMn. As y E Afw, we have y = S"x for
some J C £ X But as y £ J V n , Sny = 0 and so S2"JC = 0. Hence
x E N2n = Nn, implying Snx = 0. Thus y = Snx = 0, showing that
^ H M , = {0}.
We claim that for p = max{rc,m} we have X = Nn®Mp. Indeed,
given x E X, we have Sp;t E M p . But SpMp = Mp and so there is a
vector y E Mp such that Spy = Spx. Hence x-y E Np = Nn and so
x = y + (x-y) shows that A" = Nn + Mp. Could we have p > nl
Clearly not, since then Mn would strictly contain Mp and so we would
194 Chapter 13: Compact operators on Banach spaces

have NnDMn ^ {0}. Thus p — n and X is the direct sum of Nn and


Mn-, as Nn is finite-dimensional (see Exercise 4.20).
Finally, SMn = Mn+1 = Mn and

Ker(S\Mn) = KerSDMn = NxnMn C NnHMn = {0}.


Hence, by the inverse-mapping theorem, the restriction of S to Mn is
invertible. •

Putting Theorems 8 and 9 together, we arrive at the crowning


achievement of this chapter: a rather precise description of the action of
a compact operator.
Theorem 10. Let X be an infinite-dimensional Banach space and let T
be a compact linear operator on X. Then a(T) = {0,A 1 ,A 2 ,...}, where
the sequence A 1 ,A 2 ,... is either finite or tends to 0. For every A = A,
there is an integer kk ^ 1 and closed subspaces Nk = N(\; T) and
Mk = M(X; T) invariant under T, such that 7VA is finite-dimensional and

The restriction of A / - T to MA is an automorphism of Mk,

Nk = Ker(A/-r)*A + Ker(A/-r)* A ~\
and for /x = X}• ^ A = Af- we have Nk C M^.
Proof. Only the last claim needs justification. The operator /JLI—T maps
Mk into itself and Nk into itself. Furthermore, (ixI-T)\Nk is an auto-
morphism of the finite-dimensional space Nk since if we had
(/JLI—T)X = 0 for some x E Nk (x ± 0) then we would have
(Kl-T)nx = {X-n)nx ^ 0 for every n ^ 1, contradicting {\I-T)k*x =
0. Consequently, ( ^ / - T)nNk = Nk for every /i ^ 1, and so 7VA C MM.
D

There is no doubt that Theorem 10 is a very beautiful theorem. At


first sight it is not only beautiful but very impressive as well: it seems to
come close to giving us a very fine decomposition of the space into a
direct sum of generalized eigenspaces. Unfortunately, this is rather a
mirage: the theorem cannot even guarantee that our compact operator
has a non-trivial closed invariant subspace, let alone give a direct-sum
decomposition. In fact, non-trivial closed invariant subspaces do exist,
as we shall prove in Chapter 16. However, to prove the existence of
invariant subspaces we shall need some results to be proved in Chapter
15. Before we turn to that, in the next chapter we shall show that a
Chapter 13: Compact operators on Banach spaces 195

best possible decomposition can be guaranteed if we deal with a com-


pact normal operator on a Hilbert space.

Exercises

1. Let 1 ^ p < oo, a = (ai)i E lp and


A = {x = (*,•)? E /p : tal ^ k | (i = 1 , 2 , . . . ) } .
Prove that A is a compact subset of lp.
2. Let # be a closed and bounded subset of lp (1 ^ p < oo). Prove
that A' is compact if and only if for every e > 0 there is an n such
that 2°°=n k-\ P < * f o r e v e r Y x = (Xi)i E AT.
3. As in Example l(xiii) of Chapter 2, let C{k)(0,1) be the vector
space of k times continuously differentiable functions/: (0,1) —> IR
such that

I |/ ( / ) «|<".
II/IU = sup /=o
Show that Xk = (C{k)(0,1), || • ||*) is a Banach space and the formal
identity map i: Xk-^> Xk-i (/ •-* / ) is a compact operator.
4. Let T e a o ( ^ y ) ? where X is infinite-dimensional. Show, that the
closure of TS(X) = {Tx : x E X, \\x\\ = 1} in Y contains 0. (HINT:
Consider a sequence (*„) C S(X) such that | k , - x m | | ^ 1 for
n ? m.)
5. Let (en)i be an orthonormal basis of a Hilbert space H and let
T E 280Ctf> y), where Y is a normed space. Show that Ten -* 0.
6. Let Jf be a normed space such that for every finite set A C X and
every e > 0, X has a decomposition
JT= M@N
as a direct sum of two closed subspaces, such that M is finite-
dimensional,
d(a,M) <e
for every a E A, and
\\pN(x)\\ ^cd(x,M)
for some c > 0 and every J C E Z , where pN is the canonical projec-
tion onto N. Show that 96OW i s the norm closure of ^QQ(X), i.e.
196 Chapter 13: Compact operators on Banach spaces

every compact operator on X is the operator-norm limit of finite


rank operators.
7. Let A" be a Banach space with a Schauder basis (xn)i. Show that
%(X) is the closure of 98oo(AT).
8. Let (An)5° be a sequence of non-zero complex numbers tending to
0. Show that a(T) = {0,A1,A2,...} for some complex operator T
on some compact Banach space X. Show that if all A, are real
then X can be chosen to be a real Banach space.
9. Let X and Y be Banach spaces, let T E <&0(X, Y), and let
/ E <&(X, Y) be invertible. Show that I m ( / - T) is closed in Y and
has finite codimension.
10. Let X be a complex Banach space, and let T E ^(X) be such that
Tn is compact for some n ^ 1. Show that a(T) = {0,A 1 ,A 2 ,...},
where the sequence A 1 ,A 2 ,... is finite or tends to 0, and every A, is
an eigenvalue of T. What is the relationship between the sub-
spaces N(\; T) and N(fi\Tn)l
11. Let XlyX2,... be Banach spaces and let X = r H Xn be the
direct sum of these spaces, with x = (xn)i having norm

INII= Jjk.il.
Let Tn E <&(Xn) (n = 1,2,...) and set T = © n = 1 Tn, that is,
define T E <&(X) by Tx = (Txn)™. Show that" T E ^0(X) iff
Tn E®0(Xn) for every n.
12. Let I b e a Banach space, (Tn)i C 28OW> and let Tn -+ T in the
operator-norm topology. Show that U°°=1 ^«(2?) is relatively com-
pact, where B = B(X) is the unit ball of X. Show also that if
kn E cr(Tn) and \n -» A then A E er(r).
13. Let E be the space C[0,1], endowed with the Euclidean norm
Mi
7
II/II2 = "
and let T be as in Example l(iii). Show that T E %(E), i.e. T
maps the unit ball of the Euclidean space E into a relatively com-
pact set.
14. Let H be a Hilbert space and T E <&(H). Show that T E %(H) if
and only if Txn —> 0 whenever *„ converges weakly to 0, i.e.
whenever (xn, x) —> 0 for every x E H.
15. Construct a compact operator T on lp (1 ^ /? ^ 00) such that
= {0} and 0 is not an eigenvalue of T.
Chapter 13: Compact operators on Banach spaces 197

16. Let cx, c2,... be non-negative reals and


C = {x G l2 : x = (xk)i, \xk | ^ ck for every k}.
Show that if C is a compact subset of l2 then c^ —> 0. For what
sequences (ck)i is C compact?
17 + . Let AT be a compact subset of a normed space. Show that K is
contained in the closed absolutely convex hull of a sequence tend-
ing to 0: there is a sequence xn —> 0 such that K C C, where

C=co{jt l l :/! = 1,2,...} = ) £ A,-** : £ |A,.| ^ 1,/i = 1,2,...

Notes

Compact operators were first introduced and applied by Hilbert,


Grundziige einer allgemeinen Theorie der linearen Integralgleichungen,
Leipzig, 1912, and F. Riesz, Les systemes d'equations a une infinite
d'inconnus, Paris, 1913, and Uber lineare Funktionalgleichungen, Acta
Math., 41 (1918), 71-98. In presenting the Riesz theory of compact
operators we relied on Ch. XI of J. Dieudonne, Foundations of Modern
Analysis, Academic Press, New York and London, 1960, xiv + 361 pp.
Theorem 3 is due to J. Schauder, Uber lineare vollstetige funktional
Operationen, Studia Math., 2 (1930), 185-96.
The first solution of the approximation problem was published by P.
Enflo, A counterexample to the approximation problem in Banach
spaces, Acta Math., 30 (1973), 309-17; a simplified version of the solu-
tion is in A. M. Davie, The approximation problem for Banach spaces,
Bull. London Math. Soc, 5 (1973), 261-6.
14. COMPACT NORMAL OPERATORS

In the previous chapter we saw that for every compact operator T on a


Banach space X, the space can almost be written as a direct sum of gen-
eralized eigenspaces of T. If we assume that X is not merely a Banach
space, but a Hilbert space, and T is not only compact but compact and
normal, then such a decomposition is indeed possible - in fact, there is a
decomposition with even better properties. Such a decomposition will
be provided by the spectral theorem for compact normal operators: a
complete and very simple description of compact normal operators.
Thus with the study of a compact normal operator on a Hilbert space
we arrive in the promised land: everything fits, everything works out
beautifully, there are no blemishes. This is the best of all possible
worlds.
We shall give two proofs of the spectral theorem, claiming the
existence of the desired decomposition. In the first proof we shall make
use of some substantial results from previous chapters, including one of
the important results concerning the spectrum of a compact operator.
The second proof is self-contained: we shall replace the results of the
earlier chapters by easier direct arguments concerning Hilbert spaces
and normal operators.
To start with, we collect a number of basic facts concerning normal
operators in the following lemma. Most of these facts have already
been proved, but for the sake of convenience we prove them again.
Lemma 1. Let T E 28(//) be a normal operator. Then the following
assertions hold.
(a) || 7*|| = ||7**|| for every x E H.
(b) K e r 7 = KerT*.
(c) H r l = ||7|| n for every n^ 1.
198
Chapter 14: Compact normal operators 199

(d) r{T) =
(e) If A + IJL then Ker(XI-T) IKer^I-T).
(f) For every A E C, both Ker(A/-7) and (Ker(A/-r)) x are invari-
ant under both 7and T*.
(g) If H is the orthogonal direct sum of the closed subspaces Ho and
H1 invariant under T then with To = T\H0 and Tt = T\H1 we
have
II71I =
and 7} is a normal operator on //,-, with 7f = r * | / / , • ( / = 1,2).
Proof, (a) As T*T = 7T*, we have
||7i|| 2 = (7JC,7JC) = (T*Tx,x) = (TT*x,x) = (T*x,T*x) = \\T*x\\2.
(b) By part (a), we have Tx = 0 iff 7*JC = 0.
(c) If S E S8(//) is hermitian then
||5JC||2 = (5JC,5X) = (5*5x,x) = (S2x,x) ^ \\S2\\\\x\\2.
From this it follows that ||5 2 || = ||5|| 2 , and by induction on m we get
||52-|| = u^ip-^ This implies that ||5«|| = ||5||n f()r eyery n^X
is hermitian,
Uril 2 = H W y l l = \\(TT*)n\\ = \\TT*\\n = \\T\\2n.
(d) By (c) and the spectral-radius formula (Theorem 12.9),
r(T) = lhn\\T"\\Vn = \im\\T\\ = \\T\\.
(In fact, (c) and (d) are equivalent: if S E ^h{X) for a complex Banach
space X then r(S) = \\S\\ iff ||S"|| = ||5f for every n & 1.)
(e) If Tx = AJC and Ty = fxy then T*y = jly because by (b) we have
y E Ker(M/-r) = Ker(/Z/-r*). Therefore
\(x,y) = (Tx,y) = (x,T*y) = (x,jiy) = p(x,y),
and so if A ^ /JL then (x,y) = 0.
(f) As XI- T commutes with T and 7*, Ker(A/- T) is invariant under
both 7and T*.
Also, let (x,y) = 0 for all y E Ker(A/- 71). Then, since Ker(A/- T) is
invariant under T*, for y E Ker(A/- 71) we have

,j) = (x,T*y) =0.


Hence 7JC E (Ker(A/- T))1. Similarly,
200 Chapter 14: Compact normal operators

(T*x,y) = (x,Ty) = 0
for every y E Ker(A7- T) and so T*x = (Ker(A7- T))1.
(h) Let x = ho + hx, with ht E Ht (i = 0,1). Then
IMI2 = M 2 + I N I 2 , r* = 77io+ Th1 = ro/
and

« max{||ro||2, ||7k
= max{||r o || 2 ,||r 1 || 2 }W 2 .
Thus ||r|| ^ maxlll^H, ||r 2 ||}. The reverse inequality is obvious.
Finally, as Ho and Ht are invariant under T, it follows that H1 = H^
and Ho = H1 are invariant under T*. D

It is worth emphasizing that Lemma 1 is a collection of elementary


and simple facts, except for part (d), which is based on the spectral-
radius formula.
Let us see then the first incarnation of the spectral theorem, claiming
the existence of a spectral decomposition for a compact normal operator.
Theorem 2. Let T E 96(if) be a compact normal operator. For an
eigenvalue A of T, let Hx = Ker(r-AZ) be the eigenspace of T belong-
ing to A, and denote by Px the orthogonal projection onto Hk. The
operator T has countably many non-zero eigenvalues, say A1?A2,
Furthermore, dim//Afc < °° for every k, the projections PA are orthogo-
nal, i.e. PkkPkl = 0 if k + /, and

where the series is convergent in the norm of 2B(iJ).


Proof. By Theorem 13.8 and Lemma 1 (e), we have to prove only (1).
Given e > 0, choose n ^ 1 such that |A^ | < e for k > n. Set

Then Ho and / ^ are invariant under S and 71. With 7} = 7^)//,- and
5, = S|//; (i = 1,2), we have To = So and Sx = 0. Therefore, by
Chapter 14: Compact normal operators 201

Lemma l(g),
\\T-S\\=m&x{lT0-S0\\,\\Tl-Sl\\}=\\n\.
But 7i is a compact normal operator and so, by Theorem 13.8, ||7\||
is precisely the maximum modulus of an eigenvalue of 7\. As every
eigenvalue of Tt is an eigenvalue of T, by our choice of n we have
||7i|| < e. Hence (1) does hold. •

Let us state two other versions of the spectral theorem.


Theorem 3. Let T be a compact hermitian operator on an infinite-
dimensional Hilbert space H. Then one can find a closed subspace HQ
of H, a (finite or countably infinite) orthonormal basis (xn) of Ho, and a
sequence of complex numbers vn -> 0, such that if x = 2 cnxn+x> >
where x' E HQ~ , then
Tx = 2 vncnxn.
n
Proof. Let A l9 A 2 ,... and // A i ,// A 2 ,... be as in Theorem 2. Take a
(necessarily finite) orthonormal basis in each Hkk and let (xn) be the
union of these bases. Let Ho be the closed linear span of the orthonor-
mal sequence (xn), and set vn = A*, if xn E HXk. D

Corollary 4. Let T be a compact normal operator on a Hilbert space //.


Then H has an orthonormal basis consisting of eigenvalues of T. D

In fact, compact normal operators are characterized by Theorem 2 (or


Theorem 3). Let {xy : y E f} be an orthonormal basis of a Hilbert
space H, and let T E &(#) be such that Txy = vyxy. Then T is com-
pact iff
|{y: \vy\ ^ 6 , yer\\ < °° (2)
for every e > 0 (see Exercise 2).
Our proof of Theorem 2 was based on two substantial results:
Theorem 13.8 concerning compact operators on Banach spaces, and the
spectral-radius formula. We shall show now how one can prove
Theorem 2 without relying on these results. It is a little more con-
venient to prove Theorem 2 for compact hermitian operators; it is then a
simple matter to extend it to normal operators.
Recall that the numerical range V{T) of a Hilbert space operator
T E »(H) is
202 Chapter 14: Compact normal operators

{(Tx,x):xES(H)}
and the numerical radius v(T) is
v{T) =sup{|A|:AE V(T)}.
If T E 26(//) is hermitian, i.e. T* = T, then its numerical range is real
since
(Tx,x) = (x,T*x) = (x,Tx) =
for every x E H and so (7JC,JC) is real. In fact, T E 9S(//) is hermitian
iff its numerical range is real. Also, the spectrum of a hermitian opera-
tor is real. We shall not make use of any of the results proved about
numerical ranges; the next lemma is proved from first principluses.
Lemma 5. Let Tbe a hermitian operator. Then \\T\\ = u(T).
Proof. Set v = v(T), so that |(7JC,JC)| ^ i;||jc||2 for every x E H. We
have to show that || T\\ ^ v.
Given x E S(H), let y E S{H) be such that Tx = \\Tx\\y. Then
(Tx,y) = (x,Ty) = ||7;t||andso
llTrll = (Tx v) = \

= MW2+IMI2} = y.
Hence ||7JC|| ^ v for every x E 5(//) and so ||r|| =^ v, as claimed. D

Theorem 6. Let U be a compact hermitian operator on H. Then [/ has


an eigenvalue of absolute value || U\\.
Proof. Set
a = inf (Ux,x) and fc = sup (C/JC,JC)

so that V(U) = [a,b]. By Lemma 5, \\U\\ = max{-a,fe}. Replacing U


by -£/, if necessary, we may assume that \\U\\ = b > 0. We have to
show that b is an eigenvalue of U.
By the definition of b, there is a sequence (*„) C S(H) such that
(Uxn,xn) -> 6. Since £/ is a compact operator, by replacing (xn) by a
subsequence, we may suppose that (Uxn) is convergent, say Uxn —> y 0 .
Then ||yo|| ^ 6 because (Uxn,xn) -* b and ||xn|| = 1. As
Chapter 14: Compact normal operators 203

\\Uxn-bxn\\2 = \\Uxnf-7b{Uxn,xn) + b2\\xnf


^\\U\\2-2b(Uxn,xn)+b2
= 2b2-2b(Uxn,xn).sp-.O5
and
(Uxn,xn)-+b,
we have
(U-b)xn^0.
Therefore

Put x0 = yo/b. Then, on the one hand, Uxn —> y0 = bx0 and, on the
other hand, Uxn —» Ux0. Consequently we have Ux0 = bxQ. Since
||JCO|| ^ 1 (in fact, ||JCO|| = 1), b is indeed an eigenvalue of U. •

Let us now see how Theorem 6 may be used to deduce Theorem 2 for
compact hermitian operators. For the sake of variety, we restate
Theorem 2 in the following form.
Theorem 7. Let H be a Hilbert space and let U E 2fc(//) be a compact
hermitian operator. Then there is a (possibly finite) sequence (A^) of
real numbers and a sequence (Hk) of linear subspaces of H such that
(a) A* -> 0;
(b) dimHk < oo;
(c) HkLHt ioxk± /,
(d) if x = 2L xk+x, where JC^ E //^ and i E Hk for every /:, then
Ux =

Proof. Let \ y , (y E T) be the non-zero eigenvalues of U and let Hy be


the eigenspace belonging to A r : Hy = Ker(i/-A y /). We know that
Hy±H8i£yf8.
Let us show first that dimHy < « and, for every e > 0, there are only
finitely many Ay with |Ay| ^ €. Suppose not. Then, by taking an
orthonormal basis in each Hy with |Ay| ^ e, we find that there is an
infinite orthonormal sequence (xn)™ such that Uxn = fxnxn, where
\/jin\ ^ 6. But then (Uxn)™ does not contain-a convergent subsequence,
contradicting the compactness of U.
204 Chapter 14: Compact normal operators

This implies that the non-zero eigenvalues may be arranged in a


sequence (A^) such that with Hk = Ker(U—\kI) the conditions (a)-(c)
are satisfied.
Then, as each Hk is invariant under U, so is the closed linear span M
of all the Hk and, consequently, so is M1. Denote by U the restriction
of U to M1. Then U E 2S(M-1) is also a compact hermitian operator.
As a non-zero eigenvalue of U is also a non-zero eigenvalue of [/, it
follows from the definition of M and from Theorem 6, that U = 0.
If the sequence (A^) of non-zero eigenvalues is finite then we are
done. Otherwise, let
00

x = 2 xk + x, where xk E Hk and x E M-
k=i

Put

= 2 ** + * and y{n) = £ kkxk.


k=l k=l
Then f/x^ = y{n) and x (n) -> x. As
00

y(«) _y= 2
A: = l
the continuity of U implies that Ux = y, proving (d). D

Before we recover from Theorem 7 the full force of Theorem 2, let us


show that compact hermitian operators are rather like real numbers.
An operator T E 2S(//) is said to be positive if it is hermitian and
V{T) C [0,oo), i.e. (Tx,x) ^ 0 for every x E H. Note that if T is any
(bounded linear) operator on a Hilbert space then T*T and TT* are
positive (hermitian) operators:
(T*Tx,x) = (Tx,Tx) = \\Tx\\2 and (TT*x,x) = \\T*x\\2.
Theorem 8. A compact positive operator U on a Hilbert space has a
unique positive square root V. Every hermitian square root of U is
compact.
Proof. Let A 1 ,A 2 ,... be the non-zero eigenvalues of U, let Hk be the
eigenspace belonging to A^ and let M be the closed linear span of the
Hk. Then M 1 = Ker U and A^ > 0 for every k. Define V E 26(i/) by
Vx = \fk~kx if x E Hk and Vx = 0 if x E M1. Then V is a positive
square root of U.
Chapter 14: Compact normal operators 205

Now let W be a hermitian square root of U. Note that W commutes


with U and so it commutes with U-kkh UW = W2W = WW2 = WU.
Therefore Hk = Ker(U-\kI) is invariant under W, and so is ML =
Kerf/. Then W \M = 0 and so W\ML = 0.
2 L

Also, let (x1,...,xn) be an orthonormal basis of Hk consisting of


eigenvectors of W: Wxt = /n,-*,-, say. As W2 = U, we have JJL2 = \k and
so fa = ±\fk~k-+ 0. Since V^k~* 0, by relation (2) the operator W is
compact. Furthermore, if W is positive then /*,- = V^it an<^ s o W ^s Pre~
cisely V. D

For T E 2So(^) the unique positive square root (T*T)X^2 of the com-
pact positive operator T*T, guaranteed by Theorem 8, is called the
modulus or absolute value of T, and is denoted by \T\. Note that
|||r|jc|| = ||7JC|| for every* E H, because
|||r|*|| 2 = <|r|jc,|r|x> = ((T*T)1/2X,(T*T)1/2X)

= (T*Tx,x) = (Tx,Tx) = \\Tx\\2.


In fact, \T\ is the unique positive operator with this property (see Exer-
cise 4).
Let us see then how the spectral theorem for compact normal opera-
tors can be recovered from Theorems 7 and 8. To be precise, we shall
deduce the version given in Corollary 4 from these theorems.
Given a compact normal operator T, how can we get an orthonormal
basis consisting of eigenvectors of 77
Let Ht, H2,... be the eigenspaces belonging to the non-zero eigen-
vectors of the compact positive operator U = TT* = TT*, and let
// 0 = K e r [ / = PI Ht = (lm{// 1 ,// 2 ,...}) ± .

Since TU = UT, each Hk is invariant under T. For * E Ho we have


||7*||2 = (Tx,Tx) = (T*Tx,x) = (Ux,x) = 0
and so T\H0 = 0. Furthermore, for k ^ 1 the restriction of T to Hk is
normal; as Hk is finite-dimensional, we know from linear algebra that
Hk has an orthonormal basis consisting of eigenvectors of T (see Exer-
cise 5). Take the union of these bases, together with an orthonormal
basis of Ho.
It is easily seen that a similar assertion holds for several commuting
compact normal operators.
206 Chapter 14: Compact normal operators

Theorem 9. Let T1,...,Tn be commuting compact normal operators on


a Hilbert space H. Then H has an orthonormal basis consisting of com-
mon eigenvectors of all the Tt.
Proof. For every / i £ C and k = l , . . . , / i , the eigenspace Ker(/Jil—Tk)
is invariant under all the 7}. Hence H is the orthogonal direct sum of
the subspaces

1=1

All these spaces are finite-dimensional, with the possible exception of


HQ,...,O- Taking an orthonormal basis of each H^,...,MII, the union of
these bases will do. •

As our final theorem concerning abstract operators in this chapter, let


us note that our results, say Theorem 2 or Theorem 3, give a complete
characterization of compact normal operators up to unitary equivalence.
Two operators T, T E 26(7/) are said to be unitarily equivalent if for some
unitary operator U we have T = U~lTU = U*TU, i.e. if they have the
same matrix representation with respect to some orthonormal bases.
Let X be the collection of functions n: C\{0} —> {0,1,2,...} whose
support {A E C\{0}: n(A) ^ 1} has no accumulation point (i.e. in C there
is no accumulation point other than 0). In particular, the support is
finite or countably infinite. The following result is easily read out of
Theorem 2 (see Exercise 14).
Theorem 10. Let H be an infinite-dimensional complex Hilbert space
and let ^N{H) be the collection of compact normal operators on H. For
T E %N{H) and A E C\{0} set
nT(\) = dimKer(A/-7).
Then the correspondence T'»-» nT defines a surjection ^>^{H) —> X;
furthermore, T and T are unitarily equivalent iff nT = nr. •

We close this chapter by showing how the spectral theorems we have


just proved enable us to solve a Fredholm integral equation.
Let / = [a,b] for some a < b, and write E for the Euclidean space
C(I) endowed with the inner product

ft) git) &

and norm ||/|| 2 = (/,/) 1 / 2 . Thus the completion of E is L2(0,1).


Chapter 14: Compact normal operators 207

Let K(s,i) G C(/x/), and for/ G E define Uf G E by

Then # is the kernel of the integral operator U. It is easily checked that


U E 2S(£, C(/)) and £/ maps the unit ball of E into a relatively compact
set in C(I). Indeed, this follows from the Arzela-Ascoli theorem, since
if \K(s,t)-K(s',t)\ < e for all t then \(Uf)(s) -(Uf)(s')\ ^ e\\f\\2 by the
Cauchy-Schwarz inequality.
As the formal identity map C(I) —» E, where / • - > / , is continuous, U
extends to a compact operator on L2(0,1); for simplicity, we write U for
this extension as well. In fact,
rb
(Uf)(s) = t K(s,t)f(t)dt
* n

for/GL2(0,l).
From now on we suppose also that K(s,t) = K(t,s); in this case U is
easily seen to be a hermitian operator. We consider a Fredholm integral
equation
h
g(s) = K(s,t)f(t) dt-\f(s) =
Ja

For what values of A can we solve this equation, and what can we say
about the solution? As we shall see, this question is ideally suited for
the theory we have at our disposal.
Let us prove three quick lemmas before giving the answer.
Lemma 11. Let (<pn) be an orthonormal sequence of eigenvectors of U
with non-zero eigenvalues (An) guaranteed by each of Theorems 2, 3
and 7, and Corollary 4, such that if f-L<pn for every n then Uf = 0.
(The sequences ((pn) and (\n) may thus be finite or infinite). Then each
kn is real,
l\\n<pn(t)\2^N
n

for every x, where Af depends only on K(s, t), and


rb
2A^ \ \ \K(s,t)\2dsdt.
n Ja Ja

Proof. For 0 ^ t ^ 1 put kt(s) = K(s,t). Then, by Bessel's inequality,


for every m we have
208 Chapter 14: Compact normal operators

v5! fb 2
fb
2
X kt(s)<pH(s) ds = 2 |A»9»W| « |*,(s)|2<k.

Therefore

\K<Pn(t)\2 dx ^ \ \K(s,t)\2dsdt. D
n= \

Lemma 12. If <pn E L2(0,1) is an eigenvector of U with eigenvalue


Xn ?0then<pn E
Proof. By the Cauchy-Schwarz inequality, Uf E C{I) for / E L2(0,1).
Hence U<pn = \n<pn 6 C(I). D

Lemma 13. Let g E L2(0,1), f = Ug and let (cn) be the Fourier


coefficients of / with respect to (<pn). Then the series 2 CnVni51) c o n "
verges absolutely and uniformly to f(s) in [0,1].
Proof. Let g = 2 an(Pn + g be the decomposition of g in L2(0,1),
where g-L<pn for every n. As £/: L2(0,1) —> C(/) is continuous,
cn = Xnan and Ug = 0, the uniform convergence follows. Furthermore,
for every finite set F of indices,
,2 / \2

nfEF

I l«J 2 )f 2 \K<Pn(s)\2) ^ N\\gg


IF I \neF /
by Lemma 11. •

We are now ready to answer our question about the Fredholm


integral equation.
Theorem 14. If A £ a(U) and g E C(I) then the Fredholm integral
equation
Uf-Af = g
with a hermitian kernel K(s,t) = #(/,$) has a unique solution / given by

where (<?„) is an orthonormal basis of eigenvectors of U, the series is


Chapter 14: Compact normal operators 209

absolutely and uniformly convergent in [0,1] and the (an) are the
Fourier coefficients of g with respect to (<pn), i.e.

-r. rb
g(t)<pn(t) dt.

Proof. The unique solution in L2(0,1) is / = (U—A) lg. Clearly, / has


Fourier coefficients an/{\n—k). By Lemma 13, g + A/ = UfE C(I) and

g~\~\f = / (p

where the series is absolutely and uniformly convergent. •

There are numerous other applications of the elementary spectral


theory of compact hermitian operators to integral equations, notably to
the Sturm-Liouville equation. However, a proper account of these
applications would be much longer than we have space for.
In conclusion, let us point out that the spectral theorem for compact
normal operators is only the beginning of the story. The result we
proved is a very simple version of a spectral theorem for normal (not
necessarily compact) operators. With every normal operator one can
associate a so-called spectral measure containing all the information
about the operator up to unitary equivalence.

Exercises

1. Let Ho be a closed subspace of a Hilbert space H invariant under


a normal operator T E 2S(//). Is Hx = H^ necessarily invariant
under Tl
2. Prove relation (2), i.e. show that if {xy : y E F] is an orthonormal
basis of a Hilbert space H and T E 28(//) is given by Txy = vyxy,
where vy E C, then T E S60(/f) iff for every e > 0 there are only
finitely many vy of modulus at least e.
3. Let V E 26(//) be such that V2 is a compact positive operator. Is
V necessarily compact? And if V2 = II
4. Let S,TE ®0(H) be such that ||5JC|| = \\Tx\\ for all x E H, with 5
a positive operator. Show that S = (T*T)1/2 = \T\.
5. Let T E 2S(C") be a normal operator (i.e. TT* = T*T). Use ele-
mentary linear algebra to show that the matrix of T is diagonal in
some orthonormal basis.
210 Chapter 14: Compact normal operators

6. Let U be the compact operator defined by a hermitian kernel, as in


Theorem 14. Show that if U is positive then K(s,s) ^ 0 for every
s (0 ^ s ^ 1). Show also that

K(s,t)- 2 kkq>k(s)<pk(t)

is the kernel of a positive operator.


Deduce that

where the series is absolutely and uniformly convergent.


7. Let (xn)i be an orthonormal basis consisting of eigenvectors of a
compact hermitian operator £/, with Uxn = knxn. Let /JL1 ^
/JL2 ^ . •. be the multiset {Xn : An > 0} arranged in a decreasing
order, with }>i,)>2>--- * e corresponding eigenvectors. Putting it
another way: let /AX ^ /i 2 ^ • • • > 0 be the sequence of non-
negative eigenvalues repeated according to their multiplicities.
Show that
/xn = max{(Ux,x): \\x\\ = 1, x±yt for i = l , . . . , / i - l } .
Show also that
fin = minmax{(£/*,*): x E Hn-l9 \\x\\ = 1},

where the minimum is over all (rt-l)-codimensional subspaces


^i-i-
Finally, show that
fxn = maxmin{(Ux,x): x E Hn, \\x\\ = 1},

where the maximum is over all n -dimensional subspaces Fn.


8. Let U be a positive hermitian operator. Show that
\\Ux\\A ^ (Ux,x)(U2x,Ux)
for every vector x. Deduce from this that \\U\\ = v(U).
9. Let U E 2S0(#) be a positive hermitian operator with Ker U = {0}.
Show that there is a sequence of hermitian operators (£/„)* C
» ( # ) such that UnUx-+ x and {/£/„* -* JC for every x E H. Can
one have Un U —> / as well?
10. Let [/ E 2&oC#) be hermitian. Prove that Im £/ is a closed sub-
space of H iff U has finite rank.
Chapter 14: Compact normal operators 211

11. Let T E <&o(H). Prove that Tis


(a) normal iff H has an orthonormal basis consisting of eigenvec-
tors of T;
(b) hermitian iff it is normal and all its eigenvalues are real;
(c) positive iff it is normal and all its eigenvalues are non-
negative reals.
12. Let U E ^o(H) be hermitian. Prove that there are unique positive
operators U+, f/_ E 2S0(#) s u c h that
U = £/+-£/_ and £/+£/_ = £/_!/+ = 0.
13. Prove the Fredholm alternative for hermitian operators: Let U be a
compact hermitian operator on a Hilbert space H and consider the
following two equations:
Ux-x = 0 (2)
and
Ux-x = x0 (3)
where JC0 E H. Then
(a) the only solution of (2) is x = 0, and then (3) has a unique
solution,
or
(b) there are non-zero solutions of (2), and then (3) has a solu-
tion iff x0 is orthogonal to every solution of (2); furthermore,
if (3) has a solution then it has infinitely many solutions: if x
is a solution of (3) then x' is also a solution iff x-x' is a solu-
tion of (2).
14. Give a detailed proof of Theorem 10. In particular, check that the
map *€#(//) —• A . T is a surjection.
15. Let T E 2&oC#) be normal and, as in Theorem 10, for A E C set
nT(k) = dim Ker(A/-T). Prove that nT(\) ^ 1 for every A E C
(including A = 0) iff there is a cyclic vector for T, i.e. a vector
x0 E H such that lin{jc0, 7JC0, T2X0, ...} is dense in //.

Notes

There are many good accounts of applications of the spectral theorem for
compact hermitian opertors to differential and integral equations. We
followed J. Dieudonne, Foundations of Modern Analysis, Academic
Press, New York and London, 1960, xiv + 361 pp. Here are some of the
212 Chapter 14: Compact normal operators

other good books to consult for the Sturm-Liouville problem, Green's


functions, the use of the Fredholm alternative, etc: D. H. Griffel, Applied
Functional Analysis, Ellis Horwood, Chichester, 1985, 390 pp., I. J. Mad-
dox, Elements of Functional Analysis, 2nd edn., Cambridge University
Press, 1988, xii + 242 pp., and N. Young, An Introduction to Hilbert
Space, Cambridge University Press, 1988, vi + 239 pp.
15. FIXED-POINT THEOREMS

In Chapter 7 we proved the doyen of fixed-point theorems, the


contraction-mapping theorem. In this chapter we shall prove some con-
siderably more complicated results: Brouwer's fixed-point theorem and
some of its consequences. It is customary to deduce Brouwer's theorem
from some standard results in algebraic topology, but we shall present a
self-contained combinatorial proof.
Before we can get down to work, we have to plough through some
definitions.
A flat (or an affine subspace) of a vector space V is a set of the form
F = x + W, where W is subspace of V. If W is A:-dimensional then we
call F a k-flat. As the intersection of a set of flats is either empty or a
flat, for every set 5 C V there is a minimal flat F containing S, called
the flat spanned by S. Clearly

F = I 2 A,.^.: jcf E 5, £ A, = 1, n = 1,2,...

Let Xo9Xi,...xk be points in a vector space. We say that these points


are in general position if the minimal flat containing them is k-
dimensional, i.e. if the vectors x1— xo,x3-xo,...,xk—xo span a k-
dimensional subspace. Equivalently, they are in general position if
Mo = Mi = ' •' = M* = 0 whenever 2f =0 M/*i = ° a n d 2f =0 ^ =
° or'
in other words, if the points are distinct and {xx-x0,x2-x0,...,xk-JC0}
is a linearly independent set of vectors.
For 0 ^ k ^ n, let x0, jq,..., xk be k +1 points in IR" in general posi-
tion. The k-simplex a = (JC 0 ,x 1 ,...,x k ) with vertices JC 0 ,x 1 ,...,x k is the
following subset of Un:

213
214 Chapter 15: Fixed-point theorems

M«*i • 2 Mi• = 1 ? M* > 0 for all i


i=0 i=0
The skeleton of o- is the set {xo,Xi,...,xk} and the dimension of o- is &.
Usually we write cr* for a simplex of dimension k and call it a k-
simplex. A 0-simplex is called a vertex.
A simplex o^ is a /ace of a simplex a2 if the skeleton of o^ is a subset
of the skeleton of a2 •
Note that the closure of the simplex a — (x§,xx,...,xk) in Rn is
^ = [xo,x1,...,xk]
k k
2 Wi: 2 Mi: = 1, M.- ^ 0
i=0 1=0

= U{(xio,xil9...,xir): {io,h,...Jr} C {0,1,...,n}},

i.e. the closure of or is precisely the union of all faces of a, including


itself. Also, a is precisely co{x0,Xi,...,xk}, the convex hull of the ver-
tices, and a is the interior of this convex hull in the k -flat spanned by
the vertices.
A finite set K of disjoint simplices in Rn is called a simplicial complex
if every face of every simplex of K is also a simplex of K. We also call
K a simplicial decomposition of the set \K\ = \J {a: a E K}, the body
of K. If K is a simplicial complex and a,r G K then the closed sim-
plices <x and f are either disjoint or meet in a closed face of both.
We are ready to prove the combinatorial basis of Brouwer's theorem.
Lemma 1. (Sperner's lemma) Let AT be a simplicial decomposition of a
closed n-simplex a = [xo,Xi,...,xn]. Let 5 be the set of vertices of K
and let y: 5—> {0, l,...,n} be an (n + 1)-colouring of 5 such that the
colours of the vertices contained in a face [xio, xix,... ,xt ] of a belong to
{/ 0 ,ij,...,i r }. Call an rc-simplex an multicoloured if the vertices of an
are coloured with distinct colours. Then the number of multicoloured
n-simplices of K is odd.
Proof. Let us apply induction on n. For n = 0 the assertion is trivial;
so assume that n ^ 1 and the result holds for n — 1.
Call an (w-l)-face of K marked if its vertices are coloured with
0 , 1 , . . . , n — 1, with each colour appearing once. For an ^-simplex
an G K, denote by m(an) the number of marked (rt-l)-faces of an.
Note that a multicoloured ^-simplex has precisely one marked (n-1)-
face, and an n-simplex, which is not multicoloured, has either no
Chapter 15: Fixed-point theorems 215

marked face or two marked faces. Therefore the theorem claims that
m{K) = 2 m(an) (1)
n
a GK

is odd.
Now let us look at the sum in (1) in another way. What is the contri-
bution of an (n — l)-simplex an~l E K to m{K)l If o-n~~x is not marked,
the contribution is 0. In particular, if an~l is in a closed (n-l)-face of
a other than a0 = [xo,Xi,... ,* n -i] then the contribution of crn~l is 0.
If crn~l E K is in a0 and is marked then the contribution of an~l is 1
since an~l is a face of exactly one n-simplex of K. Furthermore, if
an~l is in a, i.e. in the interior of the original n-simplex, then vn~l
contributes 1 to m(an) if vn~l is a face of <rn: as there are two such
n-simplices an, the total contribution of orn~l to m{K) is 2. Hence,
modulo 2, m(X) is congruent to the number of marked (n — l)-simplices
in a0. By the induction hypothesis, this number is odd. Therefore so is
m(K), completing the proof. •

Given points xo,xi,...,xk of Un in general position, for every point x


of the A:-dimensional affine plane Pk through xo,Xi,...,xk, there are
unique reals Aj, A 2 ,..., A^ such that
k
x = xo+ 2 Kbi-xo)'
i=\
Hence, there are unique reals /Ao^i'-••»/*'* such that x = 2f=0 ^Lt|'JC|'
and 2 = 0 Mi = 1- These /z, (i = 0,1,...,k) are called the barycentric
coordinates of * with respect to (jto,*!,...,**). Also, if 2 = 0 M/ = 1
then 2f=0 Mi*/ e ^ - Furthermore, the closed half-space of P^ contain-
ing xk and bounded by the (fc-l)-flat spanned by xo,x1,...,xk_i is
characterized by ixk ^ 0.
The barycentric coordinates can be used to define a very useful simpli-
cial decomposition. Given a simplicial complex K, the barycentric sub-
division sd K of A' is the simplicial decomposition of | AT | obtained as fol-
lows. For a simplex a = (JC0, JCJ ,...,x k ) E A' set
1 *

thus c^. is the barycentre of o\ The complex sdK consists of all sim-
plices (ca , ca ,..., ca ) such that at E K and 07 is a proper face of 07 +1
k
°
216 Chapter 15: Fixed-point theorems

To define the /--times iterated barycentric subdivision of K, set


sd°A: = K and sdrK = sdCsd'"1*) for r ^ 1. Thus sd1 K = sdK.
The mesh of K, written mesh K, is the maximal diameter of a simplex
of K. Equivalently, it is the maximal length of a 1-simplex of K. Note
that if al• = (x0, Xi,..., JCjr) (/ = 0 , 1 , . . . , k) are faces of a A>simplex
a = ak = (xo,xl9...,xk) and r = ( c ^ c ^ , . . . , ^ ) , then the diameter of
r is less than k/(k+l) times the diameter of a. Therefore, if K is any
simplicial complex then for every e > 0 there is an r such that
meshsd r # < e.
Let Y be a subset of a topological space X, and let a = {Ay : y G F\
be a collection of subsets of X. We call a a covering of Y if
YC U e r ^ V Furthermore, a is a cfosed covering if each A y is
closed, and it is an open covering if each Ay is open. In what follows,
the underlying topological space X is always Un. Sperner's lemma has
the following important consequence.
Corollary 2. Let {A0,Aly...,An} be a closed covering of a closed n-
simplex a = [x0, JCX ,..., xn ] such that each closed face [xio, JC^ ,..., JC,J of
^ is contained in (J r = 0 ^ Then f|"=0 A ^ 0-
Proof. As we may replace At by Al na, we may assume that each At is
compact. The compactness of the sets A§,Ai,...,An implies that it
suffices to show that for every e > 0 there are points at E A{ (i —
0,1,...,n) such that \at-a^\ < e if * ^ j .
Let then e > 0. Let AT be a triangulation of <x such that every simplex
of K has diameter less than 6; as we have seen, for K we may take an
iterated barycentric subdivision of a. Given a vertex x of K contained
in a face (xio,xii,...,xir) of a, we know that JCG U r = 0 ^ - Set
y(x) = min{/; : x E ^ ^ j .
The colouring y of the vertex set of K satisfies the conditions of
Lemma 1 and so K has a multicoloured n -simplex
( a o , a l 9 . . . , a n ) : y(at) = i (i = 0 , 1 , . . . , « ) .
But then at E At, as required. •

From here it is a short step to one of the most fundamental fixed-


point theorems, namely Brouwer's fixed-point theorem. A closed n-cell
is a topological space homeomorphic to a closed n-simplex.
Theorem 3. (Brouwer's fixed-point theorem) Every continuous mapping
of a closed n-cell into itself has a fixed point.
Chapter 15: Fixed-point theorems 217

Proof. We may assume that our n-cell is exactly a closed simplex


&n = [xo,Xi,...9xn]. Suppose that cp : an —> an is a continuous map,
sending a point
n
* = 2 fliX,
to
n
<p{x)
i=0
For each /, let

Then {A0,Al9...,An} is a closed covering of an. If a point


x = 2" = 0 MI^I belongs to a closed face [x^x^,...9xir] of an then /n,- = 0
r
for i £ {i o ,ii,...,i r } and so 2 y =0 M/y = 1- Since 2" = 0 M,' = 1, there is
an index y such that /JL- ^ /xz and so x E ^4,-.. Consequently,

showing that the conditions of Corollary 2 are satisfied. Thus there is a


point x in all the A,; such an * is a fixed point of <p. D

The following lemma enables us to apply Theorem 3 to a rather


pleasant class of spaces, namely the compact convex subsets of finite-
dimensional spaces, i.e. the bounded closed convex subsets of finite-
dimensional spaces.
Lemma 4. Let K be a non-empty compact convex subset of a finite-
dimensional normed space. Then K is an n-cell for some n.
Proof. We may assume that K contains at least two points (and hence it
contains a segment) since otherwise there is nothing to prove.
We may also suppose that K is in a real normed space and hence that
# is a compact convex subset of /2" = (\Rn, \\-\\) for some n. Further-
more, by replacing Rn by the flat spanned by K and translating it, if
necessary, we may assume that 0 E Int K.
Finally, let a be an n-simplex containing 0 in its interior, and define a
homeomorphism cp : K —» a as follows: for x E Rn define
n(x) = nK(x) = inf{t: t > 0, x E tK},
and
218 Chapter 15: Fixed-point theorems

m(x) = m&(x) = inf{t: t > 0, x E to},


and for x E K set
0 if x = 0,

Corollary 5. Let K be a non-empty compact convex subset of a finite-


dimensional normed space. Then every continuous map f:K—>K has
a fixed point.
Proof. This is immediate from Theorem 3 and Lemma 4. •

Our next aim is to prove an extension of Corollary 5 implying, in par-


ticular, that the corollary is true without the restriction that the normed
space is finite-dimensional. This is based on the possibility of approxi-
mating a compact convex subset of a normed space by compact convex
subsets of finite-dimensional subspaces. Unfortunately, the simple
lemma we require needs a fair amount of preparation.
Let S = {xi,...,xk} be a finite subset of a normed space X. For
e > 0 let N(S, e) be the union of the open balls of radius e centred at
X\ >•• • j Xk.

k
N(S,e) = U D(Xi,e).

For x E N(S,e) define Af-(jt) = max{0,6-||JC-^||} (i = 1,...,£) and set


AW = Z* =1 AJ(JC). If x E N(S,e) then x belongs to at least one open
ball D(xi9€), and for that index i we have A,W > 0. Hence AW > 0
for every x E N(S9e). Define the Schauder projection (pSe : N(S,e) -»
co{xl9...,xk] by

1=1

Here

A;X; I Af ^ 0, 2 A; =
l i= \
is the convex hull of the points Jtj,..., xk: the intersection of all convex
sets containing all the points x-l,...,xk. This convex hull is, in fact,
compact, since it is the continuous image of a closed (k-\)-simplex in
R*. Indeed, if (e,-)* is the standard basis of 1$ = (Uk, || • ||), say, then the
Chapter 15: Fixed-point theorems 219

closed simplex a = [ely...,ek] is a bounded closed subset of /2*, and so


it is compact. Furthermore, cp : a —> cofo , . . . , x k } , given by
k k k
2 A,-e,- »-> 2 A/Jc,-, where A, ^ 0 and 2 A,- = 1,
1=1 i=i i=i

is a continuous map.
Lemma 6. The Schauder projection cps € is a continuous map from
N(S, e) to cofo ,..., xn } and
\\<PsJx)-x\\<e (2)
foralljceM5,6).
Proof. Only (2) needs any justification. If x E N(S,e) then

But if A,(JC) > 0 then ||JC,--JC|| < e and so

\\Xi-x\\<€. D

Here then is the promised extension of Corollary 5, Schauder3s fixed-


point theorem.
Theorem 7. Let A be a (non-empty) closed convex subset of a normed
space X and let / : A —> A be a continuous map such that K = /(A) is
compact. Then / has a fixed point.
Proof. Let n ^ 1. As /C is compact, there is afiniteset
£„ = {x1,...,xkn} c A:

such that

Set X^ = COJJCX,...,^ } and denote by <pw the Schauder projection


<Psn : M^/i» V w ) ~~* ^n • We have ifn C i4, so by Lemma 6 the restric-
tion of <pn°f to /JCW is a continuous map of Kn into itself. Hence, by
Corollary 5, there is a point xn E #„ such jthat (pn(f(xn)) = xn. There-
fore, by (2),
220 Chapter 15: Fixed-point theorems

\\f(xn)-xn\\<y (3)
As each/(jcn) belongs to the compact set K, the sequence (f(xn))i has a
convergent subsequence, say f(xnk) -> x as k —> <», where x G AT. But
then, by (3), JCW —> JC as k —> <», and so f(x) = x. D

As a beautiful application of Brouwer's theorem, we prove Perron's


theorem concerning eigenvalues of positive matrices.
Theorem 8. A matrix whose entries are all positive has a positive eigen-
value with an eigenvector whose coordinates are all positive.
Proof. Let A = (a^) be an nxn matrix with atj > 0 for all / and j . Let
(e,-)i be the standard basis in IR". The closed (n — l)-simplex a =
[ei,...,en] is a 'face' of the unit sphere

The continuous map a—> a given by x »-» >1JC/||>1JC||I has a fixed point
•^ = fe)i • Clearly, Ax = AJC for some A > 0 and xt > 0 for all /. D

From Theorem 7 it is a short step to a version of the Markov-


Kakutani fixed-point theorem. An affine map of a vector space V into
itself is a map of the form x »-» XQ + S(x), where 5 : V —> V is a linear
map. Equivalently, 7 : V—» V is an affine map if

whenever xt G V, A, ^ 0 and 2 " = 1 A,- = 1.


Theorem 9. Let K be an non-empty compact convex subset of a normed
space X and let 3F be a commuting family of continuous affine maps on
X such that T(K) C K for all 7 G 9. Then some JC0 G K is a fixed
point of all the maps T G 8F.
Proof. For T G ^ let # r be the set of fixed points of T in K:
KT= {xE K: Tx = x}.
By Theorem 7, KT ^ 0 and, as T is a continuous affine map, KT is a
compact convex subset of K. If 5 G SF then 5 maps # r into itself since
if Tx = x then T(Sx) = S(Tx) = Sx and so Sx E KT. Consequently, if
for some 7\,..., Tn E & and 5 G <&
Chapter 15: Fixed-point theorems 221

n KTI * 0,
i= l
then H w =1 KTi is a compact convex set mapped into itself by S. Hence,
by Theorem 7,

Ksnf) KTi?0.
i= l

This implies that the family of sets {KT: T E 3F} has the finite-
intersection property. As each KT is compact, there is a point x0 which
belongs to every KT, i.e. Tx0 = x0 for every I G ? . •

One should remark that it is easy to prove Theorem 9 without relying


on Theorem 7. Indeed, for T E 9 and n ^ 1 the affine map
n
1

maps JKT into itself and 3^* = {T{n): T E y , n ^ 1} is a commuting fam-


ily of affine maps of K into itself. From this it follows that the system
of compact sets {S(K): 5 E 2P} has the finite-intersection property.
Hence there is a point JC0 such that x0 E T^n\K) for every r E SF and
n ^ 1.
This point JC0 is a fixed point of every 1 6 ? . Indeed, if T{n\yn) =
x0 for yn E K then

Since (yn)i is a bounded sequence, we have r(jc0) = x0.

Exercises

1. Let X be a Banach space and let / : 2?(X) —> ^ be a contraction


from the closed unit ball into X (i.e. d(f(x),f(y)) ^ *:</(*,>0 for all
x,y E B(A0 and some fc < 1). By considering the map g(x) =
j{x + f(x)}, or otherwise, prove that if f(S(X)) C B(X) then / has
a fixed point.
2. Deduce the following assertion from Corollary 2.
Let {Ao,A!,...,i4w} be a closed covering of a closed simplex
ef = [jt 0 ,*i,...,* w ] such that, for each i (0 ^ i ^ n) the set Af is
disjoint from the closed (n-l)-face o://l~1) not containing xt (i.e.
'opposite' the vertex *,). Then n " = 0 Ai ¥" 0-
222 Chapter 15: Fixed-point theorems

3. Use the result in the previous exercise to prove that if <x =


[xo,xi,...,xn] is a closed simplex and / : cf—> (7 is a continuous
map such that for every closed (A:-1)-face f of a we have
/(f) C f, then / is a surjection.
4. Prove that Brouwer's fixed-point theorem is equivalent to each of
the following three assertions, where Bn = B(/") and Sn~x =
S(/f). (In fact, we could take Bn = B{X) and Sn~l = S(X) for
any n -dimensional real normed space X.)
(i) Sn~l is not contractible in itself, i.e. there is no continuous
map 0:Sn~1x[O9i]-^ Sn~x such that for some x0 E S"" 1 we
have <P(x,0) = x0 and <P(x, 1) = x for all JC E Sn~\
(ii) There is no retraction from 5 " onto S"1"1, i.e. there is no
continuous m a p / : Bn -» 5"" 1 such that/(;c) = x for all JC G S"" 1 .
(iii) Whenever / : Bn —» (Rn is a continuous map without a fixed
point then there is a point x E Sn~1 such that x = A/(JC) for some
0 < A < 1.
5. Let C be a closed convex subset of a Hilbert space //. Show that
for every x E H there is a unique point of C nearest to x, i.e.
there is a unique point <p(x) E C such that

Show also that the function cp : H —» C is continuous.


6. Combine Theorem 3 with the assertion in the previous exercise to
deduce Corollary 5.
7. Let Cn = B(l£) be the n-dimensional cube:
Cn = {* = (jt;)i E R": |JC;| ^ 1 for every /}.
The closed faces of Cn are
F+ = {x<E Cn: Xi = 1} and Ft~ = {xECn: xt = -1}
(i = l , . . . , n ) . For each i (i = l,...,/i) let i4,- be a closed subset
of Cw separating Ff and F,", i.e. CnV4, = U^UU', where t/,-+
and f/,~ are disjoint open subsets of Cn, with F^+ C f/,-+ and
F," C (/,"• Prove that n^ = 1 ^ * 0.
8. Let <p be a continuous one-to-one map from a compact Hausdorff
space K into a Hausdorff space. Show that <p is a homeomorphism
between AT and <?(#).
9. Prove that every compact metric space is homeomorphic to a
closed subset of the Hilbert cube:
I™ = {(Xi)i E l2 : \x( | ^ 2~l for every /}.
Chapter 15: Fixed-point theorems 223

10. Prove that every continuous map of the Hilbert cube I°° into itself
has a fixed point. (Check that the map / : B(l2) —> B(l2) given by
f(x) = Sx + (l — \\x\\)ei has no fixed point, where 5 is the right shift
and *! = (1,0,0,...).)
11. Show that a continuous map of B{12) into itself need not have a
fixed point.
12. Let C be a closed subset of a compact metric space K, and let / be
a continuous map of C into a normed space X. Use Schauder pro-
jections and the Tietze-Urysohn extension theorem (Theorem 6.3)
to prove that / has a continuous extension F: K —> X.

The aim of the next three exercises is to make it easy for the reader
to prove another beautiful fixed-point theorem.

13. Let a, b, c and x be points in a Hilbert space such that


b = i(a + c) and
0 < r ^ \\x-a\\ ^ ||*-6|| ^ ||JC-C|| ^ r + e ^ 2r.
Deduce from the parallelogram law that

where 882 = 2re + e2, and so ||tf-c|| ^ 2\/2r6.


14. Let C be a subset of B(l2), and let / : C—• C be a non-expansive
map, i.e. let / be such that d(f{x)J{y)) ^ d(x,y) for all x,y E C.
Suppose xlyx2 and a = lOtj+j^) E C, and ||/(*;)-*,|| ^ e for
i = 1,2. Set c = f(a) and b = \(a + c). Assuming that

check that
lk-6||s ! |k-fl||
and
Ik -4 ^ lki-«ll+*.
Deduce from the result in the previous exercise that

15. Let C be a non-empty closed convex subset of B{12), and let


/ : C —» C be a non-expansive map. Set
224 Chapter 15: Fixed-point theorems

Fn = {xEC:\\f(x)-x\\^l/n},
and show that Fn ^ 0 for all n.
Put dn = inf{||x||: x E Fn} and note that the monotone increas-
ing sequence (dn)i converges to some d ^ 1. Use the result of the
previous exercise to show that
diamFn = sup{||x-)>||: x,y G /„} -> 0 as n -* ».
Conclude from this that / has a fixed point.
16. Let K(s,t) be continuous for 0 ^ s,t ^ 1, and let f(t,u) be continu-
ous and bounded for 0 ^ t ^ 1 and —» < « < oo. Suppose that
|tf(M)| ^ M and \f(t,u)\ ^ N for all s, t and w (0 ^ 5,^ ^ 1).
Define an operator T: C[0,1] -> C[0,1] by

(Tu)(s) = f K(s,t)f(tM*)) &.

Check that T maps C[0,1] into the closed ball of radius MN, say
C= {uE C[0,l]: ||w|| ^ M M -
Apply the Arzela-Ascoli theorem to show that TC is a relatively
compact subset of C[0,1].
Make use of the Schauder fixed-point theorem to prove that the
Hammerstein equation

u{s) = t K(s,t)f(t,u(t))dt
Jo
has a continuous solution.

Notes

This chapter is based on the book of J. Dugundji and A. Granas, Fixed


Point Theory, vol. I, Polish Scientific Publishers, Warsaw, 1982, 209 pp.,
which is a rich compendium of beautiful results. Another interesting
book on the topic is D. R. Smart, Fixed Point Theorems, Cambridge
University Press, 1974. The more usual approach to fixed-point
theorems is via homology, homotopy and degrees of maps; this can be
found in most books on algebraic topology. The case n = 3 of
Brouwer's fixed-point theorem (Theorem 3) was proved in L. E. J.
Brouwer, On continuous one-to-one transformations of surfaces into
themselves, Proc. Kon. Ned. Ak. V. Wet. Ser. A, 11 (1909), 788-98;
the first proof of the full result was given by J. Hadamard, Sur quelques
Chapter 15: Fixed-point theorems 225

applications de Vindice de Kronecker; Appendix in J. Tannery, Introduc-


tion a la Thiorie des Fonctions d'une Variable, vol. II, 2me ed., 1910;
Brouwer himself proved the general case in 1912. Sperner's lemma is
from E. Sperner, Neuer Beweis fiir die Invarianz der Dimensionzahl und
des Gebietes, Abh. Math. Sem. Hamb. Univ., 6 (1928), 265-72, and
Theorem 7 is from J. Schauder, Der Fixpunktsatz in Funktionalrdumen,
Studia Math., 2 (1930), 171-80. The original version of Theorem 9 is in
A. A. Markoff, Quelques theorimes sur les ensembles abeliens, C. R.
Acad. Sci. URSS (N.S), 1 (1936), 311-3.
16. INVARIANT SUBSPACES

Given a complex Banach space X, which operators T E <3b(X) have non-


trivial closed invariant subspaces? This question, the so-called invariant-
subspace problem, is the topic of this brief last chapter. Until fairly
recently, it was not known whether there was any operator T without a
non-trivial (closed) invariant subspace, and it is still not known whether
there is such an operator on a (complex) Hilbert space.
Much of the effort concerning the invariant-subspace problem has
gone into proving positive results, i.e. results claiming the existence of
invariant subspaces for operators satisfying certain conditions. Our
main aim in this chapter is to present the most beautiful of these posi-
tive results, Lomonosov's theorem, whose proof is surprisingly simple.
As we remarked earlier, the Riesz theory of compact operators on
Banach spaces culminated in a very pleasing theorem, Theorem 13.8,
which nevertheless, did not even guarantee the existence of a single non-
trivial invariant subspace. This deficiency was put right, with plenty to
spare, in Chapter 14, but only for a compact normal operator on a Hil-
bert space. Now we return to the general case to prove Lomonosov's
theorem, which claims considerably more than that every compact
operator has a non-trivial invariant subspace. Before we present this
result, we need some definitions and a basic result about compact con-
vex sets.
As in Chapters 13 and 14, all spaces considered in this chapter are
complex spaces. Furthermore, as every linear operator on a finite-
dimensional complex vector space has an eigenvector, we shall consider
only infinite-dimensional spaces.
Given a Banach space X and an operator T E 2S(X), call a subspace
Y C X invariant under T or T-invariant if Y is closed and TY C Y. We

226
Chapter 16: Invariant subspaces 227

also say that Y is an invariant subspace of T. Clearly, Y = {0} and


Y = X are T-invariant subspaces for every T, so we are interested only
in other invariant subspaces, the so-called non-trivial invariant sub-
spaces.
We call a subspace Y a hyperinvariant subspace for T if Y is an 5-
invariant subspace for every S E <3l(X) commuting with T. Since T
commutes with itself, every hyperinvariant subspace is an invariant sub-
space.
Note that if T and S commute then KerT is 5-invariant since if
x E Ker T then TS(x) = ST(x) = 5(0) = 0 and so S(x) E Ker T. Hence
if T has no non-trivial hyperinvariant subspaces then either T is a multi-
ple of the identity or it has no eigenvalue, i.e. (rp(T) = 0 . In particu-
lar, if T E &0(X) and a(T) ? {0} then, by Theorem 8 of Chapter 13,
the operator T has a non-trivial hyperinvariant subspace.
What is the easiest way of constructing a T-in variant subspace? Pick
a vector x E X (x ± 0) and set
Y0(x) = lin{x,Tx,T2x,...}.
Then TY0(x) C Y0(x) and so, as T is continuous,
Y(x) = Wx) = M{x,Tx,T2x,...}
is a T-invariant subspace. Hence if T has no non-trivial invariant sub-
space then Y(x) = X for all x ^ 0. A vector x is said to be a cyclic vec-
tor for T if Y(x) = X.
Thus if T has no non-trivial invariant subspace then every non-zero
vector is a cyclic vector for T. The converse of this is also trivially true:
if every non-zero vector is a cyclic vector for T then T has no non-trivial
invariant subspace since if Y is a non-trivial invariant subspace then no
vector in Y is a cyclic vector for T.
Unfortunately, but not surprisingly, this shallow argument is no
advance on the invariant-subspace problem; nevertheless, it tells us that
we have to concentrate on cyclic vectors.
In the proof of Lomonosov's theorem, we need a basic result concern-
ing closed convex hulls of compact sets.
Theorem 1. (Mazur's theorem) The closed convex hull of a compact set
in a Banach space is compact.
Proof. Let A be a compact subset of a Banach space X and let
K = coA. We have to show that K is totally bounded, i.e. for every
e > 0 it contains a finite e-net. Since A is compact, it contains a finite
228 Chapter 16: Invariant subspaces

56-net, say {x1,...,xn}. Thus {xl9...,xn} CA and for every x E A


there is an xt such that |JC —jcf-|| < 56. The set P = cof*!,...,x n } is also
compact, and so it contains a finite 56-net { y i , . . . , y m } . Furthermore,
the set

is convex and contains A, and therefore contains coA. But then with
M = {xE. X: \\x-yi\\ < | e for some 1}
we have coA C Pi€ C M and so {y±,... ,ym} is an e-net in K = do A.
D

Here then is the main result of the Chapter.


Theorem 2. (Lomonosov's first theorem) Let T be a non-trivial compact
operator on an infinite-dimensional complex Banach space X. Then T
has a non-trivial hyperinvariant subspace.
Proof. We may assume that ||r|| = 1. Set
SA = {T}' = {SB a(J0 : ST = TS}
and pick a point x0 E X such that ||JCO|| > ||r||. Set Bo = B(xo,l) and
note that 0 <£ Bo and 0 E 7 ^ .
Suppose first that there is a point y0 E Jf (y0 ^ 0) such that
l|7>o-*oll ^ 1 (1)
for every T E: SA. Then we are done since
Y = hn{T'y0 : T E
is a non-trivial hyperinvariant subspace of X.
Suppose then that there is no y0 ^ 0 satisfying (1). Then for every
y E X (y =fi 0) there is an operator T E &4 such that

Since TB0 is a compact set not containing 0, there are operators


Tl9..., Tn E SA such that for every y E 7 ^ there is a 7} (1 ^ 1 ^ n)
satisfying
ll^-Joll < I- (2)
Now we define a map 1/^: TB0 —> X reminiscent of the Schauder pro-
jection. For y E 77?0 and 1 ^ / ^ n set
Chapter 16: Invariant subspaces 229

ki(y) = max{0, \~\\Tj\\}


and

Hy) = £ A,O0.
1=1

Relation (2) implies that X(y) > 0 for every y E 77^. Therefore we
may define

This map i//: TB0 —> Jf is continuous and so T5 0 is a compact subset of


Bo. Consequently, by Mazur's theorem, K = c6TB0 is a compact con-
vex subset of Bo. Hence

is a continuous map of a compact convex set into itself and so, by


Schauder's theorem (Theorem 15.7), it has a fixed point z0 E K:

Set

Then 5 E 5A, Sz0 = z0 + 0, and so


y = Ker(/-5) ^ {0}
is a T-invariant subspace. As 5 is compact, Y is finite-dimensional. But
then T\Y is an operator on a complex finite-dimensional space and so it
has an eigenvalue A. However then Ker(A7— T) is a non-trivial hyperin-
variant subspace for T. •

A slight variation in the proof shows that an even larger class of


operators have hyperinvariant subspaces.
Theorem 3. (Lomonosov's second theorem) If T E ^{X) commutes
with a non-zero compact operator and is not a multiple of the identity
then it has a hyperinvariant subspace.
230 Chapter 16: Invariant subspaces

Proof. Suppose TT0 = T0T for some To E SBQC^O (^o ^ 0). Proceed as
in the proof of Theorem 2 but consider T0B0 instead of TBQ and so put
K = coT 0 B 0 . Then we obtain a fixed point z0 E K of if/°T0: K^> K,
i.e.

5z 0 = z 0
for

Therefore
Y= Ker(/-5) ^ {0}
is a finite-dimensional T-invariant subspace, and now the proof is com-
pleted as before. •

Let us point out the following special case of Theorem 3.


Corollary 4. Let S,T E ^(X) be commuting operators such that 5 com-
mutes with a non-zero compact operator, and is not a multiple of the
identity. Then T has a non-trivial invariant subspace. •

As a rather special case of this corollary one obtains a theorem of


Aronszajn and Smith, proved considerably earlier.
Corollary 5. Every compact operator on an infinite-dimensional complex
Banach space has a non-trivial invariant subspace. •

An extension of this result, first proved by Bernstein and Robinson,


needs only a little work.
Corollary 6. Let X be an infinite-dimensional complex Banach space
and let T E ®CY) be such that p(T) E ®0(X) for s o m e non-zero com-
plex polynomial p(z). Then T has a non-trivial invariant subspace.
Proof. Let
n
a
p(z) = 2 k*k, withan + 0.
k=0
If p(T) ± 0 then, as Tp{T) = p(T) T, the assertion follows from
Theorem 2.
Chapter 16: Invariant subspaces 231

If, on the other hand, p(T) = 0, then anTn = - ^ 1 * akTk, and so


Y{x) = lin{jc, 7JC, ..., r"" 1 *} is a ^-invariant subspace for every x ^ 0.
n
In spite of the simplicity of its proof, Lomonosov's second theorem is a
very powerful result. At the moment, it is not clear how large a class of
operators Corollary 4 applies to; in fact, for a while it was not clear that
there is any operator T E ^h{X) which is not covered by Corollary 4.
The invariant-subspace problem for Banach spaces was solved, in the
negative, only fairly recently: Per Enflo and Charles Read constructed
complex Banach spaces and bounded linear operators on them which do
not have non-trivial invariant subspaces. The original proofs were for-
midably difficult and the spaces seemed to be rather peculiar spaces.
Later, Charles Read gave an easily accessible proof, and showed that his
construction works, in fact, on l1.
In view of these great results, the invariant-subspace problem for Hil-
bert spaces has become a very major problem in functional analysis. In
fact, it is not impossible that the answer is in the affirmative even on
reflexive spaces, i.e. that every bounded linear operator on an infinite-
dimensional reflexive complex Banach space has a non-trivial invariant
subspace.

Exercises

1. Let I be a non-separable Banach space. Show that every


T E <3&(X) has a non-trivial invariant subspace.
2. Show that the following result can be read out of the proof of
Theorem 2. Let SA be a subalgebra of ^(X) whose elements do
not have a non-trivial common invariant subspace. Then if
T E SftoUQ anc* T 7^ 0 then there is an operator A E SA such
that K e r ( / - ^ 7 ) ^ {0}.
3. Let Tly...,Tn E ?&o(X) be commuting operators. Show that they
have a non-trivial common invariant subspace.
4. Deduce from Theorem 1 and Exercise 5.5 the following extension of
Theorem 4.10. If the unit ball of a. Banach space X is <7-compact then X
is finite-dimensional.
5 + + + . Solve the invariant-subspace problem for Hilbert spaces.
232 Chapter 16: Invariant subspaces

Notes

Mazur's theorem is from S. Mazur, Uber die kleinste konvexe Menge, die
eine gegebene kompakte Menge enthdlt, Studia Math., 2 (1930), 7-9.
Theorems 2 and 3 are from V. I. Lomonosov, On invariant subspaces of
families of operators, commuting with a compact operator (in Russian),
Funk. Analiz i ego Prilozh, 7 (1973), 55-6; to be precise, Theorem 3 is
given as a remark added in proof. Corollary 5 is from N. Aronszajn
and K. Smith, Invariant subspaces of completely continuous operators,
Ann. Math., 60 (1954), 345-50.
The invariant-subspace problem for Banach spaces was solved in P.
Enflo, On the invariant subspace problem in Banach spaces, Acta Math.,
158 (1987), 213-313, and C. J. Read, A solution to the Invariant Sub-
space Problem, Bull. London Math. Soc, 16 (1984), 337-401.
A simplified and stronger version of Enflo's solution can be found in
B. Beauzamy, Un operateur sans sous-espace invariant non-trivial:
simplification de Vexample de P. Enflo, Integral Equations and Operator
Theory, 8 (1985), 314-84.
Read's result concerning lx is in A solution to the Invariant Subspace
Problem on the space lx, Bull. London Math. Soc, 17 (1985), 305-17.
An interesting account of the results concerning the invariant-
subspace problem can be found in B. Beauzamy, Introduction to Opera-
tor Theory and Invariant Subspaces, North Holland, Amsterdam, 1988,
xiv + 358 pp.
INDEX OF NOTATION

B(X), closed unit ball, 22 CQ(L) , space of continuous functions


B(xo,r), closed ball of radius r and vanishing at infinity, 91
centre JC0, 22 CR(L), space of bounded continuous
£r(jc0), closed ball of radius r and cen- real-valued functions, 93
tre x0, 22 C0R(L), space of continuous real-
9&WQ, space of bounded linear opera- valued functions vanishing at
tors on X, 28 infinity, 93
<3l(X, Y), space of bounded linear
operators, 28 d(X9 Y), Banach-Mazur distance
, Y), the space of compact opera- between X and Y, 66
tors, 186 D(x,r), open ball, 20
C Y), the space of finite rank D(jco,r), open ball of radius r and cen-
operators, 186 t r e ^ , 22
0B2WO> t h e s P a c e o f Hilbert-Schmidt £>r(jt0), open ball of radius r and cen-
operators, 187 tre JC0, 22
8A(a), the resolvent set of a in the
ca, the barycentre of the simplex a, algebra^, 167
215 A, closed unit disc in the complex
co 5, the closed convex hull of S, 55 plane, 96
co 5, the convex hull of 5, 55
C*-algebra, 167 fvg, the join of / and g, 93
C(K), space of continuous functions /7\g, the meet of / and g, 93
on a compact Hausdorff space K, f\S, restriction of a / to 5, 25
23 ^b{S), space of bounded functions on
C(L), space of bounded continuous
functions on L, 23 5,23
CC(L), space of continuous functions
with compact support, 91 Im T, image of T, 28
C®(L), space of continuous real-
valued functions with compact k-simplex, 213
support, 93 K*, annihilator of K, 164
K°, polar of K, 158

233
234 Index of notation

/ r norm, 23 a(X, &), weak topology generated by


/p-norm, 23 9, 115
linS, linear span, 38 cr(X,X*), weak topology on a normed
linS, closed linear span, 38 space, 115
linZ, linear span, 21 o{X*,X)9 weak-star topology, 116
LQO, one-point compactification of L, (TA{a), the spectrum of a in the alge-
96 b r a ^ , 167
a
L, preannihilator of L, 164 aap(T), the approximate point spec-
°L, prepolar of L, 158 trum of the operator T, 169
!£{X, Y), space of linear operators, 28 (7c(r), the continuous spectrum of the
operator T, 169
mesh K, the mesh of the simplicial <rcom(T), the compression spectrum of
complex A', 216 the operator 7\ 168
cr p (r), the point spectrum of the
PF, orthogonal projection onto F, 136 operator T, 168
<p-mean, 5 aT(T), the residual spectrum of the
operator T, 169
rn(t), Rademacher function, 143
r(7), the spectral radius of T, 174
T-invariant, 226
Rad#, the radical of the Banach alge-
T*, adjoint of T, 31
bra B, 178
||r|| 2 , Hilbert-Schmidt norm, 187
p{T), the resolvent set of the operator
||r|| HS , Hilbert-Schmidt norm, 187
r, 168

sd K, the barycentric subdivision of K, v(T), the numerical radius of T, 202


215 V(T), the numerical range of T, 201
S1, set of vectors orthogonal to S, 135
S(X), unit sphere, 22 x1, set of vectors orthogonal to x, 135
S(xo,r), sphere of radius r and centre
x0, 22 X, completion of X, 35
Sr(x0), sphere of radius r and centre X', space of linear functionals on X,
*o>22 28,45
ai<(°\ set of finite subsets of o-, 114 X*, dual of X, 31
cr{T), the spectrum of the operator T, A"*, space of bounded linear function-
167 als on X, 28, 45
INDEX OF TERMS

absolute value of an operator, 205 Banach's fixed-point theorem, 101


absolutely convergent series, 36 Banach-Mazur distance, 66
absolutely convex set, 27 Banach-Steinhaus theorem, 78
adjoint of an operator, 155 barycentre, 215
adjoint operator, 31 barycentric coordinates, 215
affine hyperplane, 46 bary centric subdivision, 215
affine map, 220 basic sequence, 72
affine subspace, 213 basis, 19, 83
Alaoglu's theorem, 118 canonical, 37
algebra, commutative, 92 Hamel, 42
algebraic dual of a normed space, 45 Schauder, 37, 83
AM-GM inequality, 1 standard, 37
analytic, 171 basis constant, 73, 83
annihilator of a set, 164 Bernstein and Robinson, theorem of,
annihilator of a subspace, 158 230
approximate eigenvector, 169 Bessel's inequality, 147
approximate point spectrum, 169 biorthogonal system, 64
approximation problem, 189 normalised, 64
arithmetic mean, 1, 6 Bishop-Phelps-Bollobas theorem, 122
weighted, 7 body, 214
Arzela-Ascoli theorem, 90 bounded below, 162
Auerbach system, 65 bounded linear operator, 28
bracket notation, 28
Banach algebra, 92 Brouwer's fixed-point theorem, 216
unital, 32
Banach limit, 59 canonical basis, 37
Banach space, 21 Carleson's theorem, 150

235
236 Index of terms

Cauchy sequence, 21 contraction with constant k, 101


Cauchy-Schwarz inequality, 9, 131 contraction-mapping theorem, 101
Cesaro summable, 150 convergent series, 36
chain, 50 convex function, 3
classical function spaces, 27 convex functional, 47
classical sequence spaces, 26 convex hull, 55
closed, 19 convex subset, 2
closed n-cell, 216 countably compact, 89
closed convex hull, 55 covering, 216
closed covering, 216 closed, 216
closed-graph theorem, 80 open, 216
closed linear span, 38 cyclic vector, 211, 227
closure, 19
degenerate form, 130
coarser topology, 20
dense subset, 33
commutative algebra, 92
dense topological space, 75
compact, 89
derivative, 105
countably, 89
differentiable map, 105
relatively, 89
dimension, 214
sequentially, 89
direct sum of subspaces, 39
compact operator, 186
dissipative operator, 182
compact space, 21
distance, 19
complete, 21
divisor of zero, 180
complete orthogonal set of vectors,
topological, 180
141
dominate, 48
completely regular topological space,
dual of an operator, 155
98
dual space, 31
completion of a metric space, 33
completion of a normed space, 35 eigenspace, 168
complex unital Banach algebra, 167 eigenvalue, 168
compression spectrum, 168 eigenvector, 168
concave function, 3 approximate, 169
concave functional, 53 Enflo's theorem, 189, 231
conjugate, 10 equicontinuous, 90
constant k equicontinuous at a point, 90
contraction with, 101 equivalence class of functions, 24, 25
Lipschitz condition with, 101 equivalent norms, 29
continuous maps, 20 essential supremum, 25
continuous spectrum, 169 Euclidean space, 23, 132
contractible in itself, 222 extension of a linear functional, 47
contraction, 101 extreme point, 125
Index of terms 237

face, 214 Hadamard's inequality, 151


marked, 214 Hahn-Banach extension theorem, 50
Fejer's theorem, 150 complex form, 50
finer topology, 20 Hamel basis, 42
finite character, 116 Hammerstein equation, 224
finite-intersection property, 116 harmonic mean, 6
finite rank operator, 186 Hausdorff topology, 21
fixed point, 101 Hermite polynomials, 144
flat, 213 hermitian form, 8, 130
form hermitian form, positive, 8
degenerate, 130 hermitian operator, 159
hermitian, 130 on a normed space, 182
non-degenerate, 130 Hilbert space, 27, 132
positive, 130 Hilbert-Schmidt norm, 187
symmetric, 130 homeomorphic spaces, 20
Fourier coefficent, 145, 149 homeomorphism, 20
Fourier series, 145 hyperinvariant subspace, 227
Fredholm alternative for hermitian hyperplane, 46
operators, 211 Holder's inequality, 9
Fredholm integral equation, 207 for functions, 12
function
concave, 3 image, 28
convex, 3 in general position, 213
strictly concave, 3 induced topology, 19
strictly convex, 3 inequality
fundamental set of vectors, 141 AM-GM, 1
Cauchy-Schwarz, 9
Gelfand transform, 183 Holder's, 9
Gelfand's spectral-radius formula, 174 Minkowski's, 10
Gelfand-Mazur theorem, 174 initial segment, 124
Gelfand-Naimark theorem, 158 inner product, 131
generalized limit, 59 inner-product space, 131
geometric mean, 1 Int, interior, 22
weighted, 7 invariant subspace, 226, 227
Gluskin's theorem, 70 invariant-subspace problem, 226
Gram determinant, 152 invariant under, 226
Gram-Schmidt orthogonalization pro- inverse, 167
cess, 142 inverse-mapping theorem, 80
invertible, 167
Haar system, 83 involution, 167
238 Index of terms

isometrically isomorphic spaces, 29 mean


isometry, 162 arithmetic, 6
isomorphic spaces, 29 harmonic, 6
isotropic vectors, 130 quadratic, 6
meet, 93
Jensen's theorem, 3 mesh, 216
John's theorem, 68 method of successive approximations,
Johnson's uniqueness-of-norm 103
theorem, 179 metric, 19
join, 93 metric space, 19
completion of a, 33
kernel, 28, 108 Minkowski functional, 28
kernel of an integral operator, 207 Minkowski's inequality, 10, 131
Krein-Milman theorem, 126 for functions, 12
modulus of an operator, 205
Laguerre polynomials, 144 multicoloured simplex, 214
lattice operations, 93
Laurent series, 171 n-dimensional Euclidean space, 27
Lebesgue number, 99 neighbourhood, 19
left shift, 32 neighbourhood base, 20
Legendre polynomial, 143 non-degenerate form, 130
linear functional, 28 non-expansive map, 223
linear operator, 28 non-trivial, 227
bounded, 28 norm, 18, 19
unbounded, 28 Hilbert-Schmidt, 187
linear span, 21, 38 operator, 29
Lipschitz condition with constant k, of a functional, 30
101 smooth, 51
local theory of Banach spaces, 70 supremum, 23, 92
locally compact space, 91 uniform, 23, 92
Lomonosov's first theorem, 228 norm topology, 21
Lomonosov's second theorem, 229 normal operator, 161
normalised biorthogonal system, 64
marked face, 214 normed algebra, 92
Markov-Kakutani fixed-point unital, 92
theorem, 220 normed space, 18
maximal element, 50 completion of a, 35
maximal ideal space, 183 nowhere dense set, 76
Mazur's theorem, 227 nowhere dense subset, 47
meagre set, 76 numerical radius, 202
Index of terms 239

numerical range, 201 quadratic mean, 6


spatial, 175 quotient norm, 38
quotient normed space, 38
one-point compactification, 96
open covering, 216 Rademacher function, 143
open-mapping theorem, 79 radical, 178
operator ideal, 188 Read's theorem, 231
operator norm, 29 regular point, 167
order, 49 relatively compact, 89
partial, 49 residual spectrum, 169
ordered set, 124 resolvent, 167
orthogonal complement, 136 resolvent identity, 181
orthogonal direct sum, 135 resolvent set, 167
orthogonal matrix, 140 restriction of a function, 25
orthogonal set of vectors, 141 Riemann-Lebesgue lemma, 152
orthogonal subspaces, 135 Riesz representation theorem, 137
orthogonal vectors, 130 Riesz-Fischer theorem, 145
right shift, 32
parallelogram law, 133
Parseval's identities, 147 scalar product, 131
partial order, 49 Schauder basis, 37, 83
partial sum, 149 Schauder projection, 218
partially ordered set, 50 Schauder system, 83
partition of unity, 100 Schauder's fixed-point theorem, 219
Perron's theorem, 220 second dual, 156
point, 19 self-adjoint operator, 159
point spectrum, 168 seminorm, 41
polar of a set, 158 separable, 22
polarization identities, 132 separate, 89
positive form, 130 separates the points strongly, 96
positive hermitian form, 8 separation theorem, 54
pre-Hilbert space, 132 sequentially compact, 89
preannihilator set of the first category, 76
of a set, 164 set of the second category, 76
of a subspace, 158 set system of finite character, 116
prepolar of a set, 158 simplex, 213, 214
principle of uniform boundedness, 77 multicoloured, 214
probability, 5 simplicial complex, 214
product topology, 20, 115 simplicial decomposition, 214
Pythagorean theorem, 133 skeleton, 214
240 Index of terms

smooth norm, 51 norm, 21


spatial numerical range, 175 product, 20, 115
spectral decomposition, 200 strong operator, 165
spectral measure, 209 stronger, 20
spectral radius, 174 subspace, 19
spectrum, 167 weak, 115
approximate point, 169 weaker, 20
compression, 168 total set of vectors, 141
continuous, 169 totally, 89
point, 168 totally ordered set, 50
residual, 169 translate of a subspace, 46
Sperner's lemma, 214 triangle inequality, 18
standard basis, 37 Tukey's lemma, 116
Stone-Weierstrass theorem, 95 Tychonov's theorem, 117
for complex functions, 96
strictly concave function, 3 unbounded linear operator, 28
strictly convex function, 3 uniform closure, 93
strong operator topology, 165 uniform norm, 23, 92
stronger topology, 20 uniformly bounded, 90
Sturm-Liouville equation, 209 unital Banach algebra, 32
sub-basis for a topology, 114 unital normed algebra, 92
subadditive functional, 48 unitarily equivalent operators, 206
subreflexive space, 122 unitary operator, 161
subspace, 21 universal property, 115
subspace topology, 19 upper bound, 50
sum of a series, 36 Urysohn's lemma, 86
superadditive functional, 53
vanishing at oo, 96
support functional, 51, 175
vector, 19
support plane, 51
vertex, 214
supremum norm, 23, 92
Volterra integral operator, 108
symmetric form, 130
system set of vectors, 141 weak topology, 115
weak-star topology a(X*,X), 116
Tietze-Urysohn extension theorem, weaker topology, 20
topological divisor of zero, 180 weakly bounded, 81
topological space, 19 weight, 5
topology, 19 weighted arithmetic mean, 7
coarser, 20 weighted geometric mean, 7
finer, 20 well-ordered set, 124
Hausdorff, 21
induced, 19 Zorn's lemma, 50

You might also like