Sie sind auf Seite 1von 343

Canadian Mathematical Society

Societe mathematique du Canada


Editors-in-Chief
Redacteurs-en-chef
Jonathan Borwein
Peter Borwein

Springer
New York
Berlin
Heidelberg
Barcelona
Hong Kong
London
Milan
Paris
Singapore
Tokyo
CMS Books in Mathematics
Ouvrages de mathematiques de la SMC
1 HERMAN/KuCERAfSIMSA Equations and Inequalities
2 ARNOLD Abelian Groups and Representations of Finite Partially Ordered Sets
3 BORWEIN/LEWIS Convex Analysis and Nonlinear Optimization
4 LEVIN/LuBINSKY Orthogonal Polynomials for Exponential Weights
5 KANE Reflection Groups and Invariant Theory
6 PHILLIPS Two Millennia of Mathematics
7 DEUTSCH Best Approximations in Inner Product Spaces
Frank Deutsch

Best Approximation
in Inner Product Spaces

Springer
Frank Deutsch
Department of Mathematics
Pennsylvania State University
417 McAllister Bldg.
University Park, PA 16802-6401
USA
deutsch@math.psu.edu

Editors-in-ChieJ
Redacteurs-en-cheJ
Jonathan Borwein
Peter Borwein
Centre for Experimental and Constructive Mathematics
Department of Mathematics and Statistics
Simon Fraser University
Burnaby, British Columbia V5A IS6
Canada

Mathematics Subject Classification (2000): 41A50, 41A65

Library of Congress Cataloging-in-Publication Data


Deutsch, F. (Frank), 1936-
Best approximation in inner product spaces 1 Frank Deutsch.
p. cm. - (CMS books in mathematics ; 7)
Includes bibliographical references and index.

1. Inner product spaces. 2. Approximation theory. 1. Title. II. Series.


QA322.4 .D48 2001
515' .733-dc21 00-047092

Printed on acid-free paper.


ISBN 978-1-4419-2890-0 ISBN 978-1-4684-9298-9 (eBook)
DOI 10.1007/978-1-4684-9298-9
© 2001 Springer-Verlag New York, Inc.
Softcover reprint of the hardcover 1st edition 2001
All rights reserved. This work may not be translated or copied in whole or in part without the
written permission of the publisher (Springer-Verlag New York, Inc., 175 Fifth Avenue, New York,
NY 10010, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use
in connection with any form of information storage and retrieval, electronic adaptation, computer
software, or by similar or dissimilar methodology now known or hereafter developed is forbidden.
The use of general descriptive names, trade names, trademarks, etc., in this publication, even if the
former are not especially identified, is not to be taken as a sign that such names, as understood by
the Trade Marks and Merchandise Marks Act, may accordingly be used freely by anyone.

Production managed by Michael Koy; manufacturing supervised by Jeffrey Taub.


Photocomposed copy prepared using the author's AMS-'IEX files.

9 8 765 432 1

SPIN 10784583

Springer-Verlag New York Berlin Heidelberg


A member 01 BertelsmannSpringer Science+Business Media GmbH
To Mary: a wonderful wife, mother, and mema
An approximate answer to the right problem is worth a good deal more than an
exact answer to an approximate problem.

-John Thkey

What is written without effort is in general read without pleasure.

-Samuel Johnson
PREFACE

This book evolved from notes originally developed for a graduate course, "Best
Approximation in Normed Linear Spaces," that I began giving at Penn State Uni-
versity more than 25 years ago. It soon became evident. that many of the students
who wanted to take the course (including engineers, computer scientists, and statis-
ticians, as well as mathematicians) did not have the necessary prerequisites such
as a working knowledge of Lp-spaces and some basic functional analysis. (Today
such material is typically contained in the first-year graduate course in analysis.)
To accommodate these students, I usually ended up spending nearly half the course
on these prerequisites, and the last half was devoted to the "best approximation"
part. I did this a few times and determined that it was not satisfactory: Too much
time was being spent on the presumed prerequisites. To be able to devote most of
the course to "best approximation," I decided to concentrate on the simplest of the
normed linear spaces-the inner product spaces-since the theory in inner product
spaces can be taught from first principles in much less time, and also since one can
give a convincing argument that inner product spaces are the most important of
all the normed linear spaces anyway. The success of this approach turned out to
be even better than I had originally anticipated: One can develop a fairly complete
theory of best approximation in inner product spaces from first principles, and such
was my purpose in writing this book.
Because of the rich geometry that is inherent in inner product spaces, most of the
fundamental results have simple geometric interpretations. That is, one can "draw
pictures," and this makes the theorems easier to understand and recall. Several
of these pictures are scattered throughout the book. This geometry also suggests
the important role played by "duality" in the theory of best approximation. For
example, in the Euclidean plane, it is very easy to convince yourself (draw a picture!)
that the distance from a point to a convex set is the maximum of the distances from
the point to all lines that separate the point from the convex set. This suggests
the conjecture (by extrapolating to any inner product space) that the distance
from a point to a convex set is the maximum of the distances from the point to
"hyperplanes" that separate the point from the set. In fact, this conjecture is true
(see Theorem 6.25)! Moreover, when this result is formulated analytically, it states
that a certain minimization problem in an inner product space has an equivalent
formulation as a maximization problem in the dual space. This is a classic example
of the role played by duality in approximation theory.
Briefly, this is a book about the "theory and application of best approximation
in inner product spaces" and, in particular, in "Hilbert space" (i.e., a complete in-
ner product space). In this book geometric considerations playa prominent role in
developing and understanding the theory. The only prerequisites for reading it are
some "advanced calculus" and a little "linear algebra," where, for the latter subject,
viii PREFACE

a first course is more than sufficient. Every author knows that it is impossible to
write a book that proceeds at the correct pace for every reader. It will invariably
be too slow for some and too fast for others. In writing this book I have tried to
err on the side of including too much detail, rather than too little, especially in the
early chapters, so that the book might prove valuable for self-study as well. It is
my hope that the book proves to be useful to mathematicians, statisticians, engi-
neers, computer scientists, and to any others who need to use best approximation
principles in their work.
What do we mean by "best approximation in inner product spaces"? To explain
this, let X be an inner product space (the simplest model is Euclidean n-space)
and let K be a nonempty subset of X. An element Xo in K is called a best approx-
imation to x from K if Xo is closest to x from among all the elements of K. That
is, Ilx - xoll = inf{llx - yll lyE K}. The theory of best approximation is mainly
concerned with the following fundamental questions: (1) Existence of best approx-
imations: When is it true that each x in X has a best approximation in K? (2)
Uniqueness of best approximations: When is it true that each x in X has a unique
best approximation in K? (3) Characterization of best approximations: How does
one recognize which elements in K are best approximations to x? (4) Continuity
of best approximations: How do best approximations in K to x vary as a function
of x? (5) Computation of best approximations: What algorithms are available for
the actual computation of best approximations? (6) Error of approximation: Can
one compute the distance from x to K, i.e., inf{llx - yll lyE K}, or at least get
good upper and/or lower bounds on this distance?
These are just a few of the more basic questions that one can ask concerning
best approximation. In this book we have attempted to answer these questions,
among others, in a systematic way. Typically, one or more general theorems valid
in any inner product space are first proved, and this is then followed by deducing
specific applications or examples of these theorems. The theory is the richest and
most complete when the set K is a closed and convex set, and this is the situation
that the bulk of the book is concerned with. It is well known, for example, that if
K is a (nonempty) closed convex subset of a Hilbert space X, then each point x in
X has a unique best approximation PK(x) in K. Sets that always have unique best
approximations to each point in the space are called Chebyshev sets. Perhaps the
major unsolved problem in (abstract) best approximation theory today is whether
or not the converse is true. That is, must every Chebyshev set in a Hilbert space
be convex? If X is finite-dimensional, the answer is yes. Much of what is known
about this question is assembled in Chapter 12.
The book is organized as follows. In Chapter 1, the motivation for studying best
approximation in inner product spaces is provided by listing five basic problems.
These problems all appear quite different on the surface. But after defining inner
product spaces and Hilbert spaces, and giving several examples of such spaces,
we observe that each of these five problems is a special case of the problem of best
approximation from a certain convex subset of a particular inner product space. The
general problem of best approximation is discussed in Chapter 2. Existence and
uniqueness theorems for best approximations are given in Chapter 3. In Chapter
4 a characterization of best approximations from convex sets is given, along with
several improvements and refinements when the convex set K has a special form
(e.g., a convex cone or a linear subspace). When K is a Chebyshev set, the mapping
x >-+ PK(x) that associates to each x in X its unique best approximation in K is
PREFACE ix

called the metric projection onto K. In Chapter 5 a thorough study is made of the
metric projection. In particular, PK is always nonexpansive and, if K is a linear
subspace, PK is just the so-called orthogonal projection onto K. In Chapter 6 the
bounded linear functionals on an inner product space are studied. Representation
theorems for such functionals are given, and these are applied to deduce detailed
results in approximation by hyperplanes or half-spaces. In Chapter 7 a general
duality theorem for the error of approximation is given. A new elementary proof
of the Weierstrass approximation theorem is established, and explicit formulas are
obtained for the distance from any monomial to a polynomial subspace. These are
then used to establish an elegant approximation theorem of Muntz.
In Chapter 8 we seek solutions to the operator equation Ax = b, where A is a
bounded linear operator from the inner product space X to the inner product space
Y, and bEY. If X and Yare Euclidean n-space and Euclidean m-space respec-
tively, then this operator equation reduces to m linear equations in n unknowns.
To include the possible situation when no solution x exists, we reformulate the
problem as follows: Minimize II Ax - bll over all x in X. This study gives rise to the
notions of "generalized solutions" of equations, "generalized inverses" of operators,
etc.
The theory of Dykstra's cyclic projections algorithm is developed in Chapter
9. This algorithm is an iterative scheme for computing best approximations from
the intersection K = nl' K i , where each of the sets Ki is a closed convex set, in
terms of computing best approximations from the individual sets K i . When all
the Ki are linear subspaces, this algorithm is called von Neumann's method of
alternating projections. It is now known that there are at least fifteen different
areas of mathematics for which this algorithm has proven useful. They include
linear equations and linear inequalities, linear prediction theory, linear regression,
computed tomography, and image restoration. The chapter concludes with several
representative applications of the algorithm.
In Chapter 10 we consider the problem of best approximation from a set of
the type K = C n A-1(b), where C is a closed convex set, A is a bounded linear
mapping from X to Y, and bEY. This problem includes as a special case the
general shape-preserving interpolation problem, which is another one of the more
important applications of best approximation theory. It is shown that one can
always replace the problem of determining best approximations to x from K by
the (generally easier) problem of determining best approximations to a certain
perturbation of x from the set C (or from a specified extremal subset of C).
In Chapter 11 the general problem of (finite) interpolation is considered. Also
studied are the problems of simultaneous approximation and interpolation, simul-
taneous interpolation and norm-preservation, simultaneous approximation, inter-
polation, and norm-preservation, and the relationship among these properties. The
last chapter (Chapter 12) examines the question of whether every Chebyshev set
in a Hilbert space must be convex.
Throughout each chapter, examples and applications have been interspersed with
the theoretical results. At the end of each chapter there are numerous exercises.
They vary in difficulty from the almost trivial to the rather challenging. I have also
occasionally given as exercises certain results that are proved in later chapters. My
purpose in doing this is to allow the students to discover the proofs of some of these
important results for themselves, since for each of these exercises all the necessary
machinery is already at hand. Following each set of exercises there is a section
x PREFACE

called "Historical Notes," in which I have tried to put the results of that chapter
into a historical perspective. The absence of a citation for a particular result should
not, however, be interpreted as a claim that the result is new or my own. While I
believe that some of the material included in this book is new, it was not always
possible for me to determine just who proved what or when.
Much of the material of the book has not yet appeared in book form, for example,
most of Chapters 9-12. Surprisingly, I have not seen even some of the more basic
material on inner product spaces in any of the multitude of books on Hilbert space
theory. As an example, the well known Frechet-Riesz representation theorem states
that every bounded linear functional on a Hilbert space has a "representer" in the
space (Theorem 6.10), and this result is included in just about every book on
Hilbert space theory. In contrast to this, not every bounded linear functional on
an (incomplete) inner product space has such a representation. Nevertheless, one
can specify exactly which functionals do have such representations (Theorem 6.12),
and this condition is especially useful in approximation by hyperplanes (Theorem
6.17) or half-spaces (Theorem 6.31).
If we had worked only in Hilbert spaces, parts of the book could have been
shortened and simplified. The lengthy Chapter 6, for example, could have been
considerably abbreviated. Thus we feel an obligation to explain why we spent the
extra time, effort, and pages to develop the theory in arbitrary inner product spaces.
In many applications of the theory of best approximation the natural space to work
in is a space of continuous or piecewise continuous functions defined on an interval,
and endowed with the L 2 -norm (i.e., the norm of a function is the square root of
the integral of the square of the function). Such spaces are inner product spaces
that are not Hilbert spaces. Moreover, for a variety of reasons, it is not always
satisfactory to work in the Hilbert space completion of the given inner product
space. For example, a certain physical problem might demand that a solution, if
any exists, be a piecewise continuous function. Thus it seemed important to me to
develop the theory of best approximation in any inner product space and not just
in a Hilbert space.
In what way does this book differ from the collection of books available today on
approximation theory? In most of these books, if any best approximation theory
in a Hilbert space setting was included, the sum total rarely amounted to more
than one or two chapters. Here we have attempted to produce the first systematic
study of best approximation theory in inner product spaces, and without elaborate
prerequisites. While the choice of topics that one includes in a book is always a sub-
jective matter, we have tried to include a fairly complete study of the fundamental
questions concerning best approximation in inner product spaces.
Each of the first six chapters of the book depends on the material of the chapter
preceding it. However (with the exception of Chapters 9 and 10, each of which
depends upon a few basic facts concerning adjoint operators given in Chapter 8,
specifically, Theorems 8.25-8.33), each of the remaining six chapters (7-12) depends
only on the first six. In particular, Chapters 7-12 are essentially independent of
one another, and anyone of these may be studied after one has digested the first
six chapters. The reader already familiar with the basic fundamentals of Hilbert
space theory can skim the first six chapters to obtain the relevant notation and
terminology, and then delve immediately into any of the remaining chapters. My
own experience with teaching courses based on this book has shown that in a one-
semester course it is possible to cover essentially all of the first seven chapters and
PREFACE xi

at least two or three of the remaining ones. One way of reducing the amount of
time needed to cover the material would be to concentrate on just the Hilbert space
case, and to omit the results that require more involved technicalities necessitated
by the possible incompleteness of the space. Also, for those students whose back-
ground already includes an elementary course in Hilbert space theory, the instructor
should be able to cover the whole book in one semester. Actually, one can even
regard this book as an introduction to the theory of inner product spaces developed
simultaneously with an important application in mind.
Finally, for the sake of completeness and in deference to the experts, we have
included some examples and exercises that do require general measure and integra-

are clearly marked with a *


tion theory (which are not prerequisites for this book). Such examples and exercises
and can be skipped without any loss of continuity.

7 12

CHAPTERINTERDEPENDENCY*

*A chapter is dependent on every chapter to which it is


connected by a sequence of rising lines.
xii PREFACE

Acknowledgments
I have been fortunate to have had the sound advice of many of my students
over the years that I have taught the topics in this book. They have helped me
to see how to present the material in a way that is most beneficial to a student's
understanding, and they forced me to clarify some arguments that were originally
somewhat opaque.
I am grateful to the following people, each of whom read certain parts of earlier
drafts of the book and offered valuable constructive cr.iticisms: Heinz Bauschke,
Ward Cheney, Charles Groetsch, Hein Hundal, Vic Klee, Wu Li, Daniel Murphy,
Adrian Ocneanu, Bob Phelps, Allan Pinkus, Ivan Singer, Don Salmon, Leonid
Vlasov, Joe Ward, and Isao Yamada. Moreover, Heinz Bauschke read the en,tire
manuscript, offered many useful critical comments, and corrected numerous typos.
In short, he was an editor's dream.
I am also indebted to Bob Seeds and Frances Weller of the Penn State Mathe-
matics Library for their assistance in so many ways related to reference materials.
In addition, I am grateful to Doug Arnold who bailed this TeX novice out of some
serious difficulties with his "TeXpertise."
Finally, I owe a large debt of gratitude to Jun (Scott) Zhong, who, as a graduate
student about ten years ago, typed the first drafts of several of these chapters, and
to Kathy Wyland, who has handled the bulk of the typing chores over the last
several years with her usual good speed and accuracy.

University Park, Pennsylvania, USA FRANK DEUTSCH


xiii

CONTENTS

Preface vii
Acknowledgments xii

Chapter 1. Inner Product Spaces 1


Five Basic Problems ....................................... 1
Inner Product Spaces ...................................... 2
Orthogonality .............................................. 8
Topological Notions ........................................ 10
Hilbert Space .............................................. 14
Exercises. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
Historical Notes ............................................ 19

Chapter 2. Best Approximation 21


Best Approximation ........................................ 21
Convex Sets ................................................ 22
Five Basic Problems Revisited. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Exercises .............................. . . . . . . . . . . . . . . . . . . . . . 30
Historical Notes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

Chapter 3. Existence and Uniqueness of


Best Approximations 33
Existence of Best Approximations .......................... 33
Uniqueness of Best Approximations ........................ 35
Compactness Concepts ..................................... 38
Exercises. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . 39
Historical Notes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . 40

Chapter 4. Characterization of Best Approximations 43


Characterizing Best Approximations ........................ 43
Dual Cones ................................................ 44
Characterizing Best Approximations from Subspaces ........ 50
Gram-Schmidt Orthonormalization ......................... 51
Fourier Analysis ............................................ 54
Solutions to the First Three Basic Problems ................ 61
Exercises ................................................... 64
Historical Notes ............................................ 69

Chapter 5. The Metric Projection 71


Metric Projections onto Convex Sets ....................... 71
Linear Metric Projections .................................. 77
xiv CONTENTS

The Reduction Principle ................................... 80


Exercises ................................................... 84
Historical Notes ............................................ 87

Chapter 6. Bounded Linear Functionals and Best


Approximation from Hyperplanes and Half-Spaces 89
Bounded Linear Functionals ................................ 89
Representation of Bounded Linear Functionals .............. 93
Best Approximation from Hyperplanes ..................... 97
Strong Separation Theorem ................................ 102
Best Approximation from Half-Spaces ...................... 107
Best Approximation from Polyhedra ........................ 109
Exercises ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 7
Historical Notes ............................................ 122

Chapter 7. Error of Approximation 125


Distance to Convex Sets .................................... 125
Distance to Finite-Dimensional Subspaces .................. 129
Finite-Codimensional Subspaces ............................ 133
The Weierstrass Approximation Theorem ................... 139
Muntz's Theorem .......................................... 143
Exercises. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
Historical Notes ............................................ 151

Chapter 8. Generalized Solutions of


Linear Equations 155
Linear Operator Equations. ... .. ...... . . . . . .. . . . . .. . . . . . . . . 155
The Uniform Boundedness and Open Mapping Theorems... 164
The Closed Range and Bounded Inverse Theorems.......... 168
The Closed Graph Theorem................................ 169
Adjoint of a Linear Operator ............................... 171
Generalized Solutions to Operator Equations ............... 177
Generalized Inverse ........................................ 179
Exercises ................................... . . . . . . . . . . . . . . . . 187
Historical Notes ............................................ 191

Chapter 9. The Method of Alternating Projections 193


The Case of Two Subspaces ................................ 193
Angle Between Two Subspaces ............................. 197
Rate of Convergence for Alternating Projections
(two subspaces) ........................................ 201
Weak Convergence ......................................... 203
Dykstra's Algorithm ....................................... 207
The Case of Affine Sets .................................... 215
Rate of Convergence for Alternating Projections. . .. . . . . ... . 217
Examples .................................................. 226
Exercises ................................................... 230
Historical Notes ............................................ 233
CONTENTS xv

Chapter 10. Constrained Interpolation frOIn a


Convex Set 237
Shape-Preserving Interpolation ............................. 237
Strong Conical Hull Intersection Property (Strong CHIP) ... 238
Affine Sets ................................................. 247
Relative Interiors and a Separation Theorem ................ 251
Extremal Subsets of C ..................................... 261
Constrained Interpolation by Positive Functions ............ 270
Exercises ................................................... 277
Historical Notes ............................................ 283

Chapter 11. Interpolation and ApproxiInation 287


Interpolation ............................................... 287
Simultaneous Approximation and Interpolation ............. 292
Simultaneous Approximation, Interpolation, and
Norm-preservation ..................................... 294
Exercises ................................................... 295
Historical Notes ............................................ 298

Chapter 12. Convexity of Chebyshev Sets 301


Is Every Chebyshev Set Convex? ........................... 301
Chebyshev Suns ............................................ 302
Convexity of Boundedly Compact Chebyshev Sets .......... 304
Exercises ................................................... 306
Historical Notes ............................................ 307

Appendix 1. Zorn's LeInIna 311

Appendix 2. Every Hilbert Space Is £2(I) 312

References 315

Index 331
CHAPTER 1

INNER PRODUCT SPACES

Five Basic Problems


To motivate the subject matter of this book, we begin this chapter by listing
five basic problems that arise in various applications of "least-squares" approxi-
mation. While these problems seem to be quite different on the surface, we will
later see that the first three (respectively the fourth and fifth) are special cases of
the general problem of best approximation in an inner product space by elements of
a finite-dimensional subspace (respectively convex set). In this latter formulation,
the problem has a rather simple geometric interpretation: A certain vector must
be orthogonal to the linear subspace.
The remainder of the chapter is devoted to defining the main spaces of interest:
the inner product spaces and Hilbert spaces, giving some examples of these spaces,
and recording a few of their elementary, but useful, properties.
Problem 1. (Best least-squares polynomial approximation to data) Let
I j = 1,2, ... , m} be a table of data (i.e., the graph of a real function
{( t j , x(tj))
x defined on the ti's). For any fixed integer n < m, find a polynomial p(t) =
2:7=0 ai ti , of degree at most n, so that the expression
m

2)X(tk) - P(tkW
k=l

is minimized.
Problem 2. (Solution to an over-determined system of equations) Consider the
linear system of m equations in the n unknowns Xl, X2, ... , Xn:
allXI + al2X2 + ... + alnXn = bl
a2lxI + a22x2 + ... + a2nXn = b2

In matrix-vector notation, this can be written as Ax = b. In the absence of further


restrictions on A or b, this system may fail to have a solution x. The next best
thing we can ask for is to make the residual vector r := Ax - b "small" in some
sense. This suggests the following problem: Find a vector x = (XI,X2,'" ,xn ) that
minimizes the expression

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
2 INNER PRODUCT SPACES

Problem 3. (Best least-squares polynomial approximation to a function) Let


x be a real continuous function on the interval (a, b]. Find a polynomial p(t)
I:~=o aiti, of degree at most n, such that the expression

J
b

(x(t) - p(t)fdt
a

is minimized.
Problem 4. (A control problem) The position 8 of the shaft of a dc motor
driven by a variable current source u is governed by the differential equation

(4.1) 81/(t) + 8'(t) = u(t), 8(0) = 8' (0) = 0,


where u(t) is the field current at time t. Suppose that the boundary conditions are
given by

(4.2) 8(1) = 1, 8'(1) = 0,

and the energy is proportional to J~ u 2 (t)dt. Find the function u having minimum
energy in the class of all real continuous functions on (0, 1] for which the system
(4.1) and (4.2) has a solution 8.
Problem 5. (Positive constrained interpolation) Let {Xl, X2, ... , x n } be a fi-
nite set of real square-integrable functions on (0,1] and let {b l ,b2 , ... ,bn } be real
numbers. In the class of all square-integrable functions X on (0,1] such that

x(t) ~ 0 for t E (0,1]

11
and
x(t)xi(t)dt = bi (i = 1,2, ... , n),

find the one for which

is minimized.

Inner Product Spaces


The natural setting for these and similar problems is a real inner product space.
Let X be a real linear space, that is, a linear space over the field R of real numbers.
1.1 Definition. A real linear space X is called an inner product space if
for each pair of elements x, y in X there is defined a real scalar (x, y) having the
following properties (for every x, y, z in X and a E R):
(1) (x, x) ~ 0,
(2) (x,x) = 0 if and only if x = 0,
(3) (x,y) = (y,x),
(4) (ax,y) =a(x,y),
(5) (x + y, z) = (x, z) + (y, z).
INNER PRODUCT SPACES 3

The function (-,.) on X x X is called an inner product (or scalar product) on


X. Properties (1) and (2) state that the inner product is "positive definite," and
property (3) says that it is "symmetric."
From (4), (5), and induction, we obtain that

(1.1.1)

From (3) and (Ll.l) we deduce

(1.1.2)

Properties (1.1.1) and (1.1.2) state that the inner product is "bilinear;" i.e., linear
in each of its arguments. An immediate consequence of (4) is

(1.1.3) <0, y) = °for every y E X.

Also, from property (2), we immediately deduce

(1.1.4) (x, y) = 0 for every x EX implies that y = o.


Inner product spaces are sometimes called pre-Hilbert spaces.
Unless explicitly stated otherwise, we shall henceforth assume throughout
the entire book that X is a real inner product space!
Before giving examples of inner product spaces, it is convenient to state a fun-
damental inequality. For each x E X, the norm of x is defined by

Ilxll := v(x, x) .

Theorem 1.2 (Schwarz Inequality). For any pair X,y in X,

(1.2.1) I (x, y) I ~ Ilx1111Y11 .

Moreover, equality holds in this inequality if and only if x and y are linearly de-
pendent.
Proof. If x and yare linearly dependent, say x = ay for some scalar a, then
both sides of (1.2.1) equal la111Y112. If x and yare linearly independent, then for
every scalar A we see that x - AY # 0, so that

In particular, setting A = (x, y) Ilyll-2 yields

and hence strict inequality holds in (1.2.1). •


4 INNER PRODUCT SPACES

1.3 Exrunples of Inner Product Spaces. All of the examples of linear spaces
that are considered below are spaces of real functions x defined on a certain do-
main. The operations of addition and scalar multiplication of functions are defined
pointwise:
(x + y)(t) := x(t) + y(t), (ax)(t):= ax(t).
Let T be any set and let F(T) denote the set of all real functions defined on
T. It is easy to verify that F(T) is a linear space with zero element the function
identically O. Thus, in each example below, to verify that X is a linear space is
equivalent to verifying that X is a (linear) subspace of F(T) (for some T). And for
this it suffices to verify that x + y and ax are in X whenever x and y are in X and
aER
(1) For any positive integer n, l2(n) will denote the functions x : {I, 2, ... , n} -t
]R with the inner product
n
(x,y) = Lx(i)y(i).
1

The inner product space l2(n) is called Euclidean n-space. The norm is given by
1

Ilxll = [~IX(i)12]"
Observe that there is a one-to-one correspondence between each x E l2(n) and
its ordered array of values

(x(1),x(2), ... ,x(n)) E ]Rn.

Thus we may regard the space l2(n) and]Rn as one and the same. This identification
of l2(n) with ]Rn will aid our geometric intuition. (We can "draw pictures" in]Rn
for n = 1,2, or 3!) More often than not, this geometric intuition in ]Rn will enable
us to conjecture results that are true in any inner product space.
(2) l2 will denote the space of all real-valued functions x defined on the natural
numbers N := {I, 2, ... , n, . .. } with the property that L~ x 2(i) < 00. If we define
00

(1.3.1) (x,y) = Lx(i)y(i),

then l2 is an inner product space. To see this, first note that by the Schwarz
inequality in b(n), we have, for each X,y in b,

for all n. Hence, by passing to the limit as n tends to infinity,


00

Llx(i)y(i)1 ::; c < 00,


INNER PRODUCT SPACES 5

so L~ x(i)y(i) converges absolutely, and (1.3.1) is well-defined. The remaining


inner product properties are easily verified.
To verify that l2 is actually a linear space, it suffices to show that x + y and
ax are in l2 whenever x and yare in b and a E R Clearly, ax E l2 if x E b. If
x, Y E b, then for every n E N, again using Schwarz's inequality in l2(n),

n n n n
~:=rx(i) + y(i)j2 = l:)2(i) + 2Lx(i)y(i) + Ly2(i)
1 1 1 1
1 1

S ~)2(i) + 2 [~X2(i)]" [~y2(i)]" + ~y2(i)


S ~X2(i) + 2 [~X2(i)]! [~y2(i)]! + ~y2(i) =: c.

Thus
00

L[x(i) + y(iW S c < 00,


1

and so x + Y E l2. Thus l2 is an inner product space. The norm in l2 is given by

(3) Let C2 [a, bJ denote the space of all continuous real functions x on the finite
interval [a, b], and define

(x, y) = lb x(t)y(t)dt.

It is easy to verify that C 2 [a, bJ is an inner product space. The norm is given by

(4) We consider a more general inner product space than that of example (3).
Let I be any closed interval of the real line R Thus I is one of the four intervals
[a, b], (-00, b], [a, 00), or lR = (-00,00). Let w be any function that is positive and
continuous on the interior of I and is integrable:

1 w(t)dt < 00.

Let C 2 (1; w) denote the set of all bounded continuous real functions on I with an
inner product defined by

(x, y) = 1 w(t)x(t)y(t) dt.


6 INNER PRODUCT SPACES

In the same way one can show that C2 (1; w) is an inner product space. The norm
is thus given by

[1 w(t)x (t) dt]:2


1

Ilxll = 2

Note that in the special case where w(t) == 1 and 1 = [a, b], C2 (1; w) is just the
space we denoted by C2 [a, bj in (3).
* (5) A somewhat larger space than C2 [a, bj is the space of all (Lebesgue) mea-
surable real functions x on [a, bj with the property that (the Lebesgue integral)
J: x 2 (t)dt < 00. If we agree to identify two functions that are equal almost every-
where, then the expression

(x, y) = lb x(t)y(t) dt

defines an inner product, and the resulting inner product space is denoted by
L 2 [a,bj. The norm is the same as in C2 [a,bj.
(6) The present example contains the first and second as special cases. Let
1 denote any index set and consider the space [2(1) of all functions x : 1 -+ lR
with the property that {i E 1 I x(i) i= O} is a countable set J = {jl,j2, ... } and
Lk X 2 (jk) < 00. Since this sum converges absolutely, every rearrangement of the
terms yields the same sum. Thus we can unambiguously define

Lx2 (j) := Lx 2 (j) = LX 2 (jk).


jEI jEJ k

The same argument that shows that lz is a linear space also shows that [2 (1) is a
linear space. Similarly, the expression

(x, y) = Lx(i)y(i)
iEI

can be shown to be an (unambiguously defined) inner product on [2(1). Thus [2(1)


is an inner product space. The norm in [2(1) is given by
1

Ilxll = [LX 2 (i)] "


iEI

Note that when 1 = {1,2, ... ,n} (respectively 1 = N = {1,2, ... }), then [2(1) is
just the space we earlier denoted by [2(n) (respectively lz). (An alternative approach
to defining the space [2(1) is given in Exercise 7 at the end of this chapter.)
* (7) This example contains the first, second, fifth, and sixth as special cases. Let
(T, S, /1-) be a measure space. That is, T is a set, S is a O"-algebra of subsets of T,
and /1- is a measure on S. Let L 2 (/1-) = L 2 (T, s, /1-) denote the set of all measurable
real functions x on T such that
INNER PRODUCT SPACES 7

(Here, as in example 5, we identify functions that are equal almost everywhere.)


Then L2(p,) is a linear space. If we define

(x,y):= h. xydp, for X,y E L 2(p,),

then L2 (p,) is an inner product space.


If T = I is a closed interval in lR and p, is Lebesgue measure on T, we often
denote L2(p,) by L2(1). Note that if T is any set, S consists of all subsets of T,
and p, is "counting measure" on T (i.e., p,(E) is the number of elements in E if E
is finite, and p,(E) = CXJ otherwise), then L 2(T,S,p,) = l2(T).

1.4 Properties of the Norm. The norm function x t-t Ilxll := V(x,x) has
the following properties:

°
(1) IIxll 2 0,
(2) IIxll = if and only if x
(3) II ax I = lalllxll,
= 0,

(4) IIx + yll :::; IIxll + lIyll ("triangle inequality").


Moreover, equality holds in the triangle inequality if and only if x or y is a
nonnegative multiple of the other.
Proof. Properties (1), (2), and (3) follow immediately from the properties of
the inner product. To prove (4), we use the Schwarz inequality to obtain

IIx + Yll2 = (x + y, x + y) = IIxll 2 + 2 (x, y) + lIyll2


:::; IIxll 2 + 21 (x, y) I + lIyll2
:::; IIxll 2 + IIxlillyll + lIyll2 = (lIxll + lIyll)2,

which implies (4).


From the proof of (4), we see that equality holds in the triangle inequality if and
only if

(1.4.1) (x,Yl = I (x,Yll = IIxIiIlYII·

Using the condition of equality in Schwarz's inequality, it follows that the second
equality in (1.4.1) holds if and only if one of x or Y is a scalar multiple of the other.
Then the first equality in (1.4.1) shows that this scalar must be nonnegative. •
We note that the norm function is a generalization of the "absolute value" func-
tion on the space of real numbers K More generally, the norm of the vector x - Y
in Euclidean 2- or 3-space is just the usual "Euclidean distance" between x and y.
This suggests that we define the distance between any two vectors x and y in an
inner product space X by IIx - ylI, and the length of x by IIxli.
Any linear space that has a function x t-t IIxll defined on it satisfying the prop-
erties (1)~(4) of Proposition 1.4 is called a norITled linear space. Thus an inner
product space is a particular kind of normed linear space. What differentiates an
inner product space from an arbitrary normed linear space is the following "paral-
lelogram law." In fact, a normed linear space is an inner product space if and only
if the parallelogram law holds (see Exercise 13 at the end of the chapter).
8 INNER PRODUCT SPACES

1.5 Parallelogram Law. For every x, y in an inner product space,

Proof. We have

Ilx ± Yl12 = (x ± y, x ± y) = IIxl1 2± 2 (x, y) + IIyl12 .


Adding these two equations yields the result. •
The geometric interpretation of the parallelogram law is that in a parallelogram,
the sum of the squares of the lengths of the diagonals equals the sum of the squares
of the lengths of all four sides (see Figure 1.5.1).

IIxll .........• x+y

lIyll

O~--------------------~~x
II x II

Figure 1.5.1. Parallelogram. law

Orthogonality
It is a well-known elementary geometric fact in 1~2 (or ]R3) that two vectors x
and yare perpendicular if and only if L:i x(i)y(i) = 0 (or L:i x(i)y(i) = 0). That
is, in ]R2 or :lR3 , two vectors x and yare perpendicular if and only if (x, y) = o.
Motivated by this, we make the following definition in any inner product space.
1.6 Definition. Two vectors x and yare said to be orthogonal, written x 1.. y,
if (x, y) = O. The vector x is said to be orthogonal to the set Y, written x 1.. Y,
if x 1.. y for every y E Y. A set A is called an orthogonal set if each pair of
distinct vectors in A are orthogonal. An orthonormal set is an orthogonal set in
which each vector has norm one. Thus {Xi liE J} is an orthonormal set if and
only if
ifi f j,
if i = j.

1. 7 Orthogonal Sets are Linearly Independent. Let A be an orthogonal


set of nonzero vectors. Then A is a linearly independent set.

t
Proof. Let {Xl, X2, ... , x n } C A and suppose L:~ Ctixi = O. Then for each k,

0= (t CtiXi, Xk) = Cti(Xi, Xk) = Ctk(Xk, Xk) = Ctkllxkl12


ORTHOGONALITY 9

implies Qk = 0, since Xk =I 0. This proves that A is linearly independent. •

The converse of Proposition 1.7 is false: There are linearly independent sets A
that are not orthogonal (e.g., A = {(I, 0), (1, I)} in 12(2)). However, in Proposition
4.11 a constructive procedure is presented (the Gram-Schmidt process) for obtaining
an orthonormal set B from A with the property that span B = span A. (Here span
A denotes the subspace spanned by A.)

1.8 Examples of Orthonormal Sets.


(1) In 12 (I), the set {ei liE I}, where ei (j) = i5ij for all i, j in I, is orthonormal.
In particular, in b(n), the vectors el = (1,0, ... ,0), e2 = (0,1,0, ... ,0), ... , en =
(0, ... ,0,1) form an orthonormal set.
(2) In C2[-1T,1Tj (or L 2[-1T,1T]), the set offunctions

III. 1 1. }
{ J2;T' Vi cost, Vi smt, ... , Vi cos nt, Vi smnt, ...

is orthonormal. This follows from the relations

J J
~ ~

cos nt cos mt dt = sin nt sin mt dt = 1Ti5nm

and

J J J
~ ~ ~

cosntdt = sinntdt = sinmtcosntdt = 0.


-~ -~

Just as in Euclidean 2-space, the Pythagorean theorem has a valid analogue in any
inner product space (see Figure 1.9.1).

1.9 Pythagorean Theorem. Let {Xl, X2,· .. , xn} be an orthogonal set. Then

In particular, if X J.. y, then

Proof.
10 INNER PRODUCT SPACES

y
X+y

lIy II

x
II X II
Figure 1.9.1. Pythagorean theorem
Again motivated by the situation in Euclidean 2 and 3-space, we can define the
angle between two vectors in an inner product space as follows.
1.10 Definition. For any two nonzero vectors x and y, the angle 0 = 8(x, y)
between them is defined by

(x, y)
8 = arccosMllYIT .

This expression is well-defined, since the Schwarz inequality implies

I (x, y) I
WM::; 1,

and moreover, 0 ::; 8 ::; 1r. Observe that x and y make an angle of at least (respec-
tively at most) 'If /2 if and only if (x, y) ::; 0 (respectively (x, y) :::: 0). Also x .l y
if and only if 8(x, y) = ~. Further, O(x, y) = 0 (respectively O(x, y) = 'If) if and
only if one of x or y is a positive (respectively negative) multiple of the other. (See
Exercise 11 at the end of the chapter.)

Topological Notions

1.11 Definition. A sequence of vectors {x n } in X is said to converge to x E X


and x is called the limit of {x n }, denoted by Xn --+ x or x = limxn, provided that
limn->oo Ilx n - xii = O. That is, for each E > 0, there exists an integer N such that

Ilx n - xii < E for all n:::: N.

The next result shows, essentially, that the norm, inner product, addition, and
scalar multiplication are all continuous functions.
1.12 Theorem. (1) For all x, y,

Illxll -llylll ::; Ilx - yll .


TOPOLOGICAL NOTIONS 11

(2) If xn -t X, Yn -t y, and an -t a, then:


(a) Ilxnl\ -t Ilxll,
(b) (xn,Yn) -t (x,y),
(c) xn+yn-tx+y,
(d) anX n -t ax.

Proof. (1) Now,

Ilxl\ = II(x - y) + yl\ :::; Ilx - yl\ + l\yll


implies
I\xl\-llyl\:::; Ilx-yl\.
Interchanging the roles of x and y yields

I\yl\- Ilxll :::; Ily - xii = Ilx - yll .


These two inequalities combine to yield (1).
(2) (a) This follows immediately from (1).
(b) Using Schwarz's inequality, we obtain
I (xn, Yn) - (x, y) I = I (xn - x, Yn) + (x, Yn - y) I
:::; I (xn - x, Yn) I + I (x, Yn - y) I
:::; Ilxn - xl\llYnl\ + Ilxl\l\Yn - yl\ -t O.
(c) Ilxn + Yn - (x + y)1\ :::; Ilxn - xii + llYn - yll-t O.
(d)
lIanxn - axl\ = I\an(xn - x) + (an - a)xl\
:::; lanlllxn - xii + Ian - all\xll -t O. •

A set K is said to be closed if x E K whenever {x n } C K and Xn -t x. That


is, K contains the limits of all its convergent sequences. K is called open if its
complement X \ K := {x E X I x ¢:. K} is closed. There are two sets that arise so
frequently that they warrant a special notation. The closed ball and open ball
centered at x with radius r are defined by
B[x,r]:= {y E X Il\x-yl\:::; r} and
B(x, r) := {y E X Ilix - yll < r}), respectively.
The justification for these names is included in the next lemma, which also records
some elementary topological properties of open and closed sets.
1.13 Lemma. (1) The empty set 0 and full space X are both open and closed.
(2) The intersection of any collection of closed sets is 'closed.
(3) The union of a finite number of closed sets is closed.
(4) Any finite set is closed.
(5) The union of any collection of open sets is open.
(6) The intersection of a finite number of open sets is open.
(7) The set U is open if and only if for each x E U, there exists an r: > 0 such
that B(x, r:) C U.
(8) Any open ball B(x, r) is open.
(9) Any closed ball B[x, r] is closed.
12 INNER PRODUCT SPACES

Proof. Statements (1)-(6) are easy consequences of the definitions and DeMor-
gan's laws. They are left as an exercise.
(7) Let U be open and x E U. If B(x, E) \ U i= 0 for each E > 0, then for each
integer n, we can choose Xn E B(x, lin) \ U. Then Xn --7 x and x E X \ U, since
X \ U is closed. This contradiction shows that B(x, E) C U for some E > 0.
Conversely, if U is not open, then K = X \ U is not closed. Thus there exist a
sequence {xn} in K and an x 1'. K such that Xn --7 x. Then Xn 1'. U for all n, so
that for each E > 0, Xn E B(x, E) \ U for n sufficiently large. Hence B(x, E) \ U i= 0,
i.e., B(x, E) rt. u, for every E> 0. This proves (7).
(8) Let C = X \ B(x, r). Let Xn E C and Xn --7 Xo. Then Ilx n - xii 2: r for all
n, and by Theorem 1.12 (2), r -<; lim Ilx n - xii = Ilxo - xii, so Xo E C. Thus C is
closed and B(x, r) is open.
The proof of (9) is similar to that of (8). •
Next we give a procedure for obtaining special closed and open sets generated
by a given set.
1.14 Definition. The closure of a set A, denoted by A, is the intersection of
all closed sets that contain A. Similarly, the interior of A, denoted by int A, is the
union of all open sets that are contained in A.
Some elementary properties of the closure and interior operations, which follow
easily from Lemma 1.13, are listed next.
1.15 Lemma. (1) A is the smallest closed set that contains A.
(2) A is closed if and only if A = A.
(3) x E A if and only if B(x, E) n A i= 0 for each E > 0.
(4) x E A if and only if there exists a sequence {x n } in A such that Xn --7 x.
(5) int A is the largest open set contained in A.
(6) A is open if and only if A = int A.
°
(7) x E int A if and only if there exists E > such that B(x, E) cA.
(8) B[x, r) = B(x, r).

Proof. We prove only statement (3) and leave the remaining statements as an
easy exercise.
Let x E A. If B(x, E) n A = 0 for some E > 0, set F = X \ B(x, E). Then F is
closed and F :) A. Hence F :) A by (1), which implies that An B(x, E) = 0. But
this contradicts x E An B(x, E). Hence B(x, E) n A i= 0 for every E > 0.

°
Conversely, if x 1'. A, then x is in the open set U := X \ A, so there exists an
E> such that B(x, E) C U. Therefore B(x, E) n A = 0, and so B(x, E) n A = 0. •
1.16 Definition. A sequence {xn} is called a Cauchy sequence if for each
E > 0, there exists an integer N such that

Ilx n - xmll < E for all n, m 2: N.

A nonempty set B is called bounded if sup{llxlll x E B} < 00.

1.17 Lemma. (1) Every convergent sequence is a Cauchy sequence.


(2) Every Cauchy sequence is bounded.
(3) The space C2 [a, b) contains a Cauchy sequence that does not converge.
TOPOLOGICAL NOTIONS 13

Proof. (1) If Xn -+ x, then

Ilxn - xmll ::; Ilxn - xii + Ilx - xmll -+ 0

as n,m -+ 00. Thus {xn} is Cauchy.


(2) Let {xn} be Cauchy. Then there exists N such that Ilx n - xmll ::; 1 for all
n, m :::: N. In particular,

for all n:::: N. Hence sUPn Ilxnll ::; max{llx111, Ilx211, ... , IlxN-IiI, 1 + IlxNII} < 00.
(3) Without loss of generality, we may assume [a, bJ = [-l,lJ. Consider the
functions Xn E C2[-1, 1J defined by

if - 1 ::; t ::; 0,
. 1
1fO::;t::;-,
n
if.!.<t<1.
n - -

(See Figure 1.17.1.)

-1 o lin

Figure 1.17.1

It is easy to verify that {xn} is a Cauchy sequence. Suppose there did exist some
x E C2 [-1, 1J such that Xn -+ x. Then, since

Ilxn - xI12 = [11 [Xn(t) - x(tWdt

= 1°-1
x2(t)dt + r n[nt - x(tWdt + r [1- x(tWdt
~
1
/
1

11/n

and Ilx n -xii -+ 0, it follows that x(t) = 0 for -1 ::; t ::; 0 and x(t) = 1 for 0 < t ::; 1.
But this contradicts the continuity of x. Hence {xn} does not converge. •
14 INNER PRODUCT SPACES

Hilbert Space
Those inner product spaces in which every Cauchy sequence converges have many
pleasant properties. Some important theorems can be proved in such spaces that
are not true in general inner product spaces. They therefore warrant a special
name.
1.18 Definition. An inner product space X is called complete, or a Hilbert
space, if each Cauchy sequence in X converges to a point in X.
1.19 Examples. (1) The space R = l2(1) is complete. This is just the statement
that every Cauchy sequence of real numbers converges to a real number. This, in
turn, is equivalent to the completeness axiom of the real numbers.
(2) The space l2(1) is a Hilbert space for any index set 1. In particular, l2(n)
and l2 are both Hilbert spaces.
To see this, let {xn} be a Cauchy sequence in l2(1). Then each Xn vanishes
except on the countable set

An := {i E 1! xn(i) of. O} .
Let A = Uj'" An, so A is also countable. Given E > 0, choose an integer N such that
!!Xn - xm!! < E for all n, m ~ N.

In particular, for all n, m 2': N,

!xn(i) - xm(i)! < E for all i EA.

That is, {xn(i) ! n = 1,2, ... } is a Cauchy sequence of real numbers for each i E A.
Since R is complete, for each i E A there exists x(i) E R such that xn(i) -7 x(i).
Defining x(i) = 0 for all i E 1 \ A, we obtain a function x : 1 -7 R. We will show
that x E l2(1) and Xn -7 x. We first observe the obvious fact that xn(i) -7 x(i) for
each i E 1. Also,
2)Xn(i) - x m (i)!2 = !!xn - xm!!2 < E2
iEA

for all n, m ~ N. In particular, for each finite subset J of A,

I)Xn(i) - xm(iW < E2


iEJ

for all n, m ~ N. Letting m -7 00 in this inequality, we obtain

I)Xn(i) - x(iW <::: E2 for all n ~ N.


iEJ

Since this is true for every finite subset J c A, it follows that


(1.19.1) !!Xn - x!!2 = 2.)xn(i) - x(iW <::: E2
iEA

for all n ~ N. This shows that for n ~ N, x - Xn E l2(1). Since l2(1) is a linear
space and Xn E b(1), we have that x = (x-x n ) +x n E l2(1). Finally, from (1.19.1)
we see that Xn -7 x.
EXERCISES 15

It can be shown that every Hilbert space can be identified with a space of type
12(1) for some index set I (see Appendix 2). Thus h(1) represents the most general
Hilbert space.
(3) Every finite-dimensional inner product space is complete. (This will follow
as a consequence of stronger results proved in Chapter 3; see Theorem 3.7.)
(4) C 2 [a, b] is not a Hilbert space. (This follows by Lemma 1.17(3).)
(5) More generally, the weighted space C 2 (l;w) (see Example 1.3(4» is not a

**
Hilbert space. (The proof is similar to that of (4).)
(6) L 2 [a, b] is a Hilbert space. This is a special case of example (7) below.
(7) L2 (/-I) is a Hilbert space. The proof of this fact can be found in almost any
book on real analysis.

Exercises

1. Limits are unique. That is, if Xn ~ x and Xn ~ y, then x = y.


2. A sequence {x n } converges to x if and only if for each E > 0, Xn E B(x, E)
for all n sufficiently large.
3. A Cauchy sequence converges if and only if some subsequence converges.
4. Verify statements (1)-(6) of Lemma 1.13.
5. Let {x n } eX, x EX. Show that Xn ~ x if and only if (xn, y) ~ (x, y) for
every y E X and Ilxnll ~ Ilxli.
6. Call a subset A of X complete if each Cauchy sequence in A converges to
a point in A. Verify that:
(a) every complete set is closed;
(b) a subset of a Hilbert space is complete if and only if it is closed.
7. An alternative, but equivalent, way of defining the space 12(1) is as follows.
Let 12 (1) denote the set of all functions x: I ~ lR such that I:iEI x 2 (i) < 00,
where

(a) Show that if x E l2(1), then the support of x,

supp x:= {i E I I xCi) i O},

is countable.
(b) Show that for each x E b(l), if J = {jl,h, ... } = supp x, then the sum
2)2(jk)
k

does not depend on the order of the indices jk in this sum and

2)2(jk) = L:>2(i) .
k iEI

(c) Show that for each x, y E b (I), the expression

Lx(i)y(i) := LX(jk)y(jk)
iEI k
16 INNER PRODUCT SPACES

is a well-defined real number that is independent of the ordering of the


indices jk E J := (supp x) U (supp y).
(d) Show that the expression

(x, y) := Lx(i)y(i)
iEI

defines an inner product on b(J).


8. (a) If {x n } is a Cauchy sequence, then {[[x n [[} converges.
(b) If {x n} is a Cauchy sequence, then {(xn,x)} converges for each x E X.
(c) Give an example in l2 of a sequence {xn} such that {(xn, x)} converges
for each x E l2, but {xn} is not a Cauchy sequence.
9. Let x E X and r > o. (a) Show that the set

A = {y E X [[[x - y[[ :<:; r} n {y E X[[[yli2': r}


is closed.
(b) Show that the set

B = {y E X [ [[x - y[[ > r}


is open.
10. Prove the "cosine law":

[[x - y[[2 = [[x[[2 + [[y[[2 - 2[[xli[[y[[ cos 8(x, y).

Interpret geometrically.
11. Verify that if x and y are nonzero vectors, then 8(x, y) = 0 (respectively
8(x, y) = 7r) if and only if one of them is a positive (respectively negative)
multiple of the other.

1+
12. Estimate
1 tn
--dt
o 1 t
from above by the Schwarz inequality. Compare with the exact answer for
n=6.
13. Let X be a real normed linear space whose norm satisfies the parallelogram
law:

Define

1
(x, y) := "4 [[[x + y[[2 - [[x - y[[2] for x, y E X.

Show that this expression defines an inner product on X such that (x, x) = v'
[[xli. Thus verify that a normed linear space "is" an inner product space
if and only if its norm satisfies the parallelogram law. [Hint: Verifying
additivity (x + y, z) = (x, z) + (y, z) requires the parallelogram law. From
this one gets (nx, y) = n (x, y) for all n E N by induction. Then obtain
EXERCISES 17

(rx, y) = r (X, y) for all rational numbers r. Finally, deduce (ax, y)


a (x, y) for any real number a by taking a sequence of rationals rn -+ a.]
14. (Weighted l2) This example generalizes the space l2(I). Let I be an index
set and w a positive function on I. Let b(1; w) denote the set of all functions
x: 1-+ lR. such that LiEI w(i)x 2 (i) < 00. Then

(x, y) = LW(i)x(i)y(i)
iEI

is a well-defined inner product on l2(1;w), and 12(1;w) is complete.


15. (Weighted C2 [a, bJ) Show that the space C2 (1; w) defined in Example
1.3(4) is an inner product space that is not complete.
16. (Discrete spaces) Let P n denote the set of all polynomials of degree at
most n:

Let T = {h, t2, ... , t m } be a set of m distinct real numbers and define
(16.1) (x, y) = L x(t)y(t).
tET

(a) If m :::: n + 1, then (16.1) defines an inner product on P n . Moreover, Pn


is complete. [Hint: Induction on n.]
(b) If m < n + 1, what property of the inner product does (16.1) fail to
possess?
(c) More generally, if w is a positive function on T and m:::: n + 1, then
(16.2) (x, y) = L w(t)x(t)y(t)
tET
defines an inner product on Pn , and Pn is complete.
17. Complete the proof of Lemma 1.15.
18. Let X be a complex linear space, i.e., a linear space over the field C of
complex numbers. X is called a complex inner product space if there
is a function (-, .; from X x X into C having the properties

°
(1) (x, x) :::: 0,
(2) (x, x) = if and only if x = 0,
(3) (x, y) = (y, x) (where the bar denotes complex conjugation),
(4) (ax,y)=a(x,y),
(5) (x + y, z) = (x, z) + (y, z).
The function (-,.; is called an inner product on X. Note that the only
difference between inner products on real spaces and on complex spaces
is that in the latter case, (x,y) = (y,x) rather than (x,y) = (y,x). (Of
course, if X is a real space, then since the inner product is real-valued,
(x, y) = (y, x) = (y, x) and all the conditions (1)-(5) hold.)
(a) Verify that the inner product is a linear function in the first argument
and conjugate linear in the second; i.e.,

(ax + j3y, z) = a (x, z) + j3 (y, z)


and
(z, ax + j3y) = a(z,x) + 7J (z,y).
18 INNER PRODUCT SPACES

(b) Show that Schwarz's inequality is valid for complex inner product spaces.
(c) Show that the parallelogram law and the Pythagorean theorem are valid
in complex inner product spaces.
(d) Define the angle O( x, y) between nonzero vectors x and y by

Re (x,y))
O(x, y) = arccos ( Ilxll Ilyll .
Show that the cosine law (Exercise 10) is valid.
(e) Show that the norm and inner product in a complex inner product space
are related by

(the "polarization identity").


(f) If X is any complex normed linear space whose norm satisfies the par-
allelogram law, then X is an inner product space (and conversely). [Hint:
Define an inner product on X suggested by the polarization identity.]
19. If I is any index set, consider the set of all functions x : I -+ <C such that
LiEf Ix(i)12 < 00. Then

(x, y) = Lx(i)y(i)
iEf

defines an inner product on this space. We call this space "complex l2(1)."
Show that it is a Hilbert space. [Hint: Proof just as for "real" l2(1).]
20. If in Exercise 15 we consider bounded continuous complex functions x on I
with the inner product

(x, y) = 1 w(t)x(t)y(t) dt,

* we obtain the inner product space called "complex C2 (1; w)."


21. Let (T, S, J.l) be a measure space and consider the linear space of all mea-
surable complex functions x on T with

If we also define
(x,y) = l xYdJ.l

(where the bar denotes complex conjugation), the resulting inner product
space obtained is called "complex L 2 (J.l)". Show that this space is a Hilbert
space. [Hint: Proof as for "real" L 2 (J.l)·]

Historical Notes
The Schwarz Inequality (Theorem 1.2) has a long history. In the specific case of
a double integral over a domain in the complex plane, it was established by Schwarz
HISTORlCAL NOTES 19

(1885; Vol. 1, pp. 223-269) in a highly original paper where he studied the problem
of a vibrating membrane. The same inequality had actually been discovered earlier
by Buniakowsky (1859), but went unnoticed by most mathematicians until 1885.
The corresponding inequality for finite sums (i.e., in £2(n)) goes back at least to
Cauchy (1821; p. 373) and, in the special case when n = 3, even to Lagrange
(1773; pp. 662-3). To appreciate the power and originality of Schwarz's (1885)
paper, Dieudonne (1981; pp. 53-54) showed how Schwarz's method could be easily
translated, with little effort, to the theory of compact self-adjoint operators on a
separable Hilbert space.
Jordan and von Neumann (1935) showed that the parallelogram law actually
characterizes those normed linear spaces that are inner product spaces (see also
Exercise 13). The reader can find 350 characterizations of inner product or Hilbert
spaces among all normed linear spaces in the unusual book of Amir (1986).
Hilbert (1906), inspired by the work of Fredholm in (1900) and (1903) on integral
equations, recognized that the theory of integral equations could be subsumed in the
theory of infinite systems of linear equations. This led him to a study of the space
that we now call £2, and to the symmetric bilinear forms on this space. While he
did not use the geometric language that later workers did, he seemed motivated by
the analogy with n-dimensional Euclidean space. In this regard, we should mention
that the axioms for finite-dimensional Euclidean space ]Rn were stated explicitly by
Weyl (1923), who regarded]Rn as the natural setting for problems in geometry and
physics.
Hilbert's work inspired a transfer of Euclidean geometry to infinite dimensions.
Indeed, Frechet (1906) introduced the notion of distance between points in an ar-
bitrary set, and the idea of a metric space. Schmidt (1908) defined what we now
call (complex) £2, along with the notions of inner product and norm, orthogonality,
closed sets, and linear subspaces. He even proved the existence of the orthogonal
projection (in our terminology, the best approximation) of any point in £2 onto any
given closed subspace. (See Chapter 2 for the definition of best approximation.)
It was von Neumann (1929a) who first defined and studied what we now call
(an abstract) Hilbert space. (Actually, he defined a Hilbert space to have the
additional properties of being infinite-dimensional and separable. It was L6wig
(1934) and Rellich (1935) who studied abstract Hilbert spaces in general.) Von
Neumann (1929a) proved the Schwarz inequality in its present form (Theorem 1.2),
that Hilbert spaces are normed linear spaces, studied orthonormal and complete or-
thonormal sets (or "abstract" Fourier analysis), orthogonal projections onto closed
subspaces, and general self-adjoint operators on Hilbert spaces. In two subsequent
papers, von Neumann (1929b), (1932) continued to develop the theory of most of
the important linear operators on Hilbert spaces. In short, much of what we know
today about Hilbert space and the linear operators on such a space can be traced
to these three fundamental papers of von Neumann.
CHAPTER 2

BEST APPROXIMATION

Best Approximation
We will describe the general problem of best approximation in an inner product
space. A uniqueness theorem for best approximations from convex sets is also
provided. The five problems posed in Chapter 1 are all shown to be special cases
of best approximation from a convex subset of an appropriate inner product space.
2.1 Definition. Let K be a nonempty subset of the inner product space X and
let x E X. An element Yo E K is called a best approximation, or nearest point,
tox from K if
Ilx - Yo II = d(x,K),
where d(x, K) := infYEK IIx - YII. The number d(x, K) is called the distance from
x to K, or the error in approximating x by K.
The (possibly empty) set of all best approximations from x to K is denoted by
PK(X). Thus
PK(X):= {y E K I IIx - yll = d(x,K)}.
This defines a mapping PK from X into the subsets of K called the metric
projection onto K. (Other names for metric projection include projector, nearest
point mapping, proximity map, and best approximation operator.)
If each x E X has at least (respectively exactly) one best approximation in
K, then K is called a proximinal (respectively Chebyshev) set. Thus K is
proximinal (respectively Chebyshev) if and only if PK(x) =1= 0 (respectively PK(x)
is a singleton) for each x E X.
The general theory of best approximation may be briefly outlined as follows: It
is the mathematical study that is motivated by the desire to seek answers to the
following basic questions, among others.
(1) (Existence of best approx;imations) Which subsets are proximinal?
(2) (Uniqueness of best approximations) Which subsets are Chebyshev?
(3) (Characterization of best approximations) How does one recognize when a
given element y E K is a best approximation to x?
(4) (Error of approximation) How does one compute the error of approximation
d(x, K), or at least get sharp upper or lower bounds for it?
(5) (Computation of best approximations) Can one describe some useful algo-
rithms for actually computing best approximations?
(6) (Continuity of best approximations) How does PK(X) vary as a function of
x (or K)?
In this book we will answer each of these questions by proving one or more
general theorems. It will turn out that most of the theorems are suggested by
rather elementary geometric considerations.

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
22 BEST APPROXIMATION

A number of applications to specific inner product spaces will also be made. In


addition, some problems will be studied that are not immediately suggested by the
above six questions. For example, it is known (see Chapter 3) that every closed
convex subset of a Hilbert space is a Chebyshev set. The question whether the
converse holds (Le., must every Chebyshev subset of a Hilbert space be convex?)
is still unanswered. However, if the Hilbert space is finite-dimensional, the answer
is affirmative (see Chapter 12).

Convex Sets
Since the theory of best approximation is the most well developed when the
approximating set K is a linear subspace, or more generally, a convex set, we will
first describe such sets.
2.2 Definition. A subset K of X is called convex ifAx+(l-A)Y E K whenever
x, y E K and 0 ::::: A ::::: 1.
Geometrically, a set is convex if and only if it contains the line segment

{AX + (1 - A)y I 0 ::::: A ::::: 1 }

joining each pair of its points X, y (see Figure 2.2.1). By a simple induction argument
one can show that K is convex if and only if I:~ AiXi E K whenever Xi E K, Ai 2': 0,
and I:~ Ai = 1. (See also Exercise 10 at the end of the chapter.)

Convex set Nonconvex set

Figure 2.2.1

2.3 Examples of Convex Sets.


(1) The empty set, the whole space X, and any singleton set are convex.
(2) The intersection of any collection of convex sets is convex.
(3) The "closed ball centered at X with radius r," which is defined by

B[x, r] := {y E X Illx - yll : : : r},


is convex.
(4) The "open ball centered at X with radius r ," which is defined by

B(x,r) := {y E X Illx - yll < r},


CONVEX SETS 23

is convex.
(5) If A and B are convex, then their sum
A + B:= {a + b I a E A, bE B}
is also convex. In particular, any translate K +x of a convex set K is
convex.

Proof. (1) This is obvious.


(2) If {Ki liE I} is a collection of convex sets, x, Y E niKi, and 0 ::; A ::; 1, then
x, Y E Ki for all i implies AX + (1- A)Y E Ki for all i. Hence AX + (1 - A)Y E niKi
and niKi is convex.
(3) Let Yl, Y2 E B[x, rJ and 0 ::; A ::; 1. Then
Ilx - [AYI + (1 - A)Y2JII = IIA(X - Yl) + (1- A)(X - Y2)11
::; Allx - Ylil + (1- A)llx - Y211
::; Ar + (1 - A)r = r
implies that AYI + (1 - A)Y2 E B[x, rJ. Thus B[x, rJ is convex.
(4) The proof is similar to (3).
(5) Let Yl, Y2 E A + Band 0 ::; A ::; 1. Then Yl = al + b1 and Y2 = a2 + b2 for
some ai E A and bi E B. Then since A and B are convex,
AYI + (1 - A)Y2 = [Aal + (1 - A)a2J + [Ab 1 + (1 - A)b2J E A + B.
Thus A +B is convex. •
2.4 Uniqueness of Best Approximations. Let K be a convex subset of X.
Then each x E X has at most one best approximation in K.
In particular, every convex proximinal set is Chebyshev.
Proof. Let x E X and suppose Yl and Y2 are in PK(x). Then (Yl + Y2)/2 E K
by convexity and

d(x, K) ::; Ilx - ~(Yl + Y2) I = II ~(x - Yl) + ~(x - Y2) II


1 1
::; 2"llx - yIiI + 2"llx - Y211 = d(x, K).
Hence equality must hold throughout these inequalities. By the condition of equal-
ity in the triangle inequality, x - Yl = p(x - Y2) for some p 2: O. But Ilx - yIiI =
d(x, K) = IIx - Y211 implies p = 1 and hence Yl = Y2· •
Alternative proofs of Theorem 2.4 are described in Exercises 8 and 9 at the end
of the chapter. One proof is based on the fact that PK(x) = K n B[x, d(x, K)J is
a convex set that is contained in a "sphere" (Le., the boundary of a ball), and the
only convex subsets of spheres are singleton sets. Another proof is based on the
parallelogram law.
Theorem 2.4 is important for our purposes. Since we will be approximating
almost exclusively by convex sets, uniqueness of best approximations is always
guaranteed whenever best approximations exist. Hence every convex proximinal set
is automatically Chebyshev, and its metric projection is a point-valued mapping.
We will use this fact frequently in the subsequent material without making explicit
reference to Theorem 2.4.
There are certain convex sets that often arise in the various applications.
24 BEST APPROXIMATION

2.5 Definitions. (1) A subset C of X is called a convex cone if ax + f3y E C


whenever x, y E C and a, f3;::: O.
(2) A nonempty subset M of X is called a (linear) subspace if ax + f3y E M
whenever x, y E M and a, f3 E R
(3) If A is any nonempty subset of X, the subspace spanned by A, written
span(A), is the set of all finite linear combinations of elements of A. That is,

Equivalently, span(A) is the smallest subspace of X that contains A (see Exercise


12 at the end of the chapter).
(4) The dimension of a subspace M, denoted by dim M, is the number of
vectors in a basis for M. In particular, if {Xl,X2, ... ,xn } is a linearly independent
set and M = span{xl,x2, ... ,Xn }, then dim M = n.

Convex cone

Figure 2.5.1

2.6 Examples of Cones and Subspaces.


(1) The empty set, the whole space X, and the set {O} consisting of just the zero
vector are all subspaces.
(2) Every subspace is a convex cone, but not conversely.
(3) The intersection of any collection of convex cones (respectively subspaces) is
a convex cone (respectively subspace).
(4) In C2 [a, b], the set of all nonnegative functions

C = {x E C 2 [a, b]1 x(t) ;::: 0 for all t E [a, b]}

is a convex cone that is not a subspace.


(5) The set of all polynomials of degree at most n,

is an (n + I)-dimensional subspace ofC2 [a, b].


Proof. (3) Let {Ci liE I} be a collection of convex cones and C = niCi. If
x, y E C and a, f3 ;::: 0, then x, y E Ci for all i implies ax + f3y E Ci for all i. Hence
ax + f3y E C, so C is a convex cone. The proof that the intersection of subspaces
is a subspace is similar.
CONVEX SETS 25

The proofs of the remaining statements are left as exercises (see Exercises 6(b),
7, and 17 at the end of the chapter). •
Certain elementary properties of the distance function allow us to replace the
problem of approximating from a set K with that of approximating from a translate
of K. Moreover, they also allow us to deduce that translates or scalar multiples of
Chebyshev sets or proximinal sets are again sets of the same kind. We record a few
of these properties next.
2.7 Invariance by Translation and Scalar Multiplication.
(1) Let K be a nonempty subset of X. Then:
(i) d(x -+- y, K -+- y) = d(x, K) for every x, y E X.
(ii) PK+y(x -+- y) = PK(X) -+- y for every x, y E X.
(iii) d(ax, aK) = lald(x, K) for every x E X and a E lR.
(iv) PaK(ax) = aPK(x) for every x E X and a E lR.
(v) K is proximinal (respectively Chebyshev) if and only if K -+- y is proximinal
(respectively Chebyshev) for any given y E X.
(vi) K is proximinal (respectively Chebyshev) if and only if aK is proximinal
(respectively Chebyshev) for any given a E R\{O}.
(2) Let M be a nonempty subspace of X. Then:
(i) d(x -+- y, M) = d(x, M) for every x E X and y E M.
(ii) PM(x -+- y) = PM(X) -+- y for every x E X and y E M.
(iii) d(ax, M) = lald(x, M) for every x E X and a E lR.
(iv) PM(ax) = aPM(x) for every x E X and a E lR.
(v) d(x -+- y, M) :s: d(x, M) -+- d(y, M) for every x and yin X.
Proof. (1) (i) For any x, y EX,
d(x -+- y, K -+- y) = zEK
inf Ilx -+- y - (z -+- y)11 = zEK
inf Ilx - zll = d(x, K).

(ii) Using (i), Yo E PK+y(x -+- y) if and only if Yo E K -+- y and Ilx -+- y - Yoll =
d(x -+- y, K -+- y) if and only if Yo - Y E K and Ilx - (YO - y) II = d(x, K) if and only
if Yo - Y E PK(X); i.e., Yo E PK(X) -+- y.
(iii) We have
d(ax, aK) = yEK
inf Ilax - ayll = lal inf Ilx - yll = lald(x, K).
yEK
(iv) If a = 0, the result is trivially true. Thus assume a =1= 0. Using (iii) we
see that Yo E PaK(ax) if and only if Yo E aK and Ilax - yoll = d(ax, aK) if and
only if ±Yo E K and lallix - ±yoll = lald(x, K). However, this is equivalent to
±YO E PK(x); i.e., Yo E aPK(x).
(v) This is an immediate consequence of (ii), and (vi) follows from (iv).
(2) (i) and (ii) follow from (1) (i) and (ii), and the fact that if M is a subspace
and y E M, then M -+- y = M.
(iii) and (iv) follow from (1) (iii) and (iv), and the fact that if M is a subspace
and a =1= 0, then aM = M.
The proof of (v) follows from the relations
d(x-+-y,M) = mEM
inf Ilx-+-y-mll= inf Ilx-+-y-(m-+-m')11
m,m'EM

:s: m,m'ElvI
inf [llx - mil -+- Ily - mill]

= mEM
inf Ilx - mil -+- inf Ily - mill = d(x, M) -+- d(y, M).
m'EM

26 BEST APPROXIMATION

We next show that the interior and closure of a convex set are convex.
2.8 Theorem. Let K be a convex set in X. Then both K and intK are convex.
Moreover, ifintK =1= 0, then for every x E intK, y E K, and 0 < A :S 1, there
follows

(2.8.1) Ax + (1 - A)y E int K,

and hence

(2.8.2) intK = K and intK = intK.

Proof. Let x, y E K and 0 :S A :S 1. By Lemma 1.15(4), we can choose sequences


{x n }, {Yn} in K such that Xn ~ x and Yn ~ y. Then AXn + (1 - A)Yn E K by
convexity, and
Ax + (1 - A)y = lim[Ax n + (1 - A)Ynl E K.
Thus K is convex.
Now fix any x E int K, y E K, 0 < A < 1, and set z = Ax + (1 - A)y. Since
x E int K, choose € > 0 such that B(x, €) C K.
Claim: B(z,A€) C K and hence z E intK.
To see this, let u E B(z, A€), and hence Ilu - zll < A€. Since y E K, it follows by
Lemma 1.15(3) that
B( A€ - Ilu - zll) K =1= 0
y, 1- A n .
In particular, choose UI E K such that

1
IluI - yll < 1 _ A[A€ - Ilu - zlll·

Now set Uo := Hu - (1 - A)UIl. Then

1 1
Iluo - xII = 11);"[u - (1 - A)UIl- xII = 11);"[u - (1 - A)UI - Axlll
1 1 (1 - A)
= );"Ilu - (1- A)UI - [z - (1- A)ylll :S );"Ilu - zll + - A - Ily - uI11
1 1
< );"Ilu - zll + );"[A€ -liu - zlll = €.

Thus Uo E B(x, €) C K. By definition of Uo and the convexity of K, u = Auo + (1-


A)UI E K. This proves the claim, and hence verifies (2.8.1).
It is clear that int K eKe K. Hence int K C K. For the reverse inclusion, let
y E K and choose any x E intK. Then by (2.8.1),

xn := _l_x + (1- ~1)Y E intK for all n.


n+l n+

Since Xn ~ y, it follows that y E int K. This proves the first equation in (2.8.2).
FIVE BASIC PROBLEMS REVISITED 27

Since K c K, we have int K c int K. For the reverse inclusion, let y E int K.
Then there exists E > 0 such that B(y, E) C K. Now let x E int K \ {y} and set
z := (1 + p)y - px, where p := 2I1x~YII' Since liz - yll = plly - xii = ~, we see that
z E B(y, E) c K. But y = (1 + p)-l z + p(l + p)-lx E int K by (2.8.1). This proves
int K C int K and verifies the second equation in (2.8.2). •
The boundary of a set S, denoted by bdS, is the closure minus the interior of
S:
bd S = S \ int S.
The following lemma shows that the boundaries of S and X \ S are the same, and
also that the interior of S is the complement of the closure of the complement of S.
2.9 Lemma. Let S be any subset of X. Then:
(1) intS=X\(X\S),
(2) int (X \ S) = X \ S,
(3) bdS = bd(X \ S).

Proof. (1) Let x E int S. Then there exists E > 0 such that B(x, E) C S. Hence
X \ B(x, E) =:l X \ S, and so

X \ B(x, E) = X \ B(x, E) =:l X \ S.

This implies that B(X,E) n (X \ S) = 0 and hence x E X \ (X \ S). This proves


that int SeX \ (X \ S).
For the reverse inclusion, let x EX \ (X \ S). Since this set is open, there exists
E> 0 such that B(x, c) eX \ (X \ S) eX \ (X \ S) = S. That is, x E int S.
(2) This follows from (1) by replacing S with X \ S.
(3) Let x E bd S( = S \ int S). If x E int (X \ S), there exists E > 0 such that
B(x, c) C X \ S, and so B(x, c) n S = 0. Hence x if. S, which is a contradiction.
Thus x if. int (X \ S).
If x if. X \ S, then there exists E > 0 such that B(x, c) n (X \ S) = 0. It follows
that B(x, c) C S and hence x E int S, a contradiction. This proves that x E X \ S.
That is, x E bd (X \ S). Hence this shows that bd S c bd (X \ S). Replacing S
by X \ S, we obtain bd (X \ S) c bdS. Combining these results, we obtain that
bdS=bd(X\S) . •

Five Basic Problems Revisited


We conclude this chapter by noting that each of the five problems stated in
Chapter 1 can be viewed as a problem of best approximation from a convex subset
of an appropriate inner product space.
Problem 1. (Best least-squares polynomial approximation to data) Let
I j = 1,2, ... ,m} be a table of data. For any fixed integer n < m, find
{(tj, x(tj ))
a polynomial pet) = I:~ Cl'.iti, of degree at most n, such that the expression

2)X(tk) - P(tkW
k=l
28 BEST APPROXIMATION

is minimized.
Let T = {tl, t2, ... , t m }, X = 12(T), and let Pn denote the subspace of X
consisting of all polynomials of degree at most n. Thus the problem may be restated
as follows: Minimize !Ix - pl12 over all p E Pn- That is, End a best approximation
to x E X from the subspace P n .
Problem 2. (Solution to an overdetermined system of equations) Consider the
linear system of m equations in the n unknowns Xl, X2, ... , xn:

or briefly, Ax = b. Find a vector x E ]Rn that minimizes the expression

Let X = 12(m). Then the problem is to find an x E]Rn that minimizes IIAx-bI12.
Letting M denote the "range of A," i.e.,

M = {y E X I y = Ax, x E ]Rn},

we see that M is a subspace of X. Thus, if we End a best approximation y E M to


b, then any x E ]Rn with Ax = y solves the problem.
Problem 3. (Best least-squares polynomial approximation to a function) Let x
be a real continuous function on the interval [a, b]. Find a polynomialp(t) = 2::~ (Xiti
of degree at most n that minimizes the expression

Let X = C 2 [a, b] and M = P n . Then the problem may be restated as follows:


Find a best approximation to x from the subspace M.
Problem 4. (A control problem) The position () of the shaft of a dc motor
driven by a variable current source u is governed by the differential equation

(4.1) (/' (t) + (}'(t) = 1t(t), 0(0) = 0' (0) = 0,

where u(t) is the field current at time t. Suppose that the boundary condition is
given by

(4.2) 0(1) = 1, 0'(1) = 0,

10
and the energy is proportional to 1 u 2 (t)dt. In the class of all real continuous
funtions u on [0,1] that are related to 0 by (4.1), find the one so that (4.2) is
satisfied and has minimum energy.
FIVE BASIC PROBLEMS REVISITED 29

If w(t) denotes the shaft angular velocity at time t, then

(4.3) (}(t) = it w(s)ds and (}'(t) = w(t).

Thus equations (4.1) and (4.2) may be written in the equivalent form

(4.4) w'(t) + w(t) = u(t), w(O) = 0,


subject to

(4.5) 11 w(t)dt = 1 and w(l) = o.


The solution of (4.4) is easily seen to be given by

(4.6) w(t) = it u(s)es-tds.

Thus, using (4.5) and (4.6), our problem reduces to that of finding u E C 2 [0, 1] that
has minimal norm in the set of all such functions that satisfy

(4.7)

and

(4.8) 11 u(s)e S- l ds = O.

Let X[O,tj denote the characteristic function of the interval [0, t]; Le., X[O,tj (s) = 1
if s E [0, t], and X[O,tj (s) = 0 otherwise. By interchanging the order of integration,
we can rewrite (4.7) as

1= 11 [it u(s)es-tds] dt = 11 [11 x[O,tj(s)u(S)es-tds] dt

= 11 [11 x[O,tj(s)u(s)es-tdt] ds = 11 u(s)e S [11e-tdt] ds


= 11 u(s)eS[-e-tl!]ds = 11 u(s)[l- eS- 1 jds.

Finally, letting X = C 2 [0, 1], Xl (t) = et-l, and X2 (t) = 1 - et-l, we see that the
problem may be stated as follows. Find u E X satisfying

and having minimal norm. If we set

v = {u E X I (U,X1) = 0, (U,X2) = I},


we see that V is a convex set and the problem may be restated as follows: Find the
best approximation to 0 from V.
30 BEST APPROXIMATION

Problem 5. (Positive constrained interpolation) Let {Xl, X2, ... , xn} be a finite
set of functions in L2[0, 1], and let b1 , b2 , ... , bn be real numbers. In the class of all
functions X E L2 [0, 1] such that

(5.1) X(t) 20 for all t E [0,1]

and
10 1
x(t)xi(t)dt = bi (i = 1,2, ... , n),

find the one such that

is minimized.
Letting X = L2[0, 1] and

K={xEXlx20, (x,xi)=bi (i=1,2,···,n)),

we see that the problem reduces to finding the best approximation to 0 from the
convex set K, i.e., determining PK(O).

Exercises

1. Show that the open ball

B(x, r) = {y E X Illx - yll < r}


is convex.
2. Show that the interval

[x, y] = {Ax + (1 - A)y I 0 ::; A ::; I}

is convex.
3. Show that a set K is convex if and only if AK + (1 - A)K = K for every
0::; A::; 1.
4. Show that the following statements are equivalent for a set C:
(1) C is a convex cone;
(2) X +y E C and px E C whenever x,y E C and p 20;
(3) aC + (3C = C for all a,(3 2 0 with a 2 + (32 i= 0, and 0 E C.
5. Show that the following statements are equivalent for a set M:
(1) M is a subspace;
(2) x + y E M and ax E M whenever x, y E M and a E JR.;
(3) aM + (3M = M for all a, (3 E JR. with a 2 + (32 i= O.
6. (1) Verify that the sum

A +B = {a + b I a E A, b E B}
of two convex cones (respectively subspaces) is a convex cone (respectively
subspace).
(2) Verify that the scalar multiple

aA := {aa I a E A}
EXERCISES 31

of a convex set (respectively convex cone, subspace) is a convex set (respec-


tively convex cone, subspace).
7. Show that every subspace is a convex cone, and every convex cone is convex.
Give examples in h (2) of a convex cone that is not a subspace, and a convex
set that is not a convex cone.
8. Give an alternative proof of the uniqueness theorem for best approximation
along the following lines. Let K be a convex set. Prove the following
statements.
(a) PK(X) = K n B[x, d(x, K)], and hence PK(X) is a convex set.
(b) The only nonempty convex subsets of the sphere

S[x, r] := {y EX/ /Ix - y/l = r}


are the singletons.
(c) PK(x) C S[x,d(x,K)].
(d) PK(x) is either empty or a singleton for each x E X.
9. Give an alternative proof of the uniqueness theorem using the parallellogram
law. [Hint: If YI, Y2 E PK(X), deduce YI = Y2 from the relation

/lYI - Y2/12 = /I (x - Y2) - (x - yII/2


= 2/lx - Y2/12 + 2/1x - YI/l 2 - 4/1x - ~(YI + Y2)/l2.]
10. Let A be a nonempty subset of X. The convex hull of A, denoted by coCA),
is the intersection of all convex sets that contain A. Verify the following
statements.
(1) coCA) is a convex set; hence coCA) is the smallest convex set that contains
A.
(2) A is convex if and only if A=co(A).
(3) coCA) = n=~ .AiXi I Xi E A, .Ai 2: 0, I:~ .Ai = 1, n EN}.
11. Let A be a nonempty subset of X. The conical hull of A, denoted by
conCA), is the intersection of all convex cones that contain A. Verify the
following statements.
(1) con (A) is a convex cone; hence conCA) is the smallest convex cone that
contains A.
(2) A is a convex cone if and only if A = conCA).
(3) conCA) = u=~ PiXi I Xi E A, Pi 2: 0, n EN}.
12. Let A be a nonempty subset of X. Verify the following statements.
(1) span(A) is the intersection of all subspaces of X that contain A.
(2) span(A) is the smallest subspace of X that contains A.
(3) A is a subspace if and only if A = span(A).
13. A set V in X is called affine if ax + (1 - a)y E V whenever x, Y E V and
a E R. The affine hull of a set S, denoted by aff(S), is the intersection of
all affine sets that contain S. Prove the following statements.
(1) V is affine if and only if V is a translate of a subspace. That is, V = M +v
for some (unique) subspace M and any v E V.
(2) The intersection of any collection of affine sets is affine. Thus aff(S) is
the smallest affine set that contains S.
(3) aff(S) = n=~ aiXi / Xi E S, ai E R, I:~ ai = 1, n EN}.
(4) The set S is affine if and only if S = aff(S).
32 BEST APPROXIMATION

14. Prove that if K is a convex cone (respectively a subspace, affine), then its
closure K is also a convex cone (respectively a subspace, affine).
15. Show that B[x,r] = B(x,r).
16. (1) Verify that the sum

A +B := {a + b I a E A, b E B}

of two affine sets is an affine set.


(2) Verify that any scalar multiple

aA:={aalaEA}

of an affine set A is an affine set.


17. Prove that P n is an (n + I)-dimensional subspace of C 2 [a, b]. [Hint: Verify
that {XQ, Xl, ... ,xn } is linearly independent, where Xi(t) = t i .]
18. If K is a convex Chebyshev set in an inner product space X and X E X \ K,
show that PK(X') = PK(x) for every x' E co{x, PK(X)}. That is, if A E [0,1]
and x' = AX + (1 - A)PK(X), then PK(X') = PK(x).
19. Show that the uniqueness theorem (Theorem 2.4) is also valid in a complex
inner product space.

Historical Notes
The term "metric projection" goes back at least to Aronszajn and Smith (1954),
while the term "Chebyshev set" goes back to Efimov and Stechkin (1958). The
term "proximinal" is a combination of the words "proximity" and "minimal" and
was coined by Killgrove (see Phelps (1957; p. 790)). That the interior and closure
of a convex set is convex (Le., Theorem 2.8) is classical. (See Steinitz (1913, 1914,
1916), and the more modern classic by Rockafellar (1970) for these and other results
on convexity.)
The control problem (Problem 4) can be found in Luenberger (1969; p. 66).
There are several good introductory books on approximation theory that include
at least one or two chapters on best approximation in inner product spaces. For
example, Achieser (1956), Davis (1963), Cheney (1966), Rivlin (1969), and Laurent
(1972). A fairly detailed overview of the general theory of best approximation in
normed linear spaces can be found in Singer (1970) and Singer (1974).
CHAPTER 3

EXISTENCE AND UNIQUENESS OF BEST APPROXIMATIONS

Existence of Best Approximations


The main existence theorem of this chapter states that every approximatively
compact set is proximinal. This result contains virtually all the existence theorems
of interest. In particular, the two most useful existence and uniqueness theorems can
be deduced from it. They are: (1) Every finite-dimensional subspace is Chebyshev,
and (2) every closed convex subset of a Hilbert space is Chebyshev.
We first observe that every proximinal set must be closed.

3.1 Proxirninal Sets are Closed. Let K be a proximinal subset of X. Then


K is closed.
Proof. If K were not closed, there would exist a sequence {x n } in K such that
Xn -+ x and x 'Ie K. Then d(x, K) -s: Ilx - xnll -+ 0, so that d(x, K) = 0. But
°
Ilx - yll > for each y E K, since x 'Ie K. This contradicts PK(x) #0. •
The converse to Theorem 3.1 is false. Indeed, in every incomplete inner product
space there is a closed subspace that is not proximinal. This general fact will follow
from results proved in Chapter 6 (see, e.g., Exercise 6(b) at the end of Chapter 6).
For the moment, we content ourselves with a particular example.

3.2 A Closed Subspace of C 2 [-1, 1] That Is Not Proxirninal. Let

Then M is a closed subspace ofC2 [-1, 1] that is not proximinal.


Proof. To see that M is closed, let {Yn} eM and Yn -+ y. Then by Schwarz's
inequality

Hence J~ y(t) dt = 0, which implies y E M, and M is closed. M is clearly a subspace


by linearity of integration.
Next define x on [-1,1] by x(t) = 1 for all t. Then x E C2 [-1, 1], and for each

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
34 EXISTENCE AND UNIQUENESS OF BEST APPROXIMATIONS

yEM,

Ilx - Yl12 = [II Ix(t) - y(tW dt = [°111 - y(tW dt + 10111- y(tW dt


= [: 11- y(tW dt + 10 [1 - 2y(t) + y2(t)] dt
1

= 1°11 -
-1
y(tW dt + 1+
10r y2(t) dt 2: 1,
1

and equality holds if and only if y(t) = 1 for -1 ~ t ~ 0 and y(t) = 0 for 0 ~ t ~ 1.
But such a y is not continuous. This proves that d(x, M) 2: 1 and Ilx - yll > 1 for
all y E M.
Next, given any 0 < E < 1, define YE on [-1,1] by

if-l~t~-E,

if -E < t < 0,
ifO~t~1.

Then it is easy to check that YE E M and Ilx - YE 112 = 1+ E/3. Hence d(x, M) ~ 1.
From the preceding paragraph, it follows that

d(x, M) = 1 < Ilx - yll for all y E M.

Thus x has no best approximation in M. •


3.3 Definition. A nonempty subset K of X is called:
(1) complete if each Cauchy sequence in K converges to a point in K;
(2) approximatively compact if given any x E X, each sequence {Yn} in K with
Ilx - Ynll -+ d(x, K) has a subsequence that converges to a point in K.
Remarks. One can easily verify (see Exercise 3 at the end of the chapter) the
following. (1) Every complete set is closed, and (2) in a Hilbert space, a nonempty
set is complete if and only if it is closed. The main reason for introducing complete
sets is that often the inner product space X is not complete, but many of the results
will nevertheless hold because the set in question is complete. An important ex-
ample of a complete set in any inner product space is a finite-dimensional subspace
(see Theorem 3.7 below).
A sequence {Yn} in K satisfying Ilx - Ynll -+ d(x, K) is called a minimizing
sequence for x. Thus K is approximatively compact if each minimizing sequence
has a subsequence converging to a point in K. It is useful to observe that every
minimizing sequence is bounded. This is a consequence of the inequality IIYnl1 ~
llYn - xii + Ilxll -+ d(x, K) + Ilxll·
3.4 Existence of Best Approximations. (1) Every approximatively compact
set is proximinal.
(2) Every complete convex set is a (approximatively compact) Chebyshev set.
Proof. (1) Let K be approximatively compact, and let x E X. Choose a
minimizing sequence {Yn} in K for x. Let {Ynj} be a subsequence that converges
to some Y E K. By Theorem 1.12,

Ilx - yll = lim Ilx - Ynj I = d(x, K).


UNIQUENESS OF BEST APPROXIMATIONS 35

Thus Y is a best approximation to x from K, and hence K is proximinal.


(2) Let K be a complete convex set and fix any x E X. Suppose that {Yn} is
minimizing for x: Ilx - Ynll -+ d(x, K). Then by the parallelogram law (Theorem
1.5),
llYn - Yml1 2= II(x - Ym) - (x - Yn)11 2
= 2(llx - Yml1 2+ Ilx - Yn11 2 ) -112x - (Ym + Yn)11 2
= 2(llx - Yml1 2 + Ilx - Yn11 2 ) - 411x - ~(Ym + Yn)112.

Since K is convex, ~(Ym + Yn) E K, and hence

(3.4.1) llYn - Yml1 2 :s; 2(llx - Yml1 2 + Ilx - Yn11 2) - 4d(x, K?


Since {Yn} is a minimizing sequence for x, it follows that the right side of (3.4.1)
tends to zero as nand m tend to infinity. Thus {Yn} is a Cauchy sequence. Since
K is complete, {Yn} converges to some point Y E K. This proves that K is approx-
imatively compact.
By the first part, K is proximinal. The conclusion now follows from Theorem
2.4 and the fact that K is convex. •
Remarks. (1) The converse of Theorem 3.4(1) is false. That is, there are prox-
iminal sets that are not approximatively compact. For example, let K = {y E
l2 Illyll = I} and consider the sequence {en}, where en(i) = Oni for all n, i. Clearly,
en E K and d(O,K) = 1 = Ilenll for each n, so {en} is a minimizing sequence for O.
But lien - emil = v'2 if n # m implies that no subsequence of {en} converges. Thus
K is not approximatively compact. However, it is easy to verify that PK(O) = K
and PK(x) = {x/llxll} if x # 0 (see Exercise 6 at the end of the chapter). Hence
K is proximinal. However, if K is convex, this cannot happen.
(2) The converse of Theorem 3.4(2) is false. That is, there are convex Chebyshev
sets that are not complete. For suppose
X = {x E £2 I x(n) = 0 except for at most finitely many n}.
Then the set K = {x E X I x(l) = I} is clearly convex, and for any x E X,
the element Y defined by y(l) = 1 and y(i) = xCi) for every i > 1 is in K and is
obviously the best approximation to x from K. However, the sequence of vectors
{xn } defined by xn(i) = Iii for every i :S; nand xn(i) = 0 for every i > n is a
Cauchy sequence in K that does not converge to any point in X. (If it converged,
its limit would be the vector x defined by xCi) = Iii for all i, and this vector is not
in x.)

Uniqueness of Best Approximations


The uniqueness of best approximations from convex sets was already observed
in Theorem 2.4. The next result states that every closed convex subset of Hilbert
space is Chebyshev.
3.5 Closed Convex Subsets of Hilbert Space are Chebyshev. Every
closed convex subset of a Hilbert space is a (approximatively compact) Chebyshev
set.
Proof. Let K be a closed convex subset of the Hilbert space X. By Theorem
3.4(2), it suffices to verify that K is complete. Let {Yn} be a Cauchy sequence in
K. Since X is complete, there exists Y E X such that Yn -+ y. Since K is closed,
Y E K. Thus K is complete. •
36 EXISTENCE AND UNIQUENESS OF BEST APPROXIMATIONS

3.6 Corollary. Every closed convex subset of a Hilbert space has a unique
element of minimal norm.
Proof. Let K be the closed convex set. By Theorem 3.5, 0 has a unique best
approximation Yo E K. That is, IIYol1 < IIYII for all Y E K\{yo}. •
Example 3.2 shows that completeness of X cannot be dropped from the hypoth-
esis of Theorem 3.5 or Corollary 3.6.
The following theorem isolates the key step in the proof of our second main
consequence of Theorem 3.4.
3.7 Finite-Dimensional Subspaces are Complete. Let M be a finite-
dimensional subspace of the inner product space X. Then:
(1) Each bounded sequence in M has a subsequence that converges to a point
inM;
(2) M is closed;
(3) M is complete;
(4) Suppose {Xl, X2, ... , xn} is a basis for M, Yk = I:~=I akiXi, and Y = I:~ aixi·
Then Yk --+ Y if and only if aki --+ ai for each i = 1,2, ... , n.
Proof. Let M = span{xI,x2, ... ,xn } be n-dimensional. We proceed to simul-
taneously prove (1) and (2) by induction on n. For n = 1, let {yd be a bounded
sequence in M. Then Yk = akX!' so {ad must be a bounded sequence in R
Thus there exists a subsequence {ak j } and ao E lR such that akj --+ ao. Setting
Yo = aOxI EM, we see that

Of course, if {yd is actually a convergent sequence, then {yd is bounded (by


Lemma 1.15), and the above argument shows that its limit is in M. This proves
(1) and (2) when n = 1.
Next assume that (1) and (2) are true when the dimension of Mis n and suppose
M = span{xI,x2, ... ,xn,xn+d is (n+l)-dimensional. Set Mn = span{x!, ... ,Xn },
so Mn is n-dimensional. Since Mn is closed by the induction hypothesis and Xn+1 rf:
M n , it follows that d(xn+1' Mn) > O. Let {Yk} be a bounded sequence in M. Then
Yk = Zk+,6k Xn+1, where Zk E Mn. Using Theorem 2.7 (2) and the fact that 0 E M n ,
we get

which implies
1,6 1< IIYkl1
k - d(xn+1' Mn)
Thus {,6d is a bounded sequence, and

implies that {Zk} is a bounded sequence. By the induction hypothesis, there is a


subsequence {Zk,} and Zo E Mn such that Zkj --+ Zo° Since {,6kJ is bounded, by
passing to a further subsequence if necessary, we may assume that ,6kj --+ ,60 E R
Hence setting Yo = Zo + ,60Xn+l, it follows that Yo EM and
UNIQUENESS OF BEST APPROXIMATIONS 37

Again, if {Yd is actually a convergent sequence, this proves that its limit is in M.
Thus (1) and (2) hold for M.
This completes the inductive step and hence the proof statements (1) and (2).
(3) To prove that M is complete, it suffices by (1) to note that every Cauchy
sequence is bounded. But this follows by Lemma 1.17(2).
(4) The proof of the "only if" part of (4) has essentially been done in the course
of establishing (1). We leave the full details to the reader (see Exercise 13 at the
end of the chapter). The proof of the "if" part follows easily from the inequality
n
[[Yk - y[[ .::; L [Ctki - Cti[[[X& •
i=l

While we needed completeness of the whole space to prove Theorem 3.5, it is


not necessary to assume completeness of the space for the next result, since it is a
consequence of the hypotheses that the sets in question are complete, and this is
enough.
3.8 Finite-Dimensional Subspaces are Chebyshev. Let X be an inner
product space. Then:
(1) Every closed subset of a finite-dimensional subspace of X is proximinal.
(2) Every closed convex subset of a finite-dimensional subspace of X is Cheby-
shev.
(3) Every finite-dimensional subspace is Chebyshev.
Proof. Statement (2) follows from (1) and Theorem 2.4. Statement (3) follows
from (2) and Theorem 3.7(2). Thus it suffices to verify (1).
Let K be a closed subset of the finite-dimensional subspace M of X. By Theorem
3.4(1), it suffices to show that K is approximatively compact. Let x E X and let
{Yn} be a minimizing sequence in K for x. Then {Yn} is bounded. By Theorem
3.7(1), {Yn} has a subsequence converging to a point Yo E M. Since K is closed,
Yo E K. Thus K is approximatively compact. •
As a specific application of Theorem 3.8, which is typical of the practical appli-
cations that are often made, we verify the following proposition.
3.9 An Application. Let X = C 2 [a, b] (or X = L 2 [a, b]), n a nonnegative
integer, and
C := {p E Pn [ p(t) 2: 0 for all t E [a, b]}.
Then C is a Chebyshev convex cone in X.
Proof. It is clear that C is a convex cone in the finite-dimensional subspace P n
of X. By Theorem 3.8, it suffices to verify that C is closed. Let {xd be a sequence
in C and Xk -+ x. Since P n is closed by Theorem 3.7, x E P n . It remains to verify
that x 2: O. If not, then x(t o) < 0 for some to E [a, b]. By continuity, there exists
a nontrivial interval [c, d] in [a, b] that contains to and such that x(t) .::; ~x(to) < 0
for all t E [c, d]. Then

[[Xk - x[[2 = lb [Xk(t) - x(tWdt 2: ld [Xk(t) - x(tWdt

2: ld [x(tWdt 2: ld [~x(toWdt
1
= 4[X(tO)[2(d - c) > O.
38 EXISTENCE AND UNIQUENESS OF BEST APPROXIMATIONS

Since the right side is a positive constant, this contradicts the fact that Xk -t x.
Hence x:::: o. •
Compactness Concepts
We next introduce two other types of "compactness." These are stronger than
approximative compactness, and will also be used in the rest of the book.
3.10 Deiinition. A subset K of X is called compact (respectively boundedly
compact) if each sequence (respectively bounded sequence) in K has a subsequence
that converges to a point in K.
The relationship between the various compactness criteria is as follows.
3.11 Lemma. Consider the following statements about a subset K.
(1) K is compact.
(2) K is boundedly compact.
(3) K is approximatively compact.
(4) K is proximinal.
Then (1) =} (2) =} (3) =} (4). In particular, compact convex sets and boundedly
compact convex sets are Chebyshev.
Proof. The implication (1) =} (2) is obvious, and (3) =} (4) was proved in
Theorem 3.4.
To prove (2) =} (3), let K be boundedly compact. If x E X and {Yn} is a
minimizing sequence in K, then {Yn} is bounded. Thus there is a subsequence that
converges to a point in K. This proves that K is approximatively compact.
The last statement is a consequence of Theorem 2.4. •
It is worth noting that the four properties stated in Lemma 3.11 are distinct.
That is, none of the implications is reversible.
An example showing that (4) oj? (3) was given after the proof of Theorem 3.4.
To see that (3) oj? (2), consider the unit ball in l2: K = {y E l2 I IIYII ::; I}.
Then K is a closed convex subset of 12 , so by Theorem 3.5, K is approximatively
compact. Taking the unit vectors en E l2 as defined in the paragraph following
Theorem 3.4, we see that {en} is a bounded sequence in K having no convergent
subsequence. Hence K is not boundedly compact.
To see that (2) oj? (1), let K = span{ed in 12. That is, K = {ael I a E ~}.
If {xn} is a bounded sequence in K, then Xn = anel for a bounded sequence of
scalars {an}. Choose a subsequence {an,} and ao E ~ such that a nk -t ao. Then
x nk -t aOel E K, so K is boundedly compact. However, the sequence Yn = nel E K
has the property that llYn -Ymll = In-ml :::: 1 for all n =f m, so it has no convergent
subsequence. Thus K is not compact.
Compact subsets of a finite-dimensional subspace can be characterized in a useful
alternative way.
3.12 Finite-Dimensional Compacts Sets. (1) Every compact set is closed
and bounded.
(2) A subset of a finite-dimensional subspace is compact if and only if it is closed
and bounded.
Proof. (1) Let K be compact. If {xn} is in K and Xn -t x, choose a subsequence
converging to a point in K. Since this subsequence must also converge to x, the
uniqueness of limits shows that x E K. Hence K is closed. If K were unbounded,
EXERCISES 39

then for each n there would exist Yn E K such that IIYnll > n. Choose a subsequence
{Yn.} converging to a point in K. Then {Yn.} is bounded by Lemma 1.15, which
contradicts llYn. II > nk -t 00. Thus K is bounded.
(2) Let K be a subset of a finite-dimensional subspace M of X. By (1), it suffices
to show that if K is closed and bounded, then it is compact. Let {x n } be a sequence
in K. By Theorem 3.7, it has a subsequence converging to a point Xo EM. Since
K is closed, Xo E K. Thus K is compact. •
The converse of (1) is false. That is, there are closed and bounded sets that are
not compact. In fact, the set K = {x E l2 Illxll ::; I} is not compact since the unit
vectors {en} in K, where en(i) = Oni, satisfy lien - emil = y'2 whenever n i= m.
Thus {en} has no convergent subsequence.

Exercises

1. For any nonempty set K, show that Id(x,K) - d(y,K)1 ::; Ilx - yll for all
x, Y E X. In particular, if Xn -t x, then d(xn, K) -t d(x, K). [Hint: For
any z E K, d(x, K) ::; Ilx - zll ::; IIx - YII + Ily - zII.]
2. Let K be a convex set, x E K, and let {Yn} in K be a minimizing sequence
for x. Show that {Yn} is a Cauchy sequence, hence bounded. [Hint: Look
at the proof of Theorem 3.4.]
3. Verify the following statements.
(a) Closed subsets of complete sets are complete.
(b) Every complete set is closed.
(c) In a Hilbert space, a nonempty subset is complete if and only if it is
closed.
4. Prove the following two statements.
(a) If K is a closed convex subset of an inner product space X that is
contained in a complete subset of X, then K is Chebyshev.
(b) Theorems 3.5 and 3.8(2) can be deduced from part (a).
5. Verify the following statements.
(a) A closed subset of a finite-dimensional subspace of X is boundedly
compact.
(b) Every finite-dimensional subspace is boundedly compact, but never com-
pact.
6. Let K denote the unit sphere in X: K = {x E X Illxll = I}. Verify:
(a) K is not convex.
(b) PK(x) = xlllxll if x E X\{O} and PdO) = K.
(c) K is proximinal, but not Chebyshev.
7. Verify the following statements.
(a) Every finite set is compact.
(b) The union of a finite number of compact sets is compact.
(c) A closed subset of a compact set is compact.
(d) The intersection of any collection of compact sets is compact.
8. If A is compact and B is closed, then A + B is closed. In particular, any
translate of a closed set is closed.
9. If X is infinite-dimensional and M is a finite-dimensional subspace, show
that there exists x E X such that Ilxll = 1 and d(x, M) = 1. [Hint: Take
any Xo E X\M and set x = d(xo, M)-l(XO - PM(XO)).]
40 EXISTENCE AND UNIQUENESS OF BEST APPROXIMATIONS

10. Show that the closed unit ball in X,

B[O, 1] = {x E X I Ilxll :::; I},

is compact if and only if X is finite-dimensional. [Hint: If X is infinite-


dimensional, use Exercise 9 to inductively construct a sequence {x n } in X
such that Ilxnll = 1 and Ilxn - xmll :::: 1 if n ¥- m.]
11. Let

M = {x E l2 I x(n) = 0 for all but finitely many n}.

Show that M is a subspace in l2 that is not closed. What is M?


12. Prove that the set

M = {x E 12 I x(2n) = 0 for all n}

is an infinite-dimensional Chebyshev subspace in 12 . What is PM(x) for any


x E 12?
13. Let C = {x E 12 I x(n) :::: 0 for all n}. Prove that C is a convex Chebyshev
subset of b, and determine Pc(x) for any x E 12 .
14. Let {pd c Pn C C 2 [a, b] and suppose Ilpk I --t O. Show that Pk --t 0
uniformly on [a, b]. That is, for each E > 0, there exists an integer N such
that sUPxE[a,bjIPk(X)1 < E whenever k:::: N. [Hint: Theorem 3.7(4).]
15. Let
X = {x E 12 I x(n) = 0 for all but finitely many n}
and M = {x E X I L~ 2~x(n) = O}. Verify the following statements.
(a) M is a closed subspace in X that is not equal to X.
(b) M.L :== {x E X I x.-l M} = {O}.
[Hint: If z E M.L, then (z, en - 2e n +l) = 0 for every positive integer n.
What does this say about z? Here ej denotes the element in X such that
ej(n) = 0 if n ¥- j and ej(j) = 1.]
(c) If some x E X has 0 as a best approximation in M, then x E M.L.
[Hint: If not, then (x, y) > 0 for some y E M. Then for A > 0 sufficiently
small, the element AY satisfies Ilx-Ayl12 = IIxl1 2 -A[2(x, Y)_AllyI12] < IlxI1 2.]
(d) No element of X \ M has a best approximation in M!
[Hint: If some x E X\M had a best approximation Xo EM, then z = X-Xo
has 0 as a best approximation in M.]
16. In the text, Corollary 3.6 was deduced from Theorem 3.5. Now use Corollary
3.6 to deduce Theorem 3.5. Hence show that Theorem 3.5 and Corollary
3.6 are equivalent. [Hint: A translate of a convex set is convex.]
17. Show that all the definitions and results of this chapter are valid in complex
inner product spaces also.
18. All the definitions and several of the results of this chapter are valid in an
arbitrary normed linear space. In particular, verify that the Theorems 3.1,
3.4(1),3.7,3.11, and 3.12 hold in any normed linear space X.

Historical Notes
The notion of "compactness" has a long history (see the books by Alexandroff
and Hopf (1935) and Kuratowski (1958), (1961)). There is an alternative definition
HISTORICAL NOTES 41

of compactness to the one that we gave that is essential, especially for defining
compactness in general topological spaces. A collection S of open subsets of a
topological space X is called an open cover for the subset K of X if the union of
the sets in S contains K. A finite sub collection of an open cover for K is called a
finite subcover if it is also an open cover for K. The set K is called compact (or,
in the older literature, bicompact) if each open cover of K has a finite subcover.
It is well-known that in a metric space this definition of compactness is equivalent
to the sequential one we gave in Definition 3.10 (see, e.g., Dunford and Schwartz
(1958; Theorem 15, p. 22)).
Bolzano (1817) established the existence of a least upper bound for a bounded
sequence of real numbers. In his (unpublished) Berlin lectures in the 1860s, Weier-
strass used Bolzano's method, which he duly credited to Bolzano, to prove that
every bounded infinite set of real numbers has a limit point (see also Grattan-
Guiness (1970; p. 74) or Kline (1972; p. 953)). The essence of the method was
to divide the bounded interval into two parts and select the part that contained
infinitely many points. By repeating this process, he closes down on a limit point
of the set. It is an easy step from this result to prove what is now called the
"Bolzano-Weierstrass theorem": A subset of the real line is (sequentially) compact
if and only if it is closed and bounded. Indeed, it is also an easy step to prove that
this holds more generally in Rn (see Theorem 3.12).
Heine (1870) defined uniform continuity for functions of one or several variables
and proved that a function that is continuous on a closed bounded interval [a, b]
is uniformly continuous. His method used the result that if the interval [a, b] has
a cover consisting of a countable number of open intervals, then it has a finite
subcover. Independently, Borel (1895) recognized the importance ofthis result and
stated it as a separate theorem. Cousin (1895) showed that the open cover in the
result of Heine and Borel need not be restricted to be countable, although this fact
is often credited to Lebesgue (1904). This result can be easily extended to any
closed bounded set in R, and even to Rn. Moreover, the converse holds as well.
The consequential result is generally called the "Heine-Borel theorem": A subset
of Rn is ("open cover") compact if and only if it is closed and bounded.
The notion of "approximative compactness" goes back at least to Efimov and
Stechkin (1961), while that of "bounded compactness" seems to originate with Klee
(1953).
The fundamental fact that every closed convex subset of a Hilbert space is Cheby-
shev dates back to Riesz (1934), who adapted an argument due to Levi (1906).
That any finite-dimensional subspace of a normed linear space is proximinal was
established by Riesz (1918).
Cauchy (1821; p. 125) gave his criterion for the convergence of a sequence {xn}
of real numbers: {xn} converges to a limit x if and only if IXn+r - xnl can be
made smaller than any positive number for all sufficiently large n and all r > O.
He proved that this condition is necessary, but only remarked (without proof)
that if the condition is satisfied, then the convergence is guaranteed. According to
historian Morris Kline (1972), Cauchy may have lacked a complete understanding
of the structure of the real numbers at that time to give a proof. That the Cauchy
criterion is a necessary and sufficient condition for the convergence of a sequence
in Rn is an easy consequence of the result in R by arguing in each coordinate
separately.
CHAPTER 4

CHARACTERIZATION OF BEST APPROXIMATIONS

Characterizing Best Approximations


We give a characterization theorem for best approximations from convex sets.
This result will prove useful over and over again throughout the book. Indeed, it
will be the basis for every characterization theorem that we give. The notion of
a dual cone plays an essential role in this characterization. In the particular case
where the convex set is a subspace, we obtain the familiar orthogonality condi-
tion, which for finite-dimensional subspaces reduces to a linear system of equations
called the "normal equations." When an orthonormal basis of a (finite or infinite-
dimensional) subspace is available, the problem of finding best approximations is
greatly simplified. The Gram-Schmidt orthogonalization procedure for construct-
ing an orthonormal basis from a given basis is described. An application of the
characterization theorem is given to determine best approximations from a trans-
late of a convex cone. Finally, the first three problems stated in Chapter 1 are
completely solved.
4.1 Characterization of Best Approximations from Convex Sets. Let
K be a convex subset of the inner product space X, x E X, and Yo E K. Then
Yo = PK(X) if and only if

(4.1.1) (x - Yo, Y - Yo) ~ 0 for all y E K.

Proof. If (4.1.1) holds and y E K, then

IIx - yol12 = (x - Yo,X - Yo) = (x - Yo,X - y) + (x - Yo,y - Yo)


~ (x-YO,x-y) ~ IIx-yollllx-yll
by Schwarz's inequality. Hence IIx - Yo II ~ IIx - YII, and so Yo = PK(X).
Conversely, suppose (4.1.1) fails. Then (x - Yo, Y - Yo) > 0 for some y E K. For
each 0 < A < 1, the element Y>.. := AY + (1 - A)YO is in K by convexity and

IIx - y>..11 2 = (x - Y>..,X - Y>..) = (x - Yo - A(Y - Yo), x - Yo - A(Y - Yo))


= IIx - Yo 112 - 2A (x - Yo, Y - Yo) + A211Y - Yo 112
= IIx - Yo 112 - A[2 (x - Yo, Y - Yo) - Ally - yoIl2].

For A > 0 sufficiently small, the term in brackets is positive, and thus IIx _ y>..11 2 <
IIx - yoll2. Hence Yo =1= PK(x). •
There are two geometric interpretations of Theorem 4.1. The first is that the
angle e between the vectors x - Yo and y - Yo is at least 90 degrees for every y E K

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
44 CHARACTERIZATION OF BEST APPROXIMATIONS

(see Figure 4.1.2). The second interpretation is that the convex set K lies on one
side of the hyperplane H that is orthogonal to x - Yo and that passes through Yo
(see Chapter 6 for a discussion of hyperplanes).
H

Y.
·x

Figure 4.1.2

Dual Cones
There is an alternative way to restate the characterization theorem, Theorem
4.1, that will prove to be quite useful later on. It involves the notion of the "dual
cone" of a given set.
4.2 Definition. Let S be any nonempty subset of the inner product space X.
The dual cone (or negative polar) of S is the set

SO: = {xEXI (x,y):::::O forall yES}.


The orthogonal complement of S is the set

s .l. := s On (-SO)={xEXI (x,y)=O forall YES}.


Geometrically, the dual cone So is the set of all vectors in X that make an angle
of at least 90 degrees with every vector in S (see Figure 4.2.1).
The following is an obvious restatement of Theorem 4.1 using dual cones.
4.3 Dual Cone Characterization of Best Approximations. Let K be a
convex subset of the inner product space X, x E X, and Yo E K. Then Yo = PK(x)
if and only if

(4.3.1)

Thus the characterization of best approximations requires, in essence, the calcu-


lation of dual cones. For certain convex sets (e.g., cones, subspaces, or intersections
DUAL CONES 45

:
,.

Figure 4.2.1 Dual cone

of such sets), there are substantial improvements in Theorem 4.3 that are possible.
Before stating them, it is convenient to list some basic properties of dual cones.
Recall that the conical hull of a set S, denoted by con(S), is the intersection
of all convex cones that contain S.
4.4 Conical Hull. Let S be a nonempty subset of X. Then:
(1) con(S) is a convex cone, the smallest convex cone containing S.
(2) S is a convex cone if and only if S = con(S).
(3) con(S) = {2:~ PiXi I Xi E S, Pi ::::: 0, n EN}.
If S is convex, then con(S) = {px I P::::: O,x E S}.
(4)
(5) °
If S is convex and E S, then con(S) = {px I p> 0, XES}.

Proof. The proofs of (1) and (2) are trivial.


(3) Let

C := { tPiXi I Xi E S, Pi ::::: 0, n EN} .

It is clear that C =:J S (e.g., if XES, x = 1 . x E C). Also, let x, Y E C and


a, f3 ::::: 0. Then x = 2:~ PiXi and Y = 2:;" liYi, where Xi, Yi E S and Pi, Ii ::::: 0.
Hence ax + f3y = 2:~ apiXi + 2:;" f3,iYi = 2:~+m 6iZi, where 6i = api and Zi = Xi
if 1 SiS n, and 6i = f3,i-n and Zi = Yi-n if n + 1 SiS n + m. Since Zi E Sand
6i ::::: 0, ax + f3y E C. Thus C is a convex cone containing S. From (1) it follows
that C =:J con(S).
Next let D be any convex cone that contains S. We want to show that D =:J C.
Let x E C. Then x = 2:~ PiXi for some Xi E S and Pi ::::: 0. We proceed by
induction on n. If n = 1, then x = P1Xl E D, since Xl ESC D and D is a cone.
Assume xED whenever x is a nonnegative linear combination of n - 1 elements
of S. Then
n n-l

X = LPiXi = LPiXi + PnXn ED,


1 1
46 CHARACTERlZATION OF BEST APPROXIMATIONS

since D is a cone. This proves D ::) C. Since D was any convex cone containing S,
take D = con(S) to get con(S) ::) C. Hence con(S) = C.
(4) Set C = {px I P 2: 0, XES}. By (3), C c con(S). Conversely, if Y E
°
con(S) \ {O}, then (3) implies that y = I:~ PiXi for some Pi > and Xi E S. Setting
P = I:~ Pi and Ai = pdp, we see that p> 0, Ai > 0, I:~ Ai = 1, x:= I:~ AiXi E S
by convexity, and y = px E C. That is, con(S) c C.
(5) This is an immediate consequence of (4) and the fact that E S. • °
The following two propositions list some of the fundamental properties of dual
cones that will aid us in many applications of Theorem 4.3.
4.5 Dual Cones. Let S be a nonempty subset of the inner product space X.
Then:
(1) So is a closed convex cone and S1- is a closed subspace.
(2) So = (8)° and S1- = (8)1-.
(3) So = [con(S)Jo = [~Jo and S1- = [span(S)j1- = [span(SW.
(4) con(S) c SOO and span(S) C S1-1-.
(5) If C is a convex cone, then (C - y)O = Co n y1- for each y E C.
(6) If M is a subspace, then MO = M1-.
(7) If C is a Chebyshev convex cone (e.g., a closed convex cone in a Hilbert
space), then
Coo = C.
(8) If M is a Chebyshev subspace (e.g., a closed subspace in a Hilbert space),
then
MOO = M1-1- = M.
(9) If X is complete and S is any nonempty subset, then SOo = con(S), S1-1- =
span( S), sooo = So, and S1-1-1- = S1-.

Proof. We will prove only the statements about dual cones. The proofs of the
analogous statements for orthogonal complements can be easily deduced from these
(or verified directly) and are left as an exercise.
(1) Let Xn E So and Xn -t x. Then for each YES,

(x, y) = lim(xn, y) ::; 0

implies x E So and So is closed. Let x, z in So and a, j3 2: O. Then, for each YES,

(ax + (3z, y) = a(x, y! + j3(z, y! ::; 0,


so ax + j3z E So and So is a convex cone.
(2) Since S c 8, then So ::) (8)°. If x E So and y E 8, choose Yn E S such that
Yn -t y. Then
(x, y! = lim(x, Yn! ::; 0
implies x E (8t. Thus So c (8)° and hence So = (8)°.
(3) Since con(S) ::) S, [con(SW C So. Let x E So and y E con(S). By Theorem
4.4, y = I:~ PiYi for some Yi E S and Pi 2: 0. Then
n
(x, Y! = LPi(X, Yi! ::; 0
1
DUAL CONES 47

implies x E [con(SW, so So C [con(SW. Thus So = [con(SW.


The second equality of (3) follows from (2).
(4) Let XES. Then for any y E So, (x,y) ::::: O. Hence x E SOo. That is,
S C SOo. Since SOo is a closed convex cone by (1), con( S) C SOo.
(5) Now, x E (C _y)O if and only if (x, c-y) ::::: 0 for all c E C. Taking c = 0 and
c = 2y, it follows that the last statement is equivalent to (x, y) = 0 and (x, c) ::::: 0
for all c E C. That is, x E Co n yJ...
(6) If M is a subspace, then -M = M implies

(7) Let C be a Chebyshev convex cone. By (4) we have C c Coo. It remains to


verify Coo C C. If not, choose x E COO\C and let Yo = Pc(x). By (5) and Theorem
4.3, we have x - Yo E (C - Yo)O = Co n Y5-- Thus

0< /Ix - yo/l2 = (x - Yo, x - Yo) = (x - Yo, x) ::::: 0,

which is absurd.
(8) This follows from (6) and (7).
(9) This follows from (3), (7), and (8). •
The sum of two sets SI and S2 is defined by SI + S2 = {x +y I x E SI, Y E S2}.
More generally, the sum of a finite collection of sets {SI, S2, ... , Sn}, denoted by
SI + S2 + ... + Sn or E~ Si, is defined by

4.6 Dual Cones of Unions, Intersections, and Sums. Let {SI, ... ,Sm} be
a finite collection of non empty sets in the inner product space X. Then
(1) (U;" Sit = n7' Si ° and (U;" Si)J.. = n;" SiJ...
(2) E;" Sf c (n;" Sir and E;" Sf c (n;" Si)J...
(3) If 0 E n;" Si, then

(4) If {C1 , C2 , ... ,Cm } is a collection of closed convex cones in a Hilbert space,
then

(5) If {Ml' M 2 , .•. , Mm} is a collection of closed subspaces in a Hilbert space,


then
48 CHARACTERIZATION OF BEST APPROXIMATIONS

Proof. We will verify the statements concerning dual cones, and leave the proofs
of the analogous statements concerning orthogonal complements as an exercise.
(1) x E niSY if and only if x E SY for each i if and only if (x, y) ::::; 0 for each
y E Si and all i if and only if (x, y) ::::; 0 for all y E UiSi if and only if x E (UiSit.
(2) Let x E 2:~n Sy. Then x = 2:~ Xi for some Xi E sy. For every y E nl" Si, we
have (x, y) = 2:~ (Xi, y) ::::; 0, so x E (nl"Si)o.
(3) x E (2:i Sit if and only if (x, s) ::::; 0 for each s E 2:i Si if and only if
(x, 2:i Si) ::::; 0 whenever Si E Si if and only if (since 0 E niSi)) (x, Si) ::::; 0 for each
Si E Si and all i if and only if x E sy for each i if and only if x E niSy.
(4) By Theorem 4.5(7), cyo = C i for each i. Also, it is easy to check that the
sum 2:i cy is a convex cone. Using part (3), we obtain

Using Theorem 4.5, it follows that

(5) This is a special case of (4). •


Remarks. (1) There is a natural generalization of Theorem 4.6 to the case of
infinitely many sets. One just needs to make the appropriate definition of the sum
of an infinite number of sets that does not entail any convergence questions. (See
Exercise 23 at the end of the chapter.)
(2) Theorem 4.6 will be particularly useful to us in Chapter 10 when we study
the problem of constrained best approximation.
(3) If Kl and K2 are two convex sets with 0 E Kl n K 2, it is not true in general
that

(4.6.1)

In fact, this result is equivalent to the condition

(4.6.2)

In other words, (4.6.1) holds if and only if the operation of taking the closed conical
hull of Kl n K2 commutes with the operation of taking the intersection. (See
Exercise 24 at the end of the chapter.) Sets that have this property will play an
important role in the characterization theorems to be developed in Chapter 10.
In the particular case where the convex set is a convex cone, Theorem 4.3 can
be strengthened by using Proposition 4.5(5).
4.7 Characterization of Best Approximations from Convex Cones.
Let C be a convex cone in X, x E X, and Yo E C. The following statements are
equivalent:
(1) Yo = Pc(x);
(2) x -- Yo E Co n Y6-;
(3) (x - Yo, y) ::::; 0 for all y E C and (x - Yo, Yo) = o.
DUAL CONES 49

The geometric interpretation of Theorem 4.7 is this: Yo is the best approximation


to x if and only if the error x - Yo is orthogonal to Yo and makes an angle of at
least 90° with each vector in C (see Figure 4.7.3).

X · Yo
"
,,
,,
,,
,
o '~~--------------------------~~
Figure 4.7.3
4.8 An Application: Best Approximation from a Translate of a Convex
Cone. Let X = C 2 [a, b] (or L2 [a, bJ), v EX, and

K = {y E X I y(t) 2: v(t) for all t E [a,b]}.

Then K is a convex Chebyshev set and

(4.8.1) PK(X) = max{x, v}

for every x E X. [Here y = max{x,v} denotes the function defined pointwise by


y(t) = max{x(t),v(t)}.] In particular, if

(4.8.2) K = {y E X I y(t) 2: ° for all t E (a, b]},

we get

(4.8.3) PK(X) = x+ = max{x, O}.

Proof. First note that K = C +v, where C = {y E X I y 2: O} is a convex cone.


Then
CO={xEXI(x,y):::;o forall y2:0}={xEXlx:::;0}.
Let x E X and Yo E C. Then by Theorem 4.7, Yo = Pc(x) if and only if x - Yo E

° ° °
Co n ycf if and only if x - Yo :::; and (x - Yo, Yo) = 0. Since Yo 2: 0, this is obviously
equivalent to x - Yo :::; and (x - Yo)YO = 0, which is equivalent to x - Yo :::; and
Yo(t) = whenever x(t) - Yo(t) < 0. That is, Yo = max{x, a}. This proves that
°
Pc(x) = max{x,O}
50 CHARACTERIZATION OF BEST APPROXIMATIONS

for any x E X. In particular, C is a Chebyshev convex cone in X. By Theorem 2.7


(1) (ii) and (v), we have that K = C + v is a convex Chebyshev set and, for any
xEX,
PK(X) = Pc+v(x) = Pc+v(x - v + v)
= Pc(x - v) +v = max{x - v,O} +v
= max{x, v} . •

*
Remarks. (1) Application 4.8 is valid in the space i!2(T) for any set T.
(2) More generally, Application 4.8 (as well as its proof) is valid in the space
L 2 (T, S, p,) for any measure space (T, S, p,).

Characterizing Best Approximations from Subspaces


There is an even simpler characterization of best approximations when the convex
set is actually a subspace. It is an immediate consequence of Theorem 4.3 and
Proposition 4.5(6).
4.9 Characterization of Best Approximations from Subspaces. Let M
be a subspace in X, x E X, and Yo E M. Then Yo = PM(x) if and only if
x - Yo E M 1.; that is,

(4.9.1) (x - Yo, y) = 0 for all y E M.

The geometric interpretation of Theorem 4.9 is clear: Yo is the best approxima-


tion to x if and only if the error x - Yo is orthogonal to M (see Figure 4.9.2). This
is the reason why PM(x) is often called the orthogonal projection of x onto M.

x,

x-Y
r 0
,

,
, )

)// Yo = PM (x)

o
M

Figure 4.9.2

4.10 Corollary (The Normal Equations). Let {Xl, X2, ... , x n } be a basis
for the n-dimensional subspace M of X. Then M is a Chebyshev subspace, and for
each x E X,
n
(4.10.1) PM(X) = LO<iXi ,
I
GRAM-SCHMIDT ORTHONORMALIZATION 51

where the scalars ai are the unique solution to the norrnal equations

n
(4.10.2) ~ai(xi' Xj) = (x, Xj) (j = 1,2, ... , n) .
i=l

In particular, if {Xl,X2, ... ,Xn} is an orthonormal basis for M, then

n
(4.10.3) PM(X) = ~(X,Xi)Xi for every x E X.

Proof. Fix x E X and Yo E M. Then Yo = L~ aixi for some scalars ai. By


Theorem 4.9, Yo = PM(X) is equivalent to (x - Yo, y) = 0 for every y E M. But the
latter is clearly equivalent to (x - Yo, Xj) = 0 for j = 1,2, ... , n; i.e., (4.10.2) holds.
If {Xl,X2, ... ,Xn } is orthonormal, then (Xi,Xj) = Oij, so (4.10.2) implies ai =
(x, Xi) for each i and hence (4.10.3) holds. •

Gram-Schmidt Orlhonormalization
In matrix-vector notation, the normal equations (4.10.2) may be rewritten as

(4.10.4) G(Xl, X2,· .. , xn)a = (3,

where
(Xn, Xl)
(X n ,X2)
1
(Xn, Xn)
is called the Gram matrix of {Xl,X2, ... ,Xn }, and a and (3 are the column vectors
a = (aI, a2, ... , an)T and (3 = ((x, Xl), (x, X2)' ... ' (x, x n >)T.
It can be shown (see Exercise 7 at the end of the chapter) that the determinant
of the Gram matrix satisfies the inequalities

and equality holds on the left (respectively right) side if and only if {Xl, X2, ... , xn}
is a linearly dependent (respectively orthogonal) set. As a practical matter, it
is better if the determinant of the Gram matrix is as large as possible. This is
because of the well-known "ill-conditioning" phenomenon: If det G(Xl, X2,· . . , xn)
is close to zero, the inverse of the Gram matrix has some entries that are very large;
hence small errors in the calculation of the vector (3 in (4.10.4) (due to rounding,
truncation, etc.) could result in very large errors in the solution vector a.
This suggests that the most desirable situation is to have the determinant of
the Gram matrix as large as possible, that is, when the set {Xl,X2, ... ,xn } is
orthogonal. Moreover, we have just seen in Corollary 4.10 that the normal equations
have an especially simple form in this case. For these (and other) reasons, one would
like to have an orthogonal basis for any given subspace. The next result describes
a constructive method for achieving this.
52 CHARACTERIZATION OF BEST APPROXIMATIONS

4.11 Gram-Schmidt Orthogonalization Process. Let {Xl, X2, ... } be a


finite or countably infinite set of linearly independent vectors. Let Mo = {O} and
Mn = span{ Xl, X2, . .. , xn} for n 2 1. Define
(4.11.1)
for each n 2 1. Then:
(1) {YI, Y2, ... } is an orthogonal set.
(2) {el, e2, ... } is an orthonormal set.
(3) span{YI,Y2, ... ,Yn} = span{el,e2, ... ,en} = Mn for each n 21.
(4) YI = Xl and Yn = Xn - ~~:II<Xn, ei)ei for n 2 2.
In particular, every finite-dimensional subspace has an orthonormal basis.
Proof. By Theorem 3.8, Mn is Chebyshev for each n 2 0 so Yn is well-defined.
Since {Xl, X2, .. . , xn} is linearly independent for each n, Xn tt M n - l . Hence Yn # 0
for all n, so en is well-defined. By Theorem 4.9, Yn E M,t_l for every n 2 1. For
any j < n,
Yj = Xj - PMj _ 1(Xj) C M j C M n - l ,
so Yn .1 Yj. By symmetry, Yn .1 Yj whenever n # j. Thus (1) holds.
Since en = Yn/IIYnll, statement (2) and the first equality in (3) follow from
(1). Since {el,e2, ... ,en } C Mn, Mn is n-dimensional, and orthonormal sets are
linearly independent by Theorem 1.7, we obtain that span{el,e2, ... ,e n } = Mn.
This completes the proof of (3).
Finally, to verify (4), it suffices to show that PMn_1(X n ) = ~~:II<Xn,ei)ei for
n 2 2. But this follows from (2), (3), and (4.10.3). •

Y2

M
,, , j
, I
, , I

,,
I ,
, I

, "
Y3 ,.. ------------------------'

e3l

Figure 4.11.2. The Gram-Schmidt theorem

In practice, one constructs the orthonormal set {el, e2, ... } from the linearly
independent set {Xl, X2, . .. } inductively as follows (see Figure 4.11.2):
GRAM-SCHMIDT ORTHONORMALIZATION 53

YI = Xl ,

n-l

Yn = xn - 2: (Xn' ei)ei , en = Yn/IIYnll,


i=l
(n 2': 2).
4.12 Examples.
(1) (Legendre polynomials) If one orthonormalizes the set of monomials
{l, t, t 2 , ... } in C 2 [-1,1], the so-called Legendre polynomials are obtained. The first
four Legendre polynomials are given by

Po(t) = 1/v'2 ,
P2(t) =
JIO 2
-4-(3t - 1),

In general,

for n 2': o.
(2) (Chebyshev polynomials) Suppose one orthonormalizes the monomials
{I, t, t 2, ... } in the weighted space C 2([-I, 1];w), where w(t) = (1- t 2)-! (recall
Example 1.3 (4) for the definition of this space). Then the so-called (normalized)
Chebyshev polynomials of the first kind are obtained. The first four Chebyshev
polynomials are given by

In general, Tn(t) = v-fTn(t) for n 2': 1, where Tn(t) := cos(n arccos t) (n =


1,2, ... ) are the "usual" Chebyshev polynomials.
(3) (Polynomials on discrete sets) Consider a set of m > 1 distinct real numbers
T = {tl,t2, ... ,tm } and the space l2(T). If one orthonormalizes the monomials
{I, t, t 2, . .. ,tn }, where n :::; m - 1, an orthonormal set of polynomials is obtained.
For example, ifti = i for each i, then l2(T) = 12(m), and the first three orthonormal
polynomials are given by

Po(t) = yIm'
1
- ~[-~(
() -v~t
PIt
)1
2 m + 1J '

andp2(t) = A[t2-(m+l)t+i(m+l)(m+2)], where the positive scalar A is chosen


so that IIp211 = 1.
As the above examples show, the Gram-Schmidt orthonormalization procedure
can be used to compute orthonormal polynomials of arbitrarily high degree. In
practice, however, the following three-term recurrence relation for orthonormal
polynomials is more efficient for their actual computation.
54 CHARACTERIZATION OF BEST APPROXIMATIONS

4.13 Three-Term Recurrence. Let Po, Pl, P2, . .. be the sequence of orthonor-
mal polynomials obtained by orthonormalizing the monomials 1, t, t 2 , . .. via the
Gram-Schmidt procedure. Then

(4.13.1)

for some constants an, bn , and en (n = 1,2, ... ).


Proof. It follows from Theorem 4.11 that P n = span{PO,Pl, ... ,Pn}, Pn ..l P n- 1 ,
and Pn E P n\Pn- 1 for n = 1,2, .... In particular, since t Pn(t) E Pn+l, we see that
there are scalars (};i with (};n+l i 0 such that tPn(t) = 2:~~ol (};iPi(t). For any
O:S j :S n + 1,

That is,

(4.13.2) (j = 0,1,2, ... , n + 1).

But if 0 :S j :S n - 2, then tPj (t) E P n - 1 and

( 4.13.3)

since Pn ..l P n- 1 • Thus

Solving this equation for Pn+l, we obtain (4.13.1) with

(4.13.4)

For the actual computation of the orthonormal polynomials using equations


(4.13.1)-(4.13.4), the following format is convenient:

Y-l = 0, P-l = 0
Po
Yo = 1, Po = iiPoii
( 4.13.5)
Yn+l(t) = [t - (tPn(t),Pn(t)]Pn(t) - (tPn(t),Pn-l(t)Pn-l(t)
Yn+l (n = 0, 1, 2, ... ).
Pn+l = iiYn+lii
Fourier Analysis
We will next prove a generalization of formula (4.10.3) for the case where M is
not necessarily a finite-dimensional subspace, but possibly infinite-dimensional. It
will be convenient to first provide some background material that is of interest in
its own right.
FOURIER ANALYSIS 55

4.14 Theorem. Let {er, e2, . .. ,en} be a finite orthonormal set in the inner
product space X and let M := span{ el, e2,.··, en}. Then, for every x EX,
n
(4.14.1) PM(X) = I)x, ei)ei

and
n
(4.14.2) L l(x,ei)1 2 ::; Ilx11 2 .

Proof. The equality (4.14.1) is just (4.10.3). Moreover, Theorem 4.9 implies
that x - PM(x) E Ml.. Using the Pythagorean theorem (Theorem 1.9), we get

and hence

(4.14.3)

Again by the Pythagorean theorem,


n n
(4.14.4) IIPM(x)11 2 = LII(x, ei)eil1 2 = LI(x, ei)12.

Combining (4.14.3) and (4.14.4), we obtain (4.14.2). •


Inequality (4.14.2) states that the sum of the squares of the components of a
vector in various orthogonal directions does not exceed the square of the length of
the vector. It is this inequality, or more generally the inequality of Theorem 4.16
below, that goes by the name of "Bessel's inequality."
The next lemma is the key to extending Theorem 4.14 to an infinite orthonormal
set.
4.15 Lemma. Let 0 be an orthonormal set in the inner product space X, and
let x EX. Then the set

(4.15.1) Ox := {e E 0 1(x, e) # O}
is countable (Le., either empty, finite, or countably infinite). Moreover, if we write
Ox = {el,e2,"'}' then

(4.15.2)

In particular, the series I:i 1(x, ei) 12 converges.


Proof. For each n E N, let

Sn:= {e E 0 II(x,e)1 2 > IlxI1 2/n}.


56 CHARACTERIZATION OF BEST APPROXIMATIONS

Since Ox = Ul" Sn, it suffices to show that each Sn is finite. But by inequality
(4.14.2), Sn contains at most n - 1 elements. This proves that Ox is countable.
By (4.14.2), we have 2:~=ll(x,eiW <:::: IIxl1 2 for every n. Since the right side of this
inequality is a constant independent of n, it follows by taking the supremum over
n that (4.15.2) holds. •
To prove Bessel's inequality when there are infinitely many orthonormal ele-
ments, we must first explain what the sum 2: eE o 1(x, e) 12 means if 0 is infinite and
possibly even uncountable. Of course, if 0 is finite, say 0 = {el' e2, . .. , en}, then
2:eEO 1(x, e) 12 := 2:~ 1(x, ei) 12. Now let 0 be an infinite orthonormal set in an inner
product space X, and let x EX. By Lemma 4.15, the set Ox = {e E 0 1 (x, e) 1= O}
is countable. Let us index the set Ox in some order, say Ox = {el' e2, ... }. By
Lemma 4.15, the series 2:~ I(x, eiW is convergent. From the theory of absolutely
convergent series, it follows that this series converges to the same sum for every
rearrangement of the terms of this sum. This shows that the ordering we initially
chose in Ox was unimportant, and we can unambiguously define
00

(4.15.3) L l(x,e)1 2 := LI(x,eiW


eEO i=l

for any ordering {el' e2, ... } of the elements of Ox.


We have just proved the following generalized version of inequalities (4.14.2) and
(4.15.2).
4.16 Bessel's Inequality. If 0 is any orthonormal set in an inner product
space X, then

(4.16.1) L 1(x, e) 12 <:::: IIxl1 2 for every x E X.


eEO

This result generalizes the inequality (4.14.2). Before we can generalize (4.14.1)
to the case where M is spanned by an infinite orthonormal set, we need to define
the expression 2:eEO (x, e)e in case the orthonormal set 0 is infinite.
4.17 Definition (Fourier Series). Let 0 be an orthonormal set in the inner
product space X, and for each x E X, let Ox denote the countable set defined (as
in (4.15.1)) by Ox = {e E 0 1 (x, e) 1= OJ. We define the Fourier series of x
relative to 0 by

0 if Ox = 0,
L(x,e)e:= { n
eEO 2:1 (x, ei)ei if Ox = {el' e2, . .. ,en} is finite.

In the case that Ox = {el' e2, . .. } is countably infinite, we define


n
L (x, e)e := li~ L (x, ei)ei,
eEO i=l

provided that this limit exists. (In the remark following Lemma 4.18, we will see
that this limit is independent of the order of the terms listed in Ox.) In each of
FOURlER ANALYSIS 57

these cases, we say that the Fourier series 2:eEO(x, e)e converges. Otherwise,
we say that the Fourier series diverges. The scalars (x, e) are called the Fourier
coefficients of x relative to O.
We will see below that the Fourier series 2:eEO (x, e)e always converges whenever
X is complete or, more generally, whenever the closed subspace M := spanO is
complete in the inner product space X. In fact, in either of these cases,

L (x, e)e = PM(X) for each x E X


eEO

(see Theorem 4.22).


The next result shows that the Fourier series of every element in spanO always
converges; in fact, it converges to the element itself.
4.18 Lemma. Let 0 be an orthonormal subset of the inner product space X.
Then for each x E spanO,

(4.18.1) x= L(x,e)e.
eEO

Proof. Fix any x E spanO. If Ox = 0, then (x, e) = 0 for every e E 0, i.e.,


x E 01. = (spanO) 1. . Since spanO n (spanO)1. = {O}, it follows that x = 0 and
(4.18.1) holds.
Next let Ox = {e1' e2, ... , en}. Given E > 0, choose y E spanO such that
Ilx - yll < E. Then y = 2::"=1 Ctksk for some Sk EO and Ctk E R By Theorem 4.14,

(4.18.2)

L CtkSkl12 = IIxl1 2 - 2 L + L Ct~


m m m

Ilx - Ctk(X, Sk)


1 1 1
m-1 m-1

;::: IIxl1 2 - 2 L Ctk(X, Sk) + L Ct~


1 1

Thus by replacing y by 2:;:,,-1 CtkSk, we see that (4.18.2) is still valid. By repeating
this argument, we see that we may discard those terms in the sums 2:;:" (x, Sk)Sk
and 2:;:" CtkSk such that (x, Sk) = 0, i.e., such that Sk tJ. Ox. That is, we may always
choose y E spanO x = span{e1' ... ' en}. For any such y,

Ilx - ~(x, ei)eill ~ Ilx - yll < E


58 CHARACTERlZATION OF BEST APPROXIMATIONS

by Theorem 4.14. Since E was arbitrary, x = L~(X, ei)ei.


Finally, suppose Ox = {e1' e2, ... } is countably infinite. Given E > 0, choose
y E spanO such that Ilx-yll < E. The same argument as in the preceding paragraph
shows that y may be chosen from spanOx . Thus

for n sufficiently large. It follows that x = limn L~(X, ei)ei. If the elements of Ox
are arranged in any other order, say Ox = {e~, e~, ... }, then, for any E > 0, choose
N1 such that Ilx - L~(x,ei)eill < E for every n 2: N 1. Now choose N2 2: N1 such
that {e1' e2, .. . ,eN'} C {e~ .e~, ... ,e~,}. Then for every n 2: N 2,

Hence limn L~(X, e~)e~ = x. In every case, we have shown that (4.18.1) holds. •
Remark. To verify the statement in parentheses that was made in Definition
4.17, suppose that y = limnL~(X,ei)ei exists. Then y E spanO, and by continuity
of the inner product we deduce that (y, e) = (x, e) for every e E O. Thus Ox =
Oy. By Lemma 4.18, y = limn L~(Y, e~)e~ = limn L~(X, eDe~ for any reordering
{el' e~, ... } of Ox = Oy.
While Lemma 4.18 shows that the Fourier series for x converges when x E spanO,
it will be seen below that this series often converges for every x EX. For example,
the next theorem shows that if X is complete (or, more generally, if spanO is
complete), then the Fourier series converges jor every x E X. Before formulating
the main theorem concerning Fourier series, we make the following observation. If
x = LeEO(x, e)e and y = LeEO(y, e)e, then

(4.18.3) (x, y) = L (x, e)(e, y),


eEO

where
n
(4.18.4) L(x,e)(e,y):= li~ L(x,ei)(ei,y)
eEO i=l

and {e1' e2, . .. } is any enumeration of the (countable) set Ox U Oy.


To see this, we first note that we can write x = limn L~(X, ei)ei and y =
limn L~ (y, ei)ei. By continuity of the inner product in each argument, we deduce

n m n
(x, y) = li~ L(x, ei) l~ L(ej, y)(ei' ej) = li~ L (x, ei)(ei, y).
i=l j=l i=l

Since the scalar (x, y) is independent of the enumeration in Ox U Oy, the result
follows.
FOURlER ANALYSIS 59

4.19 Fourier Series. Let 0 be an orthonormal set in an inner product space


X. Consider the following statements:
(1) spanO = X.
(2) Every element can be expanded in a Fourier series. That is,

x = L (x, e)e for each x E X.


eEO

(3) (Extended Parseval identity)

(x, y) = L (x, e) (e, y) for every x, y EX.


eEO

(4) (Parseval identity)

IIxl12 = L l(x,e)1 2 for every x E X.


eEO

(5) 0 is a "maximal" orthonormal set. That is, no orthonormal set in X prop-


erly contains O.
(6) 0.1. = {O}.
(7) (Elements are uniquely determined by their Fourier coefficients)

(x, e) = (y, e) for every eE 0 implies that x = y.


Then (1) {=> (2) {=> (3) {=> (4) =} (5) {=> (6) {=> (7).
Moreover, if spanO is complete (for example, if X is complete), then all seven
statements are equivalent.
Proof. (1) =} (2). This is a consequence of Lemma 4.18.
(2) =} (3). This was established in (4.18.3) and (4.18.4).
(3) =} (4). Fix any x E X and put y = x in (3) to get IIxl1 2 = 2::eEO l(x,e)iZ.
(4) =? (1). If (4) holds, let x E X and Ox = {el' e2""}' Then

n
= IIxl1 2- L I(x, eiW -+ 0 as n -+ 00.
1

This proves that x E spanO and hence that (1) holds.


(4) =? (5). If (5) fails, then 0 is not maximal, so we can choose an x E X \ 0
such that 0 U {x} is an orthonormal set. Then x E 0.1. implies that (x, e) = 0 for
every eE 0, so that 2::eEO I(x, e)lZ = 0, but IIxl1 2 > O. This shows that (4) fails.
(5) =? (6). If (6) fails, then there exists x E 0.1. \ {O}. Thus the set 0 U {x/llxll}
is an orthonormal set that is strictly larger that 0; hence 0 is not maximal, and
(5) fails.
(6) =? (7). If (x, e) = (y, e) for every e E 0, then x - y E 0.1. = {O}.
(7) =? (5). If (5) fails, 0 is not maximal, and we can choose an x E X such that
o U {x} is orthonormal. Then (x,e) = 0 = (0, e) for every eE 0; but Ilxll = 1, so
(7) fails.
60 CHARACTERIZATION OF BEST APPROXIMATIONS

Finally, assume that spanO is complete. If (1) fails, choose x E X \ spanO. Since
M := spanO is complete, it follows by Theorems 3.4 and 4.9 that M is a Chebyshev
subspace and 0 # x - PM (x) E M.L c O.L. Thus (6) fails. This proves that (6)
implies (1), and hence verifies the last statement of the theorem. •

Since each closed subspace M of an inner product space is an inner product


space in its own right, we can apply Theorem 4.19 to M instead of X to obtain the
following corollary.

4.20 Corollary. Let M be a closed subspace of the inner product space X and
let a be an orthonormal subset of M. Consider the following statements.
(1) spanO = M.
(2) x = Lew (x, e)e for every x E M.
(3) (x, y) = LeEO(x, e)(e, y) for every x, y E M.
(4) IIxl1 2 = LeEO I(x, e)12 for every x E M.
(5) a is a maximal orthonormal subset of M.
M n O.L = {OJ.
(6)
x, y E M and (x, e) = (y, e) for every e E a implies x = y.
(7)
Then (1) {=} (2) {=} (3) {=} (4) =? (5) {=} (6) {=} (7).
Moreover, if M is complete, then all seven statements are equivalent.

Remark. An orthonormal set a with the property that M = spanO is often


called an orthonormal basis for M. By the preceding corollary, a is an orthonor-
mal basis for M if and only if x = LeEO(x, e)e for every x E M. Clearly, when a
is finite, this is just the usual definition of (orthonormal) basis for M. It is natural
to ask what closed subspaces actually have orthonormal bases. By the Gram-
Schmidt theorem, every finite-dimensional subspace has an orthonormal basis. We
can show, more generally, that every complete subspace of an inner product space
has an orthonormal basis. By the above corollary, it suffices to show that every
inner product space contains a maximal orthonormal set.

4.21 Existence of Maximal Orthonormal Subsets. Every inner product


space X # {OJ contains a maximal orthonormal set.
Proof. The proof uses Zorn's lemma (see Appendix 1). Let C denote the
collection of all orthonormal subsets of X. Now, C # 0, since {x/llxll} E C for any
nonzero x E X. Order C by containment: 0 1 >- O 2 if and only if 0 1 ~ O 2 . Let
T be any totally ordered subset of C. It is clear that u{ a I a E T} is an upper
bound for T. By Zorn's lemma, C contains a maximal element 0, and this must
be a maximal orthonormal set in X. •

4.22 Approximating from Infinite-Dimensional Subspaces. Let M be a


complete subspace of the inner product space X. Then:
(1) M is a Chebyshev subspace.
(2) M has an orthonormal basis.
(3) If a is any orthonormal basis for M, then

(4.22.1) PM(x) = L (x, e)e for every x E X.


eEO
SOLUTIONS TO THE FIRST THREE BASIC PROBLEMS 61

Proof. (1). By Theorem 3.4(2), M is a Chebyshev subspace.


(2). By Corollary 4.20 and Theorem 4.21, M has an orthonormal basis.
(3). For any x E X, let Xo = PM(X). By Corollary 4.20,

(4.22.2) Xo = L (xo, e)e.


eEO

Since x - Xo E Ml. by Theorem 4.9 and since M = spanO, it follows that (x-
Xo, e) = 0 for every e E O. That is, (x, e) = (xo, e) for every e E O. Substitute this
into (4.22.2) to obtain (4.22.1). •
The preceding theorem shows that if an orthonormal basis for a complete sub-
space M is available, then it is easy (in principle) to compute best approximations
from M by using formula (4.22.1). Now we give an application of this theorem.
4.23 Application. Let r be a nonempty set and X = £2(r). For each j E r,
define ej E X by

That is, ej is the function in X that is 1 at j and 0 elsewhere. Fix any non empty
subset ro c r. Then the subspace

Mro := {x E X I x(j) = 0 for every j E r \ ro}

is a Chebyshev subspace of X,

is an orthonormal basis for M ro ' and

(4.23.1) PMro (x) = L (x, e)e = L x(j)ej for every x E X.


eEOro jEro

In particular, when ro = r, we have that Or = {ej I j E r} and Mr = X. That is,


Or is an orthonormal basis for X.
To see this, first observe that Oro is clearly an orthonormal subset of Mro, and
Mro is a closed, hence complete, subspace of the Hilbert space X. Next note that
if x E Mro, then (x,ej) = xU) = 0 for every j E r \ r o, while if x E 0&0' then
xU) = (x, ej) = 0 for every j E roo Thus if x E Mro n 0&0' it follows that x(j) = 0
for every j E r, or x = O. That is, Mro not = {O}. By Corollary 4.20, Oro is an
orthonormal basis for Mro. Formula (4.23.1) now follows from Theorem 4.22.

Solutions to the First Three Basic Problems


We conclude this chapter by giving complete solutions to the first three of the
five problems initially posed in Chapter 1. The fourth (respectively fifth) problem
will be solved in Chapter 7 (respectively Chapter 10).
62 CHARACTERIZATION OF BEST APPROXIMATIONS

Problem 1. (Best least-squares polynomial approximation to data) Let


{( tj , x( tj)) I j = 1,2, ... , m} be a table of data. For any fixed integer n < m, find
a polynomial p(t) = L~ aiti, of degree at most n, such that the expression
m

k=l
is minimized.
We saw in Chapter 2 that letting T = {tl, t2, ... , t m }, X = l2(T), and M = P n ,
the problem may be restated as follows: Find the best approximation to x E X
from the (n + I)-dimensional subspace M.
By Corollary 4.10, p(t) = L~ aiti is the best approximation to x if and only if
the ai satisfy the normal equations
n
Lai(ti,tj) = (x,t j ) (j = 0, 1, ... , n) ,
i=O
where
m m

(ti, t j ) = L4t{ = Lti+ j


k=l k=l
and
m

(x, t j ) = LX(tk)t{ .
k=l
Hence the normal equations can be written as

(1.1) ~ai (~t~+j) = ~t{X(tk) (j=O,I, ... ,n) .

In particular, when we are seeking the best constant approximation to x (i.e., n = 0),
the normal equations (1.1) reduce to
m m

aOL 1 = LX(tk),
k=l k=l
or
1 m
aD = mLX(tk) .
k=l
That is, the best constant approximation to x is its mean.
Similarly, the best linear approximation to x (i.e., n = 1) is given by Pl(t) =
aD + alt, where the ai are given by

(~ t~) (~X(tk)) - (~tk) (~tkX(tk))


ao=
m (~t~) - (~tkr
and

al =
m (~tkX(tk)) - (~tk) (~X(tk))
m (~tf) - (~tk r
SOLUTIONS TO THE FIRST THREE BASIC PROBLEMS 63

Problem 2. (Solution to an overdetermined system of equations) Consider the


linear system of m equations in the n unknowns Xl,X2,··· ,Xn:

or briefly, Ax = b. Find a vector x = (Xl, X2, ... , x n ) E JRn that minimizes the
expression

We saw in Chapter 2 that letting X = l2(m) and M = {y E X I y = Ax,


x E JRn}, the problem may be restated as follows: Find the best approximation Yo
to b from the subspace M. Then choose any Xo E JRn such that Axo = Yo.
By Theorem 4.9, Yo = PM(b) if and only ifb-yo E M.l.. if and only if (b-yo, Ax) =
o for every x E JRn . At this point we make the following observation. If A*
denotes the transpose matrix of A (i.e., the (i, j) entry of A* is the (j, i) entry of
A : aTj = aji), then, for each x E JRn and y E JRm, we have

(where the inner product on the left corresponds to the space b(m), while that on
the right corresponds to the space b(n)).
Continuing from the top of the preceding paragraph, we see that Yo = PM(b) is
equivalent to (A*(b-yo), x) = 0 for every x E JRn, which is equivalent to A*(b-yo) =
0; i.e., A*b = A*yo. Now, Yo EM, and so Yo = Axo for some Xo E JR n . Thus we
can finally conclude that Xo E JRn is a solution to the problem if and only if

(2.1) A* Axo = A*b.


In particular, if the matrix A* A is nonsingular, there is a unique solution of (2.1)
given by

(2.2) Xo = (A* A)-l A*b.

However, if A* A is singular, there will always be more than one solution Xo of (2.1).
We should emphasize that this does not contradict the uniqueness guaranteed by
Theorem 2.4. For although the best approximation Yo E M to b is unique (by
Theorem 2.4), there may be more than one Xo E JRn with Axo = Yo. That is, A
may not be one-to-one.
Problem 3. (Best least-squares polynomial approximation to a function) Let
x be a real continuous function on the interval [a, bJ. Find a polynomial p(t)
L:~ a4i, of degree at most n, that minimizes the expression
64 CHARACTERIZATION OF BEST APPROXIMATIONS

Letting X = C2 [a, b] and M = P n , the problem is to find the best approxima-


tion to x from the subspace M. By Corollary 4.10, pet) = ~~ aiti is the best
approximation to x if and only if the ai satisfy the normal equations

(j=O,l, ... ,n),

or equivalently,

(3.1) L.
n

i=O l
~i
+J + 1
(bi+J+l _ ai+J+l) = r tjx(t)dt
ia
b
(j = 0, 1, ... , n) .

In particular, the best constant approximation to x (i.e., n = 0) is given by its


integral mean, or average:

iar x(t)dt.
1 b
ao = b_ a

Exercises

1. (Approximation by translates of cones). Let C be a convex cone,


z E X, and K = C + z. Let x E X and Yo E K. Verify that Yo = PK(X) if
and only if (x - Yo, y) ::::; 0 for all y E C and (x - Yo, Yo - z) = O.
2. (Approximation by affine sets, i.e., translates of subspaces). Let
M be a subspace, v E X, and V = M + v.
(a) Let x E X and Yo E V. Show that Yo = Pv(x) if and only if
x - Yo E Ml..
(b) Prove that Pv(x + z) = Pv(x) for every x E X and z E Ml..
3. (Strong uniqueness) Let K be a convex Chebyshev set in X. Prove that
for any x E X,

for every y E K.
This is a type of strong uniqueness result for best approximations: It gives
a quantitative estimate of how much larger Ilx - yll is than Ilx - PK(x)11 in
terms of Ily - PK(x)ll. [Hint: Theorem 4.1.]
4. Let C = {p E C2 [a,b] I p E P n , p 2': O}.
(a) Show that C is a Chebyshev convex cone.
(b) When n = 0, show that for every x E C2 [a, b],

Pc(x) = max { 0, b ~a lb X(t)dt}

(c) If n 2': 1, can you exhibit a formula for Pc(x)?


EXERCISES 65

5. (a) Let C = {x E l2(1) I x(i) ::::: 0 for all i E I}. Show that C is a Chebyshev
convex cone in l2 (1) and
Pc(x) = x+ := max{x, O}

*
for every x E l2(1).
(b) Let C = {x E L 2 (J.L) I x ::::: O}. Show that C is a Chebyshev convex
cone in L 2(J.L) and Pc(x) = x+ for each x E L 2(J.L). (This generalizes part
(a).)
6. (Distance to finite-dimensional subspace). Let {XI,X2, ... ,xn } be
linearly independent in X and M = span{xI, X2, . .. , x n }. Show that

d(x M)2 = g(X,XI,X2,""Xn)


, g(XI,X2,""X n )
for every x E X, where g(yl, Y2, ... , Ym) is the determinant of the m x m
Gram matrix
(YI, YI! (YI, Y2!
[ (Y2, YI! (Y2, Y2!
(YI, Ym!
(Y2, Ym!
1
.

(Ym, YI! (Ym, Y2! (Ym,Ym)

[Hint: Adjoin the equation d(x,M)2 = Ilx - P M (x)11 2 = (x - PM(x),x) =


(x,x)- L:~ Oi(Xi,X) to the normal equations (4.10.2) and solve for d(x,M)2
by Cramer's rule.]
7. (Gram determinants). Let {Xl, X2,"" xn} be a set of n vectors in X
and g(XI' X2,' .. ,xn ) the Gram determinant as defined in Exercise 6. Verify
the following statements.
(a) g( Xl, X2, ... ,xn ) is a symmetric function of the n arguments Xi·
[Hint: What happens when two of the Xi'S are interchanged?]
(b) 0 :::; g(XI' X2, ... , Xn) :::; IIxIil211x2112 ... Ilxnl12.
[Hint: Exercise 6.]
(c) The equality g( Xl, X2, ... , Xn) = 0 holds if and only if the set {Xl, ... , xn}
is linearly dependent.
(d) The equality g(XI,X2,'" ,xn) = IlxIil21Ix2112 .. '11xn 11 2 holds if and only
if {Xl, X2,' .. , Xn} is an orthogonal set.
(e) The inequality g(XI,X2) ::::: 0 is just Schwarz's inequality. (Hence, the
inequality on the left in part (b) generalizes Schwarz's inequality.)
8. (Hadamard's determinant inequality). Let A be an n x n real matrix:

A= [:: : : : : : 1
...
anl an2 ... ann
, or A = [::1
. ..
an
,

where ai = (ail, ai2, ... ,ain) denotes the ith row vector of A. We regard
each ai as an element in l2 (n).
(a) Show that
66 CHARACTERIZATION OF BEST APPROXIMATIONS

[Hint: The matrix AA* is a Gram matrix.J


(b) If laijl::; c for each i,j, then

9. Let {Xl,X2, ... ,Xn } be a basis for the subspace M. Show that for each
XEX,

X Xl xn
1
(XI,X) (Xl,Xl) (Xl,Xn)
FM(X) =X- (X2,X) (X2,XI) (X2,X n )
g(Xl, X2,·· . , Xn)
(Xn,X) (Xn, Xl) (Xn,Xn)

(The determinant on the right is, of course, understood to be the linear


combination of X, Xl, ... , Xn that one obtains by formally expanding this
determinant by cofactors of the first row.)
10. (Cauchy determinant formula). Let ai, bi be real numbers such that
ai + bj =f 0 for all i,j = 1,2, ... , n. Show that if

I I 1
Ql+b1 al+b2 Ql+bn
1 1 1
D= Q2+bl a2+b2 Q2+bn

I 1 1
Qn+bl Qn+b2 an,+bn

then

11. In the space C2[O, IJ, consider the monomials Xi(t) = t i - l (i = 1,2, ... , n).
(a) Show that (Xi,Xj) = (i + j _1)-1.
(b) In this case the Gram matrix G(X1,X2,'" ,xn) is also called the Hilbert
matrix. Show that the determinant of the Hilbert matrix is given by

[1!2!··· (n - 1)!J4
g(XI,X2, ... ,Xn )= 1!2!···(2n-l)!·

[Hint: Exercise 10 with ai = bi = i - ~.J


(c) Verify that

What practical implications does this have for solving the normal equations
when n is large?
12. Find the best approximation to X E C2 [-I, IJ from the subspace P2 in the
cases where
(a) x(t) = t 3 ,
(b) x(t) = et ,
(c) x(t) = sin 27rt.
EXERCISES 67

13. (a) Compute the first four Legendre polynomials (see Example 4.12 (1)).
(b) Compute the first four Chebyshev polynomials of the first kind (see
Example 4.12 (2)).
(c) Compute the first three orthonormal polynomials in b(m), m::::: 4 (see
Example 4.12 (3)).
14. In the weighted space C2([a, b]; w), let {PO,PI,P2, ... } denote the set of poly-
nomials obtained by orthonormalizing the set of monomials {I, t, t 2 , •.• }.
Show that the best approximation to t n from Pn - l is given by qn-l(t) :=
t n - anPn(t), where an is a constant chosen such that

anPn(t) = t n + lower-order terms.

15. Show that the zeros of the orthonormal polynomials obtained in Exercise
14 are real, simple, and lie in [a, b].
16. Prove that a finite-dimensional inner product space is complete by using
the fact that it has an orthonormal basis.
17. If {Xl, X2, ... , Xn} is an orthonormal basis for the subspace M, show that
n
X= L (x, Xi)Xi
I

for every x E M.
18. (a) Verify that for a table of "constant data," i.e., {(tj,e) I j = 1, ... ,m},
both the best constant approximation and the best linear approximation to
this data are given by p(t) == e. What about the best approximation from
Pnwhenn> I?
(b) Verify that for a table of "linear data," i.e., {( tj, atj + b) I j = 1, ... , m},
the best constant approximation is given by

Po(t) = a (~~tk) + b,
and the best linear approximation is given by

What about the best approximation from P n when n > I?


19. Let {XI,X2, ... ,xn } C X and let K = {L:~ .\Xi I Ai ::::: 0, L:~ Ai = I}.
(That is, K is the "convex hull" of {Xl,X2, ... ,xn }.) Let Yo E K. Show
that Yo = PK(O) if and only if

(i=I,2, ... ,n).

20. An alternative approach to obtaining characterization theorems for best


approximations is via differentiation. For example, finding the best approx-
imation to x from the subspace M = span{xl,x2, ... ,x n } is equivalent to
minimizing the function of n real variables
68 CHARACTERIZATION OF BEST APPROXIMATIONS

By expanding this in terms of inner products, show that f is a differentiable


function of the ai. Deduce the normal equations (4.10.2) again from the
conditions

(20.1) (i=1,2, ... ,n).

Can you justify why the necessary condition (20.1) is also sufficient?
[Hint: When L~ aT is large, so is f(a1, a2, . .. , an).] Finally, the solution
to (20.1) is unique. This can be established using the nonsingularity of the
Gram matrix G(X1,X2,."'X n ) (see Exercise 7), or it can be verified by
showing that f is a "strictly convex" function, and such functions have a
unique minimum.
21. For any collection of nonempty sets {Sl, S2, ... , Sm} in an inner product
space, show that (nr Si)O ~ L7' Sf·
22. In Theorems 4.5 and 4.6, prove the statements concerning orthogonal com-
plements.
23. (Infinite sum of sets) Suppose I is any index set, and for each i E I, let
Si be a nonempty subset of the inner product space X such that 0 E Si for
all except possibly finitely many i E I. We define the sum of the sets Si,
denoted by LiEI Si, by

{ LSi I Si E Si for all i E I, Si = 0 for all except finitely many


iEI
i} .
(Note that if I is finite, this reduces to the usual definition of the sum of
a finite collection of sets.) Prove a generalization of Theorem 4.6 valid for
any indexed collection of nonempty sets, not necessarily a finite collection.
24. Let X be a Hilbert space and C and D closed convex subsets with 0 E CnD.
Show that the following statements are equivalent.
(1) (CnD)o=co+DO;
(2) (C n D)O c Co + DO;
(3) con (C) n con(D) c --'co-n-;-(C=--n-:OD"');
(4) con(C) n con(D) = con(C n D).
While Theorem 4.6(3) shows that (1) (and hence all four statements above)
hold when C and D are Chebyshev convex cones, the next exercise shows
that these conditions do not always hold for general Chebyshev convex sets.
25. Consider the two sets in the Hilbert space £2(2) defined by

C:= {x E £2(2) I x 2 (1) + [x(2) - If <:::: 1} and D:= -C.

(a) Verify that both C and D are convex Chebyshev sets.


(b) Show that C n D = {a}, COO(C) = {x E £2(2) I x(2) ::>: a}, ron(D) =
{x E £2(2) I X(2) <:::: a}, con(C n D) = {O}, and con(C) n con(D) =
{x E £2(2) I x(2) = O}.
(c) Show that (C n D)O f Co + DO. [Hint: Either a direct proof or appeal
to (b) and Exercise 24.] This proves that Theorem 4.6(3) is not valid if the
HISTORICAL NOTES 69

Chebyshev convex cones are replaced by more general convex Chebyshev


sets.
26. Let M and N denote the following subsets of £2(4): M = span{el,e2 +
e3} and N = span{el +e2,e3 +e4}, where {el,e2,e3,e4} is the canonical
orthonormal basis for £2(4), i.e., eiU) = t5ij . Verify the following statements.
(a) M and N are 2-dimensional Chebyshev subspaces of £2(4).
(b) M n N = {O}.
(c) For each element x = (x(1),x(2),x(3),x(4)) E £2(4), we have

1
PM(x) = x(l)el + 2" [x(2) + x(3)](e2 + e3)
and

27. Let X = £2 and let {en I n = 1, 2, ... } denote the canonical orthonormal
basis of X, i.e., en(i) = t5ni for all positive integers nand i. Let Mo =
span{ e2n-l In = 1,2, ... } and Me = span{ e2n In = 1,2, ... }. Show that
Mo and Me are Chebyshev subspaces, Me = Me;, Mo + Me = £2, PMJX) =
L~ x(2n - 1)e2n-l, and PMe(x) = L~ x(2n)e2n for all x E £2.
28. An inner product space X is called separable if it contains a countable
dense subset D, that is, if D = X.
(a) If D is countable and spanD = X, show that X is separable. [Hint: The
subset of spanD consisting of those linear combinations whose coefficients
are all rational numbers is countable and dense in spanD.]
(b) Every finite-dimensional inner product space X is separable.
(c) Show that every orthonormal subset of a separable inner product space
is countable. [Hint: If e and e' are distinct members of an orthonormal set,
then lie - e/ll = .J2.]
29. Let X be a complex inner product space. Verify the following statements.
(a) The characterization theorem (Theorem 4.1) is valid if the inequality
(4.1.1) is replaced by

(4.1.1') Re (x - Yo, Y - YO! ~ 0 for every y E K .

(b) The characterization theorem (Theorem 4.3) is valid if the dual cone of
a set S is defined by So := {x E X I Re (x, y! ~ 0 for every YES}.
(c) Theorems 4.5, 4.6, 4.9, 4.10,4.11,4.16, 4.18, 4.19, 4.20, 4.21, and 4.22
are valid as stated.

Historical Notes
The characterization of best approximations (Theorem 4.1) can be traced back
at least to Aronszajn (1950a) (see also Cheney and Goldstein (1959)), although
it may be older. The characterization of best approximations from convex cones
(Theorem 4.7), which is a strengthening of Theorem 4.1 in the special case of cones,
is essentially contained in Moreau (1962).
The conical hull and dual cone of a set and many of their basic properties can be
found in Steinitz (1914) (see also Fenchel (1953)). Theorem 4.4(5) can be found in
70 CHARACTERIZATION OF BEST APPROXIMATIONS

Rockafellar (1970; Corollary 2.6.3, p. 14). Dual cones can be defined even in more
general normed linear spaces, but in this case they must lie in the dual space. See
Kato (1984) for some facts about dual cones in normed linear spaces.
In 1883, Gram (1883) generalized a problem of Chebyshev for finite sums by
considering the problem of minimizing the integral

(4.25.1)

over all possible choices of the scalars aj. By cleverly applying the "orthogonaliza-
tion process" to the 'l/Jj's (now recognized as the forerunner of the Gram-Schmidt
orthogonalization procedure, Theorem 4.11), he was able to reduce the problem
to the case where the 'l/J/s form an orthonormal set, and the aj's that minimize
(4.25.1) are the Fourier coefficients aj = J: p(t)f (t)'l/Jj (t) dt. Gram also considered
an infinite orthonormal set {'l/Jl, 'l/J2, ... } and was able to see that the condition that
the expression in (4.25.1) converged to zero as n tends to infinity was connected to
the maximality of the orthonormal set {'l/Jl, 'l/J2, ... }.
The fact that all orthonormal bases of a Hilbert space have the same cardinality
was shown by L6wig (1934; p. 31) and Rellich (1935; p. 355). L6wig (1934; p. 27)
proved that two Hilbert spaces are "equivalent" if and only if they have the same
dimension. (The dimension of a Hilbert space is the cardinality of a maximal
orthonormal basis. See Exercise 4 of Chapter 6 for the definition of "equivalent.")
The equivalence of the first three statements of Theorem 4.19, when X is a
separable Hilbert space, was proved by von Neumann (1929a; Satz 7).
A constructive approach to computing the best approximation PM(x) as given
in (4.22.1) was given by Davis, Mallat, and Zhang (1994). (See also Cheney and
Light (2000).)
CHAPTER 5

THE METRIC PROJECTION

Metric Projections onto Convex Sets


In this chapter we shall study the various properties of the metric projection onto
a convex Chebyshev set K. It is always true that PK is nonexpansive and, if K is
a subspace, even linear. There are a substantial number of useful properties that
PK possesses when K is a subspace or a convex cone. For example, every inner
product space is the direct sum of any Chebyshev subspace and its orthogonal
complement. More generally, a useful duality relation holds between the metric
projections onto a Chebyshev convex cone and onto its dual cone. The practical
advantage of such a relationship is that determining best approximations from a
convex cone is equivalent to determining them from the dual cone. The latter
problem is often more tractable than the former. Finally, we record a reduction
principle that allows us to replace one approximation problem by another one that
is often simpler.
We begin by first observing the equivalence of the "E - 5" definition of continuity
with the sequential definition.
5.1 Lemma. Let X and Y be nonempty subsets of inner product spaces, F a
mapping from X into Y, and Xo EX. Then the following statements are equivalent:
(1) F(xn) -+ F(xo) whenever Xn E X and Xn -+ Xo;
(2) For each E > 0, there exists 5 = 5(E, xo) > 0 such that IIF(x) - F(xo) I < E
whenever x E X and Ilx - xoll < 6.
Proof. (1) =? (2). Suppose (2) fails. Then there is an E > 0 such that for each
integer n ::::: 1, there exists Xn E X with Ilxn - Xo II < lin and IIF(xn ) - F(xo) II : : : E.
Then Xn -+ Xo but F(x n ) f+ F(xo). Thus (1) fails.
(2) =? (1). Suppose (2) holds, Xn E X, and Xn -+ Xo. For each E > 0, choose
5> 0 such that IIF(x) - F(xo)11 < E whenever x E X and Ilx - xoll < 5. Choose
an integer N such that Ilxn - xoll < 5 for n::::: N. Then IIF(xn ) - F(xo)11 < E for
n::::: N. That is, F(x n ) -+ F(xo), and (1) holds. •
5.2 Definition. Let X and Y be (nonempty subsets of) inner product spaces,
F: X -+ Y, and Xo EX. F is said to be continuous at Xo if both the equivalent
conditions (1) and (2) of Lemma 5.1 are satisfied. F is called (i) continuous (on
X) if it is continuous at each point of X; (ii) uniformly continuous (on X) if
for each E > 0, there exists 5 = 5(E) > 0 such that IIF(x) - F(y)11 < E whenever
x, y E X and Ilx - yll < 5; (iii) Lipschitz continuous if there exists a constant
s
c::::: 0 such that IIF(x) - F(y)11 cllx - yll for all x, y in X; (iv) nonexpansive if
IIF(x) - F(y)11 s Ilx - yll for all x, y E X.
Clearly, every nonexpansive mapping is Lipschitz continuous, every Lipschitz
continuous mapping is uniformly continuous, and every uniformly continuous map-

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
72 THE METRIC PROJECTION

ping is continuous. Let us next observe that the distance functional F(x) = d(x, K)
is nonexpansive.
5.3 Distance is Nonexpansive. Let K be a nonempty subset of X. Then for
every pair x, y in X,

(5.3.1) Id(x, K) - dey, K)I :::: Ilx - YII·


In particular, the function x t-+ d(x, K) is nonexpansive, hence uniformly con-
tinuous.
Proof. For any z E K,

d(x, K) :::: Ilx - zll :::: Ilx - yll + Ily - zll·


Taking the infimum over all z E K yields

d(x, K) :::: Ilx - yll + dey, K).


Combining this with the inequality obtained by interchanging the roles of x and y,
we obtain the result. •
While a continuous function is not generally uniformly continuous (e.g., J(t) =
lit on (0, 1)), this will be the case when the domain is compact.
5.4 Theorem. Let X be a compact subset of an inner product space, Y an
inner product space, and suppose F: X --'t Y is continuous. Then F is uniformly
continuous.
Proof. If not, then there exists E > ° such that for each n E N there are points
xn'x~ E X with Ilxn - x~11 < lin and
(5.4.1 )

Since X is compact, there is a subsequence {x nk } and Xo E X such that x nk --'t Xo.


It follows that X~k --'t Xo also. Since F is continuous at xo, there exists Ii > Osuch
that 1!F(x)-F(xo)11 < E/2 whenever Ilx-xoll < Ii. Then for k large, Ilxnk -xoll < Ii
and Ilx~k - xoll < Ii, so that

IIF(xnk ) - F(x~J I :::: 1!F(xnk ) - F(xo) II + IIF(xo) - F(X~k)ll < E,

which contradicts (5.4.1). •


Some useful properties of the metric projection onto a convex Chebyshev set are
recorded next.
5.5 Metric Projections are Monotone and Nonexpansive. Let K be a
convex Chebyshev set. Then:
(1) PK is "idempotent":

(5.5.1)

Briefly, P'i< = PK .
METRIC PROJECTIONS ONTO CONVEX SETS 73

(2) PK is "firmly nonexpansive":

(5.5.2)

(3) P K is "monotone":

(5.5.3) (x - Y, PK(X) - PK(Y)) 2': 0 for all x, y.

(4) PK is "strictly nonexpansive":

(5) PK is nonexpansive:

(5.5.5) IIPdx) - PK(y)11 :::; Ilx - YII for all x, y.

(6) P K is uniformly continuous.


Proof. (1) This follows because each Y E K is its own best approximation in
K: PK(Y) = y.
(2) We have

(x - Y, PK(x) - PK(y)) = (x - PK(X), PK(x) - PK(y))


+ (PK(X) - PK(y),PK(X) - PK(Y))
+ (PK(y) - Y, Pdx) - PK(Y))'

The first and third terms on the right are nonnegative by Theorem 4.1, and the
second term is IIPK(x) - PK (y)11 2 . This verifies (2).
(3) This is an immediate consequence of (2).
(4) Using (2), we obtain for each x, Y in X that

Ilx - YI12 = II[x - PK(X)] + [PK(x) - PK(Y)] + [PK(Y) - Y]112


= IIPK(X) - PK (y)11 2 + II[x PK(X)]- [y - PK (y)]11 2
+ 2(PK(X) - PK(y),x - PK(x) - [y - PK(Y)])
= IlPdx) - PK (y)11 2 + Ilx - PK(X) - [y - PK (y)]11 2
+ 2(PK(X) - PK(Y)' X - y) - 21IPK (x) - PK (y)11 2
2': IIPK(X) - PK (y)11 2 + Ilx - PK(X) - [y - PK (y)]11 2

This proves (4).


Finally, note that (5) follows immediately from (4), and (6) from (5). •
In practice, the element x being approximated may not always be specified ex-
actly (because of round-off, truncation, or other types of error). Thus, instead of
approximating the "ideal" element x, we will actually be approximating the ele-
ment x + ox, where ox is the error vector involved. The inequality (5.5.5) (see also
Figure 5.5.6) implies that IIPK(x+ox) - PK(x)11 :::; Iloxll. In other words, the error
incurred in the best approximation is no larger than the error in the original vector.
This is obviously a desirable property.
74 THE METRIC PROJECTION

.x

....
.y

Figure 5.5.6

The metric projection onto a Chebyshev convex cone or Chebyshev subspace has
additional, and stronger, properties.
5.6 Metric Projections onto Convex Cones. Let C be a Chebyshev convex
cone in the inner product space X (e.g., a closed convex cone in a Hilbert space).
Then:
(1) Co is a Chebyshev convex cone.
(2) For each x E X,

x = Pc(x) + Pco(x) and Pc(x).l Pco(x).

Moreover, this representation is unique in the sense that if x = Y + z for some y E C


and z E Co with y.l z, then y = Pc(x) and z = Pco(x).
(3) IIxl1 2 = IIPc(x)112 + IIPco(x)112 for all x E X. Hence, IIxl1 2 = d(x, c)2 +
d(x, co)2.
(4) Co = {x E X I Pc(x) = O} and C = {x E X I Pco(x) = O} = {x E X I
Pc(x) = x}.
(5) IIPc(x)II :::; Ilxll for every x E X; moreover, IIPc(x)1I = Ilxll if and only if
xE C.
(6) Coo = C.
(7) Pc is ''positively homogeneous":

Pc()..x) = )..Pc(x) for all x E X, )..:0:: O.

Proof. To prove (1) and (2), let x E X and Co = x - Pc(x). By Theorem 4.7,
Co E Co and Co .1 (x - co). For every y E Co,

(x - Co, y) = (Pc(x), y) :::; O.

Hence x - Co E (CO) 0. By Theorem 4.7 (applied to Co rather than C), we get that
Co = Pcc (x). This proves that Co is a Chebyshev convex cone, x = Pc(x) +Pco (x),
METRIC PROJECTIONS ONTO CONVEX SETS 75

and Pco(x) J.. Pc(x). To complete the proof of (2), we must verify the uniqueness
of this representation.
Let x = y + z, where y E C, Z E Co, and y J.. z. For each c E C,

(x - y,c) = (z,c):::: 0

and

(x - y, y) = (z, y) = o.
By Theorem 4.7, y = Pc(x). Similarly, z = Pco(x).
(3) The first statement follows from (2) and the Pythagorean theorem. Moreover,
using this and (2), we obtain

which verifies the last statement.


(4) This follows using (2). For example, x E Co if and only if x = Pco (x) if and
only if Pc(x) = o.
(5) From (3), II Pc (x) II :::: Ilxll is clear. Also from (3), IIPC(x) II = Ilxll if and only
if Pco(x) = 0, which from (4) is equivalent to x E C.
(6) This was already verified in Theorem 4.5. It also follows by applying (4) to
the Chebyshev convex cone Co rather than C. That is,

Coo = (COr = {x E X I Pco(x) = O} = C.

(7) Let x E X and>" :::: o. Then by (2), x = Pc (x) + Pco (x) and >..x = >..Pc(x) +
>"Pco(x). Since both C and Co are convex cones, >..Pc(x) E C and >..Pco(x) E Co.
By (2) and the uniqueness of the representation for >..x, we see that Pc(>..x) =
>..Pc(x). •

5.1 Remark. The representation 5.6(2) can be also expressed by the operator
equation

(5.7.1) [= Pc +Pco,

where [ is the identity mapping: [(x) = x for all x. In particular, this relationship
implies that to find the best approximation to x from either set C or Co, it is
enough to find the best approximation to x from anyone of the sets. What this
means in practice is that we can compute whichever of the best approximations
Pc(x) or Pco(x) is more tractable; the other one is then automatically obtained
from the relation x = Pc(x) + Pco(x) (see Figure 5.7.2). This "duality" idea will
be more fully exploited in later chapters.
76 AX

Figure 5.7.2
Since MO = M -L when M is a subspace, we obtain the following corollary of
Theorem 5.6.
5.8 Metric Projections onto Subspaces. Let M be a Chebyshev subspace
of the inner product space X (e.g., a closed subspace of a Hilbert space). Then:
(1) M-L is a Chebyshev subspace.
(2) x = PM (x) + P M "- (x) for each x E X. Briefly, I = PM + P M"-.
Moreover, this representation is unique in the sense that if x = y + Z, where
y E M and Z E M-L, then y = PM (x) and Z = PM"-(X).
(3) IIxl1 2 = IIPM(X)112 + IIPM"-(x)112 for all x. Hence, IIxl1 2 = d(x, M)2 +
d(x,M-L)2.
(4) M-L = {x E X I PM(x) = O} and M = {x E X I PM"-(X) = O} = {x E X I
PM(x) = x}.
(5) IIPM(X)II ::s; Ilxll for all x EX; IIPM(X)II = Ilxll if and only if x E M.
(6) MH =M.
Just as in Remark 5.7, property 5.8(2) allows us to choose whichever one of
PM(X) or PM"-(x) is the "easiest" to compute, and the other is automatically
obtained. For example, suppose M were infinite-dimensional, but M -L was finite-
dimensional. In principle, it is easy to compute PM"-(X). This amounts to solving
the normal equations of Theorem 4.10 (for M-L instead of M) to obtain PM"-(x),
and hence also PM(x) = x - PM"-(X), This "duality" idea will also be pursued in
more detail in later chapters.
If A and Bare nonempty subsets of the inner product space X, we say that X
is the orthogonal SUIll of A and B, and write X = A 83 B, if each x E X has a
unique representation of the form x = a + b, where a E A, b E B, and a J. b. We
say that X is the direct SUIll of A and B, and write X = A EB B, if each x E X
has a unique representation in the form x = a + b, where a E A and b E B. Note
that if A and B are actually subspaces, then X = A EB B if and only if X = A + B
and An B = {O}.
LINEAR METRIC PROJECTIONS 77

There are simple examples in [2(2) of a closed convex cone C such that [2(2) =
CEBCo, but b(2) =I CEBCo. (For example, take C = {x E b(2) I x(l) 2:: O}.) Also,
it is easy to see that if M and N are any pair of distinct I-dimensional subspaces of
b(2) that are not orthogonal, then [2(2) = MEBN, but b(2) =I MEBN. In contrast
to these examples, if B c A~, then X = A EB B if and only if X = A EB B.
Using this terminology, we can now state a generalization for part (2) of both
Theorems 5.6 and 5.S. This result actually characterizes the convex cones or sub-
spaces that are Chebyshev.
5.9 The Projection Theorem. Let M (respectively C) be a subspace (re-
spectivelya convex cone) in the inner product space X. Then M (respectively C)
is Chebyshev if and only if

(5.9.1) X = M EB M.l (respectively X = C EB CO).


In particular, if M (respectively C) is a closed subspace (respectively closed
convex cone) in a Hilbert space X, then (5.9.1) holds.
Proof. The last statement of the theorem follows from the first and Theorem
3.5.
If C is Chebyshev, then (5.9.1) follows from Theorem 5.6 (2).
Conversely, suppose X = C EB Co. Then for each x EX, there exists a unique
c E C and Co E Co such that x = c+co and c -L co. Hence x-c E Co and x-c -L c.
By Theorem 4.7, c = Pc(x). Hence C is Chebyshev.
Similarly, X = M EB M.l if and only if M is Chebyshev. •
Note that the proof also showed that if X = C EB Co, then x = c + co, where
c= Pc(x) and Co = Pco(x).

Linear Metric Projections


Actually, further useful properties of the metric projection onto a Chebyshev
subspace are known besides those listed in Theorem 5.S. The most important of
these is that the metric projection is "linear." We now give the definition of this
fundamental concept.
5.10 Definition. If X and Yare inner product spaces, a mapping F: X ~ Y
is called linear if and only if

(5.10.1) F(ax + f3y) = aF(x) + f3F(y)


for all x, y in X and a, (3 in R
Equivalently, F is linear if and only if it is additive,

(5.10.2) F(x + y) = F(x) + F(y),


and homogeneous,

(5.10.3) F(ax) = aF(x),

for all x, y in X and a E R


78 THE METRIC PROJECTION

Linear mappings are also commonly called linear operators. A linear operator
F is said to be bounded if there exists a constant c such that

(5.10.4) IIF(x) II :<:: cllxll for all x E X.

(To be more precise, we should have written IIF(x)lly :<:: cllxllx to differentiate
between the norms involved. However, no confusion should ever result by omitting
the space indices.)
In the special case where Y = 1ft is the scalar field, a linear operator F : X -+ 1ft
is also called a linear functional.
The property of linearity is a rather strong one. Indeed, for such mappings,
continuity and boundedness are equivalent, as is continuity at a single point!
5.11 Boundedness is Equivalent to Continuity. Let F : X -+ Y be a linear
mapping. Then the following statements are equivalent:
(1) F is bounded;
(2) F is Lipschitz continuous;
(3) F is uniformly continuous;
(4) F is continuous;
(5) F is continuous at some point.
Proof. (1) =} (2). If F is bounded, there exists c > 0 such that IIF(x)11 :<:: cllxll
for all x. By linearity,

IIF(x) - F(y)11 = IIF(x - y)11 :<:: cllx - yll

for all x, y E X. That is, F is Lipschitz continuous.


The implications (2) =} (3) =} (4) =} (5) are obvious.
(5) =} (1). Suppose F is continuous at the point Xo EX. If F were not bounded,
then for each integer n ~ 1, there exists Xn E X such that IIF(x n ) II > nllxnll. Set
Zn = xnl(nllxnll). Then Ilznll = lin and

(5.11.1)

for every n. But Zn -+ 0 implies Zn + Xo -+ Xo· By continuity of F at Xo, we obtain

which contradicts (5.11.1). Thus F must be bounded. •

For a bounded linear operator F, the smallest constant c that works in (5.10.4)
is called the norm of F and is denoted by IIFII. Thus

(5.11.2) IIFII := inf{c I IIF(x)11 :<:: cllxll for all x EX}.

5.12 Norm of Operator. Let X and Y be inner product spaces with X f {O}
and let F : X -+ Y be a bounded linear operator. Then

(5.12.1) IIF(x) II :<:: 1IFllllxli for all x E X,


LINEAR METRIC PROJECTIONS 79

and

(5.12.2) IIFII = sup IIFII(XII)11 = sup IIF(x)11 = sup IIF(x)ll·


x,,"O x IIxll9 Ilxll=l

Proof. The relation (5.12.1) follows from the first equality in (5.12.2).
To verify (5.12.2), first note that

IIFII = inf {c I IIF(x)11 :s: c


Ilxll
for all x # o} = sup IIF(x)ll.
40 Ilxll
Also,

sup IIF(x) I :s: sup IIF(x)ll:S: sup IIFII (xII) I :s: sup IIFII(xll)11
IIxll=l IIxll9 0<lIx1l9 x 40 x

= sup
X,,"O
IIF (-llxll)
x
II = sup
IIxll=l
IIF(x)ll·
Hence equality must hold throughout this string of inequalities. •
Next we consider a few examples of bounded linear operators. One of the most
important, from the point of view of approximation theory, is the metric projection
onto a Chebyshev subspace.
5.13 Linearity of Metric Projection. Let M be a Chebyshev subspace in
the inner product space X (e.g., a closed subspace of a Hilbert space). Then:
(1) PM is a bounded linear operator and IIPMII = 1 (unless M = {O}, in which
case IIPMII = 0).
(2) PM is idempotent: Pir = PM.
(3) PM is "self-adjoint":

(5.13.1) (PM (x), y) = (x, PM(y)) for all x, Y in X.

(4) For every x E X,

(5.13.2)

(5) PM is "nonnegative":

(5.13.3) (PM(x),x) 2: 0 for every x.

Proof. (1) Let x, Y in X and Q, fJ in JR. By Theorem 4.9, x - PM(x) and


Y - PM (y) are in M.L. Since M.L is a subspace,

Since aPM(x) + fJPM(y) E M, Theorem 4.9 implies that aPM(x) + fJPM(Y) =


PM(ax+fJy). Thus PM is linear. Using part (5) of Theorem 5.8, we get IIPM(X) II :s:
Ilxll for all x. Thus PM is bounded and IIPMII :s: 1. Since PM(Y) = Y for every
Y E M, Ilyll = IIPM(Y) II :s: IIPMllllyl1 implies that IIPMII 2: 1 and hence IIPMII = 1.
80 THE METRIC PROJECTION

(2) This follows from (1) of Theorem 5.5.


(3) For any x,y in X, Theorem 4.9 implies that (PM(x),y - PM(y)) = 0, and
hence

(5.13.4)

By interchanging the roles of x and y in (5.13.4), we obtain

(4) Take y = x in (5.13.4).


(5) This clearly follows from (4). •
The metric projection on a Chebyshev subspace M is commonly called the or-
thogonal projection onto M for obvious reasons.

The Reduction Principle


There is a useful "reduction principle" that can be stated as a commuting prop-
erty for a pair of metric projections. It allows us, for example, to "reduce" a problem
of best approximation in a given convex subset K of X from the whole space X
to any Chebyshev subspace M that contains K. That is, we need to consider best
approximation only to points in the "reduced" space M rather than from all points
in X.
In the next theorem, and throughout the remainder of the book, we will denote
the composition of two operators by juxtaposition. For example, the operator P APB
is defined by

5.14 Reduction Principle. Let K be a convex subset of the inner product


space X and let M be any Chebyshev subspace of X that contains K. Then:
(1) PKPM = PK = PMPK,
(2) d(x, K)2 = d(x, M)2 + d(PM(x), K)2 for every x E X.

Proof. The statement PK = PM PK is obvious, since K eM. For any y E K,


we have that y E M and x - PM(x) E M1- by Theorem 4.9, so that

(5.14.1)

Since the first term on the right of (5.14.1) is independent of y, it follows that
y E K minimizes Ilx - Yl12 if and only if it minimizes IIPM(X) - Y112. This proves
that PK(x) exists if and only if PK (PM (x)) exists and PK(X) = PK(PM(X)). Hence
PK(X) = 0 if and only if PK(PM(X)) = 0. In either case, (1) is verified. From this
and (5.14.1) we obtain (2). •

The geometric interpretation of the reduction principle is the following: To find


the best approximation to x from K, we first project x onto M and then project
PM(x) onto K. The result is PK(X) (see Figure 5.14.2).
THE REDUCTION PRINCIPLE 81

Relllark. The convexity of K is not essential in the reduction principle. That


is, there is a generalization of Theorem 5.14 that is valid for nonconvex sets in
which PK is a "set-valued" mapping (see Exercise 20 at the end of the chapter).

, x

..' ...
'"

,,

,
.
.
.,
, ,
,
~ PM (x) M

:
, ,

a~;I): PK( PM (I»

Figure 5.14.2. The reduction principle

One example of the practical value of the reduction principle is this. Suppose
we are seeking best approximations to some x E X from a closed convex subset
K of a finite-dimensional subspace M of X. It is easy-at least in principle---
to compute PM(x) (and d(x,M)). This amounts to solving the normal equations
(4.10.2). By the reduction principle, the problem of finding PK(x) (and d(x, K)) is
thus reduced to that of finding PK(PM(X)) (and d(PM(x), K)). Thus, by replacing
x with PM(X), we may assume that x E M, i.e., that X = M is finite-dimensional.
We give such an application next.
5.15 Application of Reduction Principle. Let x E CdO,I] and C =
{y E P 1 I Y 2': 0 on [0, I]}. What is Pc(x)? That is, we seek the best ap-
proximation to x from the convex cone of nonnegative linear functions. By the
reduction principle, this reduces to finding the best approximation to Pp , (x) from
C. Since Pp,(x) can be obtained from the normal equations (4.10.2) by solving
two equations in two unknowns, we assume that this has already been done. That
is, we assume that x E P 1 and we seek Pc(x). Thus we can restate the problem as
follows:
Let X = C2 [0, 1] n P1 and C = {y E X I Y 2': O}. Find Pc(x) for any x E X.
First observe that if Yl(t) = t and Y2(t) = 1- t, then X = span{Yl, Y2} and C is
the conical hull of the Yi; i.e.,

To verify the latter statement, we need only observe that if Y = aYl + /3Y2, then
Y E C if and only if f3 = y(O) 2': 0 and a = y(I) 2': O.
82 THE METRIC PROJECTION

Let x E X; hence x = aY1 + bY2. We will verify that

x if min{ a, b} 2:: 0,
o if max{a + 2b, 2a + b} :::; 0,
!(a+2b)Y2 ifa=min{a,b}<Oand
(5.15.1) Pc(x) =
max{a+2b,2a+b} > 0,
!(2a + b)Y1 if b = min{ a, b} < 0 and
max{a + 2b, 2a + b} > o.
We first show that

(5.15.2) co = {aY1 + (3Y2 I max{a + 2(3,2a + (3}:::; O}.

Now, (Y1, Y1) = I; t 2dt = ~ = (Y2, Y2) and (Y1, Y2) = 101 t(l - t)dt = i· Thus
Y = aY1 + (3Y2 E Co if and only if (y, z) :::; 0 for all z E C, which is equivalent to
(Y,Yi):::; 0 fori = 1,2 or a(Yl,Yi)+(3(Y2,Yi):::; 0 fori = 1,2. That is, a/3+(3/6:::; 0
and a/6 + f3/3 :::; 0, or 2a + (3 :::; 0 and a + 2(3 :::; O. This verifies (5.15.2).
By (4) of Theorem 5.6, it follows that Pc(x) = 0 if x E Co, i.e., if max{a +
2b,2a + b} :::; O. Also, Pc(x) = x if x E C, i.e., if a 2:: 0 and b 2:: O. For the
remainder of the proof of (5.15.1) we may assume that x f: C and x f: Co. Thus
we have that min{ a, b} < 0 and max{ a + 2b, 2a + b} > O. We consider two cases.
Case 1. a = min{a, b} < O.
Since max{ a + 2b, 2a + b} > 0, it follows that b > 0 and a + 2b = max{ a +
2b,2a + b} > o. Let Yo = !(a + 2b)Y2· Then Yo E C, x - Yo = ~(2Y1 - Y2) and
(x-Yo,Yo) = O. By (5.15.2) we see that X-Yo E Co. By Theorem 4.7, Yo = Pc(x).
Case 2. b = min{a, b} < O.
A proof similar to that of Case 1 shows that !(2a + b)Y1 = Pc(x).
This completes the verification of (5.15.1). •

We conclude this chapter with a few more examples of bounded linear operators
and an example of an unbounded linear functional.
A linear operator whose domain is finite-dimensional is always bounded.
5.16 Theorem. Let F: X --7 Y be linear and suppose X is finite-dimensional.
Then F is bounded.
Proof. By Theorem 4.11, we can choose an orthonormal basis {e1' e2, ... , en}
for X. Then each x E X has the unique representation x = L~(X, ei)ei by (4.10.3).
Let c = max{IIF(ei)111 i = 1,2, ... , n}. Then by Schwarz's inequality (in 12(n)),

IIF(x)11 = iit(x,ei)F(ei)ii :::; t l(x,ei)II/F(ei)11

:::; c t I(x, ei)1 :::; cvn [t I(x, ei)12]! = cvnllxll·

Thus F is bounded and IIFII :::; cvn. •


The next example is of a bounded linear operator from C 2 [a, b] into itself that
includes several examples of interest as special cases.
THE REDUCTION PRINCIPLE 83

5.17 Bounded Linear Operator on C2[a, b]. Let k(s, t) be a continuous real
function of two variables s, t in [a, b]. For any x E C2[a, b], define F(x) at s E [a, b]
by

(5.17.1) F(x)(s) = ib k(s,t)x(t)dt.

Then F is a bounded linear operator from C2 [a, b] into itself such that

b b ] ~
(5.17.2) IIFII::; [ illk(s,tWdtds

Proof. It is clear from (5.17.1) that F is linear. It remains to show that


F(x) E C 2[a,b] and to verify (5.17.2). Fix any So E [a,b]. By Theorem 5.4, k is
uniformly continuous on [a, b] x [a, b]. Hence, for each E > 0, there exists J = J(E) > 0
such that

Ik(s, t) - k(so, t)1 < E whenever Is - sol < J and t E [a, b].
Thus for Is - sol < J,
b
IF(x)(s) - F(x)(so) I = l i [k(S, t) - k(so, t)]x(t)dtl

::; ib Ik(s, t) - k(so, t)llx(t)ldt

::; E ib Ix(t)ldt::; EVb=a [i b IX(tWdt] ~


= EVb - a Ilxll
using Schwarz's inequality. Since E > 0 was arbitrary, F(x) is continuous at so.
Since So was arbitrary, F(x) E C 2 [a, b].
Applying Schwarz's inequality to (5.17.1), we obtain, for each s E [a, b],

IF(x)(sW = lib k(s, t)x(t) dtl2 ::; [i b Ik(s, tWdt] ib Ix(tWdt


b
= [i Ik(S,t) 12 dt] Ilx11 2.
Integrating this over all s E [a, b) yields

b Ib b
IIF(x)112 = ilF(x)(s)12ds::; a llk(s, tW dtds Ilx11 2.
This shows that F is bounded and verifies (5.17.2). •
In the next chapter we will make a detailed study of the bounded linear func-
tionals on an inner product space. For the present, we only show how easy it is
to generate a large collection of bounded linear functionals on any inner product
space.
84 THE METRIC PROJECTION

5.18 Bounded Linear Functionals. Given any z E X, define the function


f = fz on X by

(5.18.1 ) f(x) := (x, z) for all x E X.

Then f is a bounded linear functional on X and

(5.18.2) Ilfll = 114

Proof. f is clearly linear, since the inner product is linear in its first argument.
By Schwarz's inequality, If(x)1 ::::: Ilxllllzll for all x E X. Hence f is bounded and
Ilfll ::::: 114 Since IIzl12 = (z, z) = fez) ::::: Ilfllllzll, it follows that Ilfll ~ Ilzll, and
hence (5.18.2) holds. •
In the next chapter we will see that if X is a Hilbert space, the converse of
Theorem 5.18 holds. That is, every bounded linear functional f on X has the
representation (5.18.1) for some z EX. For spaces that are not complete, a repre-
sentation such as (5.18.1) is not valid for every bounded linear functional. But we
will see that every bounded linear functional is the limit of such functionals.
We conclude this chapter by giving an example of an unbounded (hence discon-
tinuous) linear functional on C 2 [a, bj.
5.19 An Unbounded Linear Functional on C2 [a, bj. Fix any point to E [a, bj
and define F on C 2 [a, bj by

(5.19.1 ) F(x) = x(to),

(F is called "point evaluation at to.") Then F is easily seen to be a linear functional


on Cda, bj. Consider any sequence {x n } in C2 [a, bj having the property that 0 :::::
xn(t) ::::: ..;n,
xn(to) = ..;n,
and xn(t) = 0 whenever It-tol ~ 1/(2n). Then Ilxnll ::::: 1
for all n, but F(x n ) = ..;n.
Thus sUPllxIl91F(x)1 = 00, and so F is not bounded.
Although we shall not do so here, it is possible to prove the existence of an
unbounded linear functional on every infinite-dimensional space.

Exercises

1. If K is any set with 0 E K, show that K n KO = {O}.


2. If K = {x E X I Ilxll ::::: I}, show that KO = Kl. = {O}. More generally,
show that if K is any set with 0 E int(K), then KO = Kl. = {O}.
3. For any set K and scalar a =1= 0, (aK)O = (sgna)KO, where sgna:= a/lal
4. Fix any z E X := C 2 [a, bj and define F on X by

F(x) := z· x for every x E X.

(The mapping F is called "multiplication by z.") Show that


(a) F : X H X is linear.
(b) F is bounded and IIFII = Ilzlloo, where Ilzlloo := maxa<t<b Iz(t)l·
5. In the space C2 [0, 1], consider the function x(t) = sin t. Suppose the trun-
cated Taylor series x(t) = t - 13t3 is used instead of x(t). If K is any convex
EXERCISES 85

Chebyshev set in C 2 [0, 1], find the largest error incurred if PK(X) is used
instead of PK(X). That is, obtain an upper bound for IIPK(x) - PK(x)ll·
6. Fix an integer n:::o: 2, and let M = {x E ben) I L:~ xCi) = O}.
(a) Show that M is a Chebyshev subspace of 12(n) having dimension n - 1.
(b) Find the best approximation in M to each of the unit basis vectors
el, e2, ... , en in 12(n). (Here ei(j) = Oij.)
(c) Find the best approximation in M to any x E ben). [Hint: It suffices
to use (b) and linearity of PM']
7. Let {Xl, X2, ... , xn} be a linearly independent set in X and define F on X
by
n
F(x) = L(X,Xi)Xi.
I

(a) Show that F is a bounded linear operator from X into itself.


(b) Obtain an upper bound for IIFII.
8. Show that the mapping F on 12 defined by
F(x)(n) = x(n + 1) (n = 1,2, ... )
is a bounded linear operator from 12 into itself with IIFII = 1. Is F one-to-
one or onto? [F is called the (backward) shift operator.]
9. Let K be an approximatively compact Chebyshev set (not necessarily con-
vex). Show that PK is continuous. [Hint: If Xn -+ xo, then {PK(x n )} is a
minimizing sequence.]
10. (The cone of positive functions) Let C = {y E C 2 [a, b] i y:::o: O}. Verify
the following:
(a) C is a Chebyshev convex cone and
Pc(x) = x+:= max{x,O} for each x E C2[a,b].
(b) Co = -C.
(c) Co is a Chebyshev convex cone and
Pco(x) = -X-:= -max{-x,O} for each x E C 2[a,b].
(d) x = x+ - x- for every x E C 2 [a, b].
11. Let X = C 2 [a,b] and
C = {y E X lyE Po, y:::o: O}.
(a) What is CO?
(b) What is Pc(x) for any x E X?
12. (Isotone regression) Let X = 12(n) and
C = {x E X I x(1) ~ x(2) ~ ... ~ x(n)}.
(a) Show that C is a Chebyshev convex cone.
(b) Show that for every x, y in X,
n
(x,y) := Lx(i)y(i)
I
n n
=x(l) Ly(i) + [x(2) - x(I)] Ly(i)
I 2
n
+ [x(3) - x(2)] L y(i) + ... + [x(n) - x(n - 1)]y(n).
3
86 THE METRIC PROJECTION

(c) Verify that Co = {y E X I E~ y(i) = 0, EZ y(i) ~ 0 for k ~ 2}.


(d) When n = 2, show that Yo = Pc(x) if and only if Yo = x when x E C
and Yo(i) = a (i = 1,2), where a = ~[x(l) + x(2)], when x rt C.
[Hint: Part (c).]
(e) Characterize Pc(x) for any x E l2(n).
[Hint: Part (c).]
13. Let Ml :J M2 :J ... :J Mn be Chebyshev subspaces. Show that

14. Let M and N be orthogonal subspaces that are Chebyshev. Show that
M + N is a Chebyshev subspace and

15. Let M and N be Chebyshev subspaces. Show that N C M if and only if


PNPM = PN= PMPN.
16. A projection onto a closed subspace M is a bounded linear operator P
from X onto M such that p2 = P. Show that a bounded linear operator
P from a Hilbert space X into itself is a projection onto M if and only if
P = F + (1 - F)PM for some bounded linear operator F: X -+ M.
17. Let X = Pn with the inner product (x, y) = J~ x(t)y(t)dt. Let

What is Co ? (q was described in Application 5.15.) Can you describe C~


when n > I?
18. Let K be a convex Chebyshev set in X. Show that for each pair x, y in
X, either IIPK(X) - PK(y) II < Ilx - yll or PK(x) - PK(y) = x - y. [Hint:
Theorem 5.5(2) and the Schwarz inequality.]
19. Let £(X, Y) denote the set of all bounded linear operators from X into
Y. With the pointwise operations of addition and scalar multiplication of
operators and the norm (5.11.2), show that £(X, Y) is a normed linear
space.
20. Prove the following generalization ofthe reduction principle (Theorem 5.14).
Let 8 be a nonempty subset of the inner product space X and let M be a
Chebyshev subspace of X that contains 8. Then for every x E X,
(1) PS(PM(X)) = Ps(x) = PM(Ps(x)),
(2) d(x,8)2 = d(x, M)2 + d(PM(X), 8)2.
[Hint: Look at the proof of Theorem 5.14.]
21. If X is a complex inner product space, the definition of dual cone and linear
operator are modified as follows:

KO:= {x E X I Re(x,y) ~ 0 for all y E K},


F(O'.x + j3y) = O'.F(x) + j3F(y) for all x, y in X, 0'., j3 E C.
Verify the following statements for a complex inner product space X.
(a) Lemma 5.1 is valid.
HISTORICAL NOTES 87

(b) Theorem 5.5 is valid provided that (x - y, PK(X) - PK(y) is replaced


by its real part where it appears in (2) and (3) of Theorem 5.5.
(c) Theorems 5.6,5.8, and 5.9 are valid.
(d) Theorems 5.11 through 5.14 and Theorems 5.16 and 5.18 are all valid.
22. Definitions 5.2 and 5.10 make sense in any normed linear space. Verify that
Lemmas 5.1,5.3,5.11,5.12, and 5.16 are all valid in arbitrary normed linear
spaces.

Historical Notes
The term metric projection was apparently first used in a paper of Aronszajn
and Smith (1954).
For results concerning the metric projection onto a closed convex subset of a
Hilbert space, perhaps the best single reference is the first part of the long two-
part paper of Zarantonello (1971). (In the second part of this same paper, he even
developed a "spectral theory" that was based on projections onto closed convex
cones that extended the known theory that was based on projections onto closed
subspaces. )
The monotonicity of the metric projection (Le., Theorem 5.5(3)) was observed
by Nashed (1968). Parts (2) and (6) of Theorem 5.6 (in a Hilbert space setting)
are due to Moreau (1962) (see also Zarantonello (1971; p. 256) and Aubin (1979;
p. 18)). The projection theorem (Theorem 5.9), for closed subspaces of a separable
Hilbert space (respectively for closed convex cones in a Hilbert space), goes back
at least to von Neumann (1929a) (respectively Moreau (1962)). The "if" part of
the projection theorem, for both the subspace and cone cases, seems new.
That boundedness is equivalent to continuity for linear operators (i.e., Theorem
5.11), and that the orthogonal projection onto a closed subspace of a Hilbert space
is linear, idempotent, and self-adjoint, are both classical results (see von Neumann
(1929a)).
The reduction principle (Theorem 5.14) can be traced to Deutsch (1965) (see
also Laurent (1972; Prop. (2.2.6), p. 46)).
CHAPTER 6

BOUNDED LINEAR FUNCTIONALS


AND BEST APPROXIMATION
FROM HYPERPLANES AND HALF-SPACES

Bounded Linear Functionals


In this chapter we study the dual space of all bounded linear functionals on an
inner product space. Along with the metric projections onto Chebyshev subspaces,
these functionals are the most important linear mappings that arise in our work.
We saw in the last chapter that every element of the inner product space X natu-
rally generates a bounded linear functional on X (see Theorem 5.18). Here we give
a general representation theorem for any bounded linear functional on X (Theorem
6.8). This implies the Fn"\chet-Riesz representation theorem (Theorem 6.10), which
states that in a Hilbert space every bounded linear functional has a "representer."
This fact can be used to show that every Hilbert space is equivalent to its dual space.
In general (incomplete) inner product spaces we can characterize the bounded lin-
ear functionals that are generated by elements, or representers, in X. They are
precisely those that "attain their norm" (Theorem 6.12). We should mention that
many of the results of this chapter-particularly those up to Theorem 6.12-can be
substantially simplified or omitted entirely if the space X is assumed complete, i.e.,
if X is a Hilbert space. Because many of the important spaces that arise in practice
are not complete, however (e.g., C 2 [a, b]), we have deliberately avoided the simpli-
fying assumption of completeness. This has necessitated developing a somewhat
more involved machinery to handle the problems arising in incomplete spaces.
Hyperplanes are the level sets of bounded linear functionals. They play an im-
portant role in the geometric interpretation of various phenomena (e.g., the "strong
separation principle", Theorem 6.23). Many problems of best approximation can
be given "dual" reformulations in terms of problems concerning best approximation
from certain hyperplanes. The usefulness of such a reformulation stems from the
fact that the distance from a point to a hyperplane and the best approximation
from a hyperplane to a point have simple explicit formulas!
Finally, we study the problem of best approximation from "half-spaces" and,
more generally, from a "polyhedron" (i.e., an intersection of finitely many half-
spaces). A half-space is a set determined by all the points on one side of a hy-
perplane. As with hyperplanes, there are simple formulas for computing best ap-
proximations from half-spaces as well as the distance to half-spaces. Moreover,
half-spaces are the fundamental building blocks for closed convex sets, since ev-
ery closed convex set in an inner product space is an intersection of half-spaces
(Theorem 6.29).
In this paragraph we state explicitly some basic properties of linear functionals
that follow as particular consequences of results proved in Chapter 5 for arbitrary
linear operators.

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
90 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

Recall that a linear functional on the inner product space X is a real function
f on X such that
(6.1.1) f(ax + j3y) = af(x) + j3f(y)
for all x, y in X and a, j3 in]E.. A linear functional f is bounded if there exists a
constant c such that

(6.1.2) If(x)1 ::; cllxll for all x EX.

The smallest of these constants c is called the norm of f and has the alternative
expressions

(6.1.3) Ilfll = sup Ifll(xII)I = sup If(x)1 = sup If(x)l·


#0 x IIxll9 Ilxll=l
Finally, we recall the following facts.
6.1 Theorem. A linear functional f is bounded if and only if it is continuous
(Theorem 5.11). Moreover, f is always bounded if X is Enite-dimensional (Theorem
5.16).
6.2 Definition. The dual space of an inner product space X, denoted by X*,
is the set of all bounded (= continuous) linear functionals on X. Further, addition
and scalar multiplication in X* are deEned pointwise:
U + g)(x) := f(x) + g(x),
(af)(x) := af(x) for all x E X, a E IE..

It is easy to verify that if f and 9 are linear functionals on X, then so are f +9


and af. Moreover, we have the following theorem.
6.3 The Dual Space is a Normed Linear Space. With the norm of each
f EX* deEned by (6.1.3), X* is a normed linear space.
Proof. Let f,g be in X*. Then, for every x E X,

IU + g)(x)1 = If(x) + g(x)1 ::; If(x)1 + Ig(x)1 ::; (Ilfll + Ilgll)llxll·


Hence f +9 E X* and Ilf + gil::; Ilfll + Ilgll· Also,
sup l(af)(x)1 = sup laf(x)1 = lal sup If(x)1 = laillfll·
IIxll=l IIxll=l IIxll=l

Thus af E X* and Ilafll = tal Ilfll. Thus X* is a linear space. Clearly, Ilfll 2': 0
and Ilfll = 0 if and only if f(x) = 0 for all x E X, i.e., f = O. Thus X* is a normed
linear space. •
Actually, it is possible to define an inner product on X* (that generates the same
norm) so that X* is even a Hilbert space! This will follow easily from a result to
be established shortly (see also Exercises 7 and 8).
From now on, we will usually use the more suggestive notation x* , y*, etc. (rather
than f, g, etc.) to denote elements of the dual space X*.
In Chapter 5 we showed how the elements of X generate bounded linear func-
tionals on X. We can say substantially more.
BOUNDED LINEAR FUNCTIONALS 91

6.4 Definition. Let A denote the "hat" mapping on X defined by x 1-7 X, where

(6.4.1) x(y) := (y, x) for every y E X.

Using Theorem 5.18 we see that x E X* and Ilxll = IIxll.


6.5 X Generates Elmnents in X*. The hat mapping X -+ X* defined
A :

as in Definition 6.4 is a one-to-one bounded linear operator, having norm one, such
that Ilxll = Ilxll for every x E X.
Proof. Let F(x) = x; i.e., F(x)(y) = (y,x) for all y. Then, for any z E X,

F(ax + jiy)(z) = (z, ax + jiy) = a(z, x) + ji(z, y)


= aF(x)(z) + jiF(y)(z) = [aF(x) + jiF(y)](z).
Thus F(ax + jiy) = aF(x) + jiF(y) and F is linear.
As already noted, IIF(x)11 = 115:11 = Ilxll for all x, so F has norm one.
Finally, since F is linear and norm-preserving, we have IIF(Xl) - F(X2)11
IIF(Xl - x2)11 = Ilxl - x211· From this it follows that F is one-to-one. •
We call x the "representer" of the functional x E X*. More generally, an element
x E X is called a representer for a functional x* E X* if x* = x, i.e., if

x*(y) = (y,x) for every y E X.

By Theorem 6.5, a functional x* E X* can have at most one representer, and the
set
X := {x I x E X}
of all functionals in X* that have representers is a subspace of X*.
It is natural to ask whether every bounded linear functional on X has a repre-
senter in X. That is, must X = X*? We shall see shortly that this is the case
precisely when X is complete. However, "most" bounded linear functionals do have
representers.
The following technical lemma contains the key step in proving the general rep-
resentation theorem for bounded linear functionals, and it is the basis for some
other interesting results as well. Loosely speaking, it states that every bounded
linear functional is close to one that has a representer.
6.6 Key Lemma. Let x* E X* with Ilx*11 = 1 and x E X with I xii :::; 1. Then

(6.6.1) Ilx* - xii:::; 8[1- x*(X)j1/2.


In particular, if E > 0 and x* (x) > 1 - (E/8)2, then

(6.6.2) Ilx* - xii < E.

Proof. To verify (6.6.1), it suffices to show that for each y E X with Ilyll = 1,

(6.6.3) Ix*(y) - (y,x)1 :::; 8[1- X*(x)F/2.


92 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

Fix any y E X with Ilyll = 1, and set z = x*(y)x-x*(x)y and (J" = sgn(z,x). Then
x*(z) = 0, and for any 0 > 0,

+ Ilx*(x)yll :::; 2,
II(J"zll :::; Ilx*(y)xll
0:::; 1- [x*(xW = [1 + x*(x)][l - x*(x)] :::; 2[1 - x*(x)],

and
Ilx - (J"ozll 2': Ix*(x - (J"oz) I = Ix*(x)l·
It follows that

Ix* (y) - (y, x) I :::; Ix*(y) - x* (y) IIxl1 21+ I(x* (y)x, x) - (x* (x)y, x) I
+ I(x*(x)y,x) - (y,x)1

:::; Ix*(y)I[1-llxI1 2] + ~«(J"oz,x) + [1- x*(x)]IIYllllxll


1
:::; 1-llx11 2 + 20 UI (J"oz112 + IIxl1 2 -11(J"oZ - x11 2] + 1- x*(x)
1
:::; 1- [x*(xW + 20[402 + 1-lx*(xW] + 1- x*(x)
1
:::; 3[1- x*(x)] + 20 + 8[1- x*(x)]
= (3 + D[1 - x*(x)] + 20.
°
If x*(x) = 1, then Ix*(y) - (y, x) I :::; 20, and since 0 was arbitrary,
Ix*(y) - (y,x)1 = 0. If x*(x) < 1, take 0 = [1- X*(x)j1/2 to obtain < 0:::; v'2 and
Ix*(y) - (y,x)1 :::; 30 2 + 30:::; 3v'20 + 30 < 80 = 8[1 x*(X)j1/2.

Thus, in either case (6.6.3) holds. If x*(x) > 1- (E/8)2, then [1- X*(X)]1/2 < E/8,
and (6.6.2) follows from (6.6.1). •

A subset D of a normed linear space X is called dense in X if for each x E X


and E > 0, there exists y E D such that Ilx - yll < E. For example, the rational
numbers are dense in the set of all real numbers.
As a consequence of Lemma 6.6, we now show that the set of functionals that
have representers in X is a dense subspace of X*.
6.7 Denseness of X in X*. Let x* E X*\{O}. If {x n } is any sequence in X
with Ilxnll :::; 1 and limx*(xn) = Ilx*ll, then {xn} is a Cauchy sequence,

(6.7.1) lim Ilx* - (1Ix* Ilin) II = 0,

and

(6.7.2) x*(x) = lim(x, Ilx*llx n ), x E X.

In particular, X is a dense subspace of X*, and the convergence in (6.7.2) is


uniform over bounded subsets of X.
REPRESENTATION OF BOUNDED LINEAR FUNCTIONALS 93

Proof. By replacing x* with Ilx*II-1x*, we may assume Ilx*11 = 1. For each


E> 0, x*(x n ) > 1 - (E/8? for n large. By Lemma 6.6, Ilx* - xnll :S E for n large,
so (6.7.1) is verified. Using Theorem 6.5, we obtain that

as n, m --+ 00. That is, {x n } is a Cauchy sequence. Suppose B is a bounded subset


of X, say Ilxll :S c for all x E B. For each E > 0, there exists an integer no such that
cllx* - xnll < E for each n ~ no. Thus for all x E B and for all n ~ no, we have

Ix*(x) - (x, xn)1 = Ix*(x) - xn(x)1 :S Ilx* - xnll Ilxll :S cllx* - xnll < E.

This verifies that the convergence in (6.7.2) is uniform over bounded sets.
That X is dense in X* follows from (6.7.1). Finally, X is a subspace of X* by
Theorem 6.5. •

Representation of Bounded Linear Functionals


Next we give a general representation theorem for bounded linear functionals on
X. It will prove quite useful throughout the remainder of the book. In particular,
it will be used in the Fnkhet-Riesz representation theorem (Theorem 6.10) and the
strong separation theorem (Theorem 6.23).
6.8 Representation and Generation of Bounded Linear Functionals. If
X is an inner product space and x* E X*, then there exists a Cauchy sequence
{x n } in X such that

(6.8.1) x*(x) = lim (x, xn) for each x E X,

and

(6.8.2) Ilx*11 = lim Ilxnll·

Conversely, if {x n } is a Cauchy sequence in X, then (6.8.1) defines a bounded


linear functional x* on X whose norm is given by (6.8.2).
°
Proof. Let x* E X*. If x* = 0, set Xn = for all n. Then (6.8.1) and (6.8.2)
hold. If x* =1= 0, choose any sequence {Yn} in X with IIYnl1 = 1 and lim x*(Yn) =
Ilx*ll. Set Xn = Ilx*IIYn· Then Ilxnll = Ilx*11 for all n, and by Theorem 6.7, {xn} is
a Cauchy sequence such that (6.8.1) and (6.8.2) hold.
Conversely, let {x n } be a Cauchy sequence in X. For any x E X,

so {(x, x n )} is a Cauchy sequence in R Thus the limit

x*(x):= lim (x,x n)

exists for each x E X and defines a real function x* on X. Moreover, for each
X,Y E X and a,/3 E lR,

x*(ax + /3y) = lim (ax + /3y, x n ) = lim [a(x, xn} + /3(y, xn)]
= a lim(x, xn} + /3lim(y, xn} = ax* (x) + /3x* (y)
94 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

implies that x* is linear.


Further, lim Ilxnll exists by (1) of Theorem 1.12, since {xn} is a Cauchy sequence,
and
IX*(X)I = lim l(x,xn)1 S; (lim Ilxnll)llxll
for each x EX. Thus x* is bounded and Ilx* II S; lim Ilx n II. To obtain the reverse
inequality, let c = supllxnll. Given any I: > 0, choose an integer N such that
n
cllx n - xmll < I: for all n, m 2: N. Then if m 2: N,

Ilx*llllxmll2: x*(xm) = lim(xm,xn)


n
= lim[(xm,xn
n
- xm) + Il xml1 2]
2:liminf[-llxmllllxn -xmll
n
+ II XmI1 2]2: -1:+ Ilxml12.

Letting m ~ 00 in this inequality, we obtain

Ilx*lllimllxmll 2:limllxm l1 2 - E-
m m

Since I: was arbitrary, it follows that Ilx*11 2: lim Ilxmll. Thus (6.8.2) holds. •
As an easy consequence of Theorems 6.7 and 6.8 we can give a proof of the
celebrated Hahn~Banach extension theorem for linear functionals. It is valid in a
more general normed linear space setting, but the particular proof given here seems
to be unique to the inner product space case and, unlike the standard proofs of the
normed linear space case, does not make any appeal to the axiom of choice.
6.9 Hahn-Banach Extension Theorem. Let M be a subspace of the inner
product space X and y* E M*. That is, y* is a bounded linear functional on M.
Then there exists x* E X* such that X*IM = y* and Ilx*11 = Ily*ll.
Proof. By Theorem 6.7, we can choose {Yn} in M with IIYnl1 = 1 such that
Ily*llYn ~ y*. In particular, {Yn} is a Cauchy sequence and

y*(y) = lim(y, Ily*IIYn), Y E M.

Defining x* on X by

x*(x) = lim(x, Ily*IIYn), x EX,

it follows from Theorem 6.8 that x* E X*, Ilx*11 = lim 11(lIy*IIYn)11 = Ily*ll, and,
obviously, x* 1M = y*. •

When X is complete, every bounded linear functional on X has a representer in


X.
6.10 Frechet-Riesz Representation Theorem. Let X be a Hilbert space.
Then X* = X. That is, for each x* E X*, there is a unique element x E X such
that

(6.10.1) x*(y) = (y,x) for each yEX,

and Ilx*11 = Ilxll·


REPRESENTATION OF BOUNDED LINEAR FUNCTIONALS 95

Proof. Let x* E X*. By Theorem 6.8, there is a Cauchy sequence {x n } in X


such that
x*(y) = lim(y,x n ), y E X,
and Ilx*11 = lim Ilxnll. Since X is complete, x:= limx n exists. By (2) (a) and (b)

°
of Theorem 1.12, Ilxll = lim Ilxnll and (6.10.1) holds. The uniqueness of x follows,
since if (y, x') = (y, x) for all y E X, then (y, x' - x) = for all y, so x' - x = 0, or
x' =x. •
This result actually characterizes Hilbert space among all inner product spaces.
That is, if X is not complete, there exists an x* E X* that fails to have a representer
(see Exercise 6 at the end of the chapter).
Left unanswered thus far has been the question of precisely which functionals in
X* have representers in X. We answer this next.
6.11 Definition. A functional x* E X* is said to attain its norm if there
exists x E X with Ilxll = 1 and x*(x) = Ilx*ll.
From (6.1.3), we see that

Ilx*11 = sup Ix*(x)1 = sup x*(x).


IIxll=l Ilxll=l

Thus x* attains its norm if and only if the "supremum" can be replaced by "max-
imum" in the definition of the norm of x* .
6.12 Norm-Attaining is Equivalent to Having a Representer. Let x* E

°
X*. Then x* has a representer in X if and only if x* attains its norm.
Proof. If x* = 0, then x = is the representer for x*, and x* attains its norm
at any y E X with Ilyll = 1. Thus we may assume x* =1= 0. By replacing x* with
x* Illx*ll, we may further assume that Ilx*11 = 1.
Let x* have the representer x. Then Ilxll = 1 and

x*(x) = (x,x) = 1 = Ilx*ll.

Thus x* attains its norm at x.


Conversely, suppose x* attains its norm at x. That is, Ilxll = 1 and x*(x) = 1.
By Lemma 6.6, x* = x. That is, x is the representer for x*. •
We now consider some particular examples of functionals that exhibit the various
representations thus far discussed.
6.13 Examples.
(1) If x* E 12(I)*, then (by Theorem 6.10) there exists a unique x E b(I) such
that
x*(y) = 2)(i)x(i) for every y E 12 (1) ,
iEI

and Ilx*11 = Ilxll·


(2) As special cases of (1), we have that if x* E 12 (respectively x* E b(n)*),
then there exists a unique x E l2 (respectively x E l2 (n)) such that
00

x*(y) = I>(i)x(i) for every y E l2


96 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

(respectively x*(y) = I:~ y(i)x(i) for every y E l2(n)) and

Ilx*11 = !lxll·
(3) If x E C 2 [a, b], then (by Theorem 5.18 or Theorem 6.5)

x*(y):= lb y(t)x(t)dt for all y E C 2 [a,b]

defines a bounded linear functional on C 2 [a, b] with Ilx* II = IIxll.


(4) If x* E C2 [a,b]* has no representer in C2 [a,b] (or, equivalently, x* fails to
attain its norm), then (by Theorem 6.7) for any sequence {x n } in C2 [a, b] with
Ilxnll :::; 1 and x*(x n ) -+ Ilx*ll, we have the representation
x*(y) = lim Ilx*lll b y(t)xn(t)dt for every y E C 2 [a,b].

(5) Consider the space X = C2 [-1, 1] and the functional x* on X defined by

x*(y) = 11 y(t)dt for all y E X.

It is clear that x* is linear and, for any y EX,

Hence x* is bounded and Ilx*11 :::; l. Consider the sequence {x n } in C2 [-1, 1] defined
by
if - 1 :::; t :::; 0,
if 0:::; t :::; lin,
if lin:::; t :::; l.
Clearly, Ilxnll :::; 1 and

X*(X n ) = 11 o
xn(t) dt = 11/n
0
ntdt + 11
lin
1 dt = 1/(2n) + (1- lin) -+ l.

Thus Ilx*11 = 1 and x*(xn) -+ Ilx*ll. By Theorem 6.7, x* has the representation

x*(y) = lim l~ y(t)xn(t)dt for every y E X.

To see that x* does not have a representer, it suffices by Theorem 6.12 to show
that x· does not attain its norm. If x* attained its norm, then there would exist
x E C2 [-1, 1] such that Ilxll = 1 and x*(x) = l. Then (by Schwarz's inequality in
CzrO,l]),

Hence equality must hold throughout this string of inequalities. It follows that
x(t) = 0 for t E [-1,0) and x(t) = 1 on (0,1]. But such an x is not continuous.
This contradiction shows that x* does not attain its norm.
BEST APPROXIMATION FROM HYPERPLANES 97

Best Approximation from Hyperplanes


The level sets of bounded linear functionals are called "hyperplanes" and are
useful for providing geometric interpretations of various phenomena.
6.14 Definition. A hyperplane in X is any set of the form

H = {y E X I x*(y) = c},

where x* E X* \ {OJ and c E R. The kernel, or null space, of a functional x* E X*


is the set
kerx* := {y E X I x*(y) = OJ.

Note that a hyperplane is never empty. For since x* f 0, choose any Xo E X


with x*(xo) f 0. Then the element y = [c/x*(xo)Jxo satisfies x*(y) = c.
We next show that if x* E X* \ {OJ, then kerx* is a maximal subspace in
X; that is, X is the only subspace that properly contains ker x*. Moreover, each
hyperplane {y E X I x* (y) = c} is a translate of ker x* .
If M and N are two subspaces of X, recall that X is the direct sum of M
and N, denoted by X = M E8 N, provided that X = M + Nand M n N = {OJ.
Equivalently, X = M E8 N if and only if each x E X has a unique representation
of the form x = y + z, where y E M and zEN (see Exercise 3 at the end of the
chapter). Recall that Theorem 5.9 established that if M is a Chebyshev subspace,
then X = M E8 M.l.
6.15 Hyperplanes Are Translates of Maximal Subspaces. Let X be an
inner product space, x* E X* \ {OJ, c E JR, M = kerx*, and H = {y E X I x*(y) =
c}. Then:
(1) For any Xl EX \ M, X = M E8 span {xd.
(2) For any Xo E H, H = M + Xo.
(3) M is a closed maximal subspace in X.
(4) H is a closed convex subset of X.
In particular, if X is complete, then Ii is Chebyshev.
Proof. (1) Fix Xl EX \ M. For each X E X, set y = X - x*(x)[X*(XI)J-IXI.
Then y E M and x = y + x*(x)[X*(XI)J-1XI. Thus X = M + span {xd. If
z E M n span {xd, then z = aXI. If a f 0, then Xl = (1/a)z E M, which is a
contradiction. Thus a = 0, and so z = 0. That is, M n span{xd = {OJ, and this
verifies (1).
(2) Fix any Xo E H. If X E M + Xo, then X = Y + Xo for some y E M implies
x*(x) = x*(xo) = c. Thus x E H. Conversely, if x E H, then y = x - Xo is in M
and x = y + Xo E M + Xo.
(3) It is clear that M is a subspace, since x* is linear; and M is closed, since x'
is continuous. It remains to show that M is maximaL If Y is a subspace in X with
Y ~ M and Y f M, choose any Xl E Y \ M. By (1),

X = M + span {xd eYe X,


so Y=X.
(4) H is the translate of a closed subspace, so it is closed by Exercise 8 at the
end of Chapter 3, and convex by (5) of Proposition 2.3. The last statement follows
from Theorem 3.5. •
98 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

In particular, in an n-dimensional space X, a hyperplane is the translate of an


(n-l)-dimensional subspace of X. For example, in Euclidean 3-space (respectively
2-space) a hyperplane is a "plane" (respectively a "line").
A hyperplane has the pleasant property that there are simple formulas for its
metric projection, as well as for its distance from any point.
6.16 Distance to Hyperplanes. Let X be an inner product space, x* E
X* \ {O}, c E R, and H = {y
E X I x*(y) = c}. Then

(6.16.1) d(x,H) = 11:*lllx*(x) - cl


for each x E X.
Proof. Assume first c = O. Then H = {y E X I x*(y) = O}, and we must show
that

(6.16.2) d(x,H) = 11:*lllx*(x)l.


Fix any x EX. For any y E H,

11:*lllx*(x)1 = 11:*lllx*(x - y)1 :::: Ilx - yll


implies that

(6.16.3) 11:*lllx*(x)l:::: d(x,H).

Conversely, given any 10 with 0 < 10 < Ilx*ll, choose z E X with liz II = 1 and
x*(z) > Ilx*ll- E. Then y:= x - x*(x)[x*(Z)]-I Z is in Hand
Ix*(x)1 Ix*(x)1
Ilx-YII= Ix*(z)l:::: Ilx*II-E
implies d(x,H) :::: Ix*(x)I(llx*11 - E)-I. Since 10 was arbitrary, it follows that
d(x,H):::: Ix*(x)lllx*II- 1 . Combining this with (6.16.3) we obtain (6.16.2).
Now suppose e i= O. Choose any Xo E H and define

Ho := H - Xo = {y E X I x*(y) = O}.

By the first part ofthe proof and the invariance by translation (Theorem 2.7(1) (i)),
we obtain
d(x, H) = d(x - Xo, H - xo) = d(x - Xo, Ho)

= 11:*lllx*(x - xo)1 = 11:*lllx*(x) - el,


which completes the proof. •
We can now characterize the hyperplanes that are Chebyshev. The functionals
having representers once again play the essential role.
BEST APPROXIMATION FROM HYPERPLANES 99

6.17 Chebyshev Hyperplanes. Let X be an inner product space, x* E X* \


{O}, c E ]F., and H = {y E X I x*(y) = c}. Then the following statements are
equivalent:
(1) H is Chebyshev;
(2) H is proximinal;
(3) Some x E X \ H has a best approximation in H;
(4) x* attains its norm;
(5) x* has a representer in X.
Moreover, if H is Chebyshev, then

(6.17.1)

for every x EX, where z is the representer for x* .


Proof. The implications (1) ==} (2) ==} (3) are obvious.
(3) ==} (4). If some x E X \ H has a best approximation Yo E H, then by
Theorem 6.16,

Ilx - yoll = Ilx~lllx*(x) - cl = 11:*lllx*(x - yo)l·

Setting z = a(x - Yo), where a = Ilx - Yoll-lsgn x*(x - Yo), we see that Ilzll = 1
and x*(z) = Ilx*ll. Thus x* attains its norm (at z).
The equivalence of (4) and (5) is just Theorem 6.12.
(5) ==} (1). Let z E X be the representer of x*. Then 2 = x* and Ilzll = Ilx*ll.
For any x E X, set y = x -llzll-2((x, z) - e)z = x - Ilx*II-2[x*(X) - e]z. Then
x*(y) = e, so that y E H, and

Ilx - yll = Ilx*II-2Ix*(x) - elllzil = Ilx*II-1Ix*(x) - eJ = d(x, H)

(using Theorem 6.16). Thus y = PH(x) by Theorem 2.4. This proves that H is
Chebyshev. In the process, we have also verified (6.17.1). •
As a consequence of (6.17.1) and Theorem 4.9, observe that the representer of a
functional x* is always orthogonal to kerx* = {x E X I x*(x) = O}.
As a simple application of Theorems 6.16 and 6.17 in Euclidean 2-space l2(2), we
obtain the well-known distance formula from a point to a line and the (not as well-
known) formula for the best approximation in the line to the point. Specifically, fix
any scalars a, b, and e, where a2 + b2 of 0, and consider the line

H = {y E l2(2) I ay(l) + by(2) = e} = {y E l2(2) I (y, z) = e},

where z = (a, b). Then for any x E l2(2),

d(x, H) = (a 2 + b2)-! lax(l) + bx(2) - el

and

(see Figure 6.17.2).


100 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

x=(x( 1), x(2))

z=(a, b)

~-----~O;c-l. _____~H~=_{~Y~E~X_I~a~y( 1) + b y(2) =c }

Figure 6.17.2

Next we give an immediate consequence of Theorem 6.17 in the space C 2 [a, b].
6.18 Best Approximation from Hyperplanes in C 2 [a, b]. Let
z E C2 [a, b] \ {O}, c E Ill, and

H= {Y E C[a, b]ll 2
b
y(t)z(t)dt = c} .
Then H is a Chebyshev hyperplane in C2 [a, b] and

(6.18.1 )

for all x E C 2 [a, b].


As a specific example of Corollary 6.18, consider C 2 [0,1] and

H = {y E C2 [0, 1] 111 y(t)t 2 dt = I} .

Then z(t) = t 2, IIzl12 = fol t 4dt = i, and


PH(x) = x ~ 5[1 1 x(t)t 2 dt ~ 1] z
for every x E C2 [0, 1]. In particular, if x(t) = et3 , we get

From Theorem 6.17 we saw that a hyperplane H is Chebyshev if a single point


that is not in H has a best approximation in H. In particular, if H is not Cheby-
shev, no point that is not in H has a best approximation in H. Sets with this
latter property are called anti-proximinal. Using the characterization of Cheby-
shev hyperplanes (Theorem 6.17), it is easy to give examples of anti-proximinal
hyperplanes. Of course, owing to part (4) of Theorem 6.15, these examples must
be in incomplete inner product spaces.
BEST APPROXIMATION FROM HYPERPLANES 101

6.19 Two Exrunples of Anti-Proxirninal Hyperplanes.


(1) Recall that in Example 3.2 we showed that the closed subspace

is not proximinal in C2 [-1, 1]. In Example (5) of 6.13, we also showed that the
functional x* defined on C 2 [-1, 1] by

x*(y):= l 1y (t)dt

is in C 2 [-1, 1]*, but fails to have a representer; hence M is a hyperplane. By


Theorem 6.17, no point in C2 [-1, 1] \ M has a best approximation in M.
(2) Let

x= {x E l2 I x(n) = 0 except for at most finitely many n}

and

H = {y E X I ~Tn/2y(n) = o}.
Then H is a hyperplane in X, and no point in X \ H has a best approximation in
H.
To verify this, define x* on X by
00

x*(y) = 2:)-n/2 y (n) for every y E X.


1

(This sum is actually a finite sum, since y(n) = 0 for n sufficiently large.) Clearly,
x* is a linear functional on X. For any y E X with Ilyll = 1, Schwarz's inequality
(in l2) implies

(6.19.1) Ix*(y)l:S; [
;;)-n
00 ] 1/2
IIYII = 1.

Thus x* E X*, Ilx*11 :s; 1, and H = kerx* is a hyperplane. Next we show that
Ilx*11 = 1. For each integer N 2': 1, define YN(n) = 2- n / 2 for all n :; Nand
YN(n) = 0 if n > N. Then YN EX, IIYNII :; 1 for all N, and
N
X*(YN) = 2:)-n = 1- TN -+ 1.
1

Thus Ilx* II = 1.
If x* attained its norm, then x*(y) = 1 for some Y E X with IIYII = 1. Thus
equality holds in (6.19.1) for this y. By the condition of equality in Schwarz's
inequality, we must have y = AZ for some A > 0, where z(n) = 2- n / 2 for all n.
But z 1. X (since z(n) =1= 0 for every n) implies Y = AZ 1. X as well. This proves
that x* does not attain its norm. By Theorem 6.17, no point in X \ H has a best
approximation in H.
102 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

Strong Separation Theorem


Hyperplanes are useful in describing certain geometric phenomena.
6.20 Definition. A hyperplane H = {y E X I x*(y) = c} is said to separate
the sets K and L if
(6.20.1) supx*(K) ::; c::; inf x*(L),
where supx*(K) := sup{x*(y) lyE K} and inf x*(L) := inf{x*(y) lyE L}.
Geometrically, H separates K and L if K lies in one of the two closed "half-
spaces" determined by H, and L lies in the other. (The two closed half-spaces
determined by H are the sets H+ := {y E X I x*(y) 2: c} and H- := {y E X I
x* (y) ::; c}.) In this chapter we will be interested only in the case where one of the
sets is a single point. In Chapter 10 we will consider the more general situation.
The next lemma implies the obvious geometric fact that if H separates K and
the point x, then x is closer to H than K (see Figure 6.20.2).

H' H

d(x, H )

d( x, H')

Figure 6.20.2

6.21 Lemma. Suppose the hyperplane H = {y E X I x*(y) = c} separates the


set K and the point x; i. e., sup x* (K) ::; c ::; x* (x). Then the hyperplane
H' = {y E X I x*(y) = supx*(K)} also separates K and x and
(6.21.1) d(x, H) ::; d(x, H') ::; d(x, K).

Proof. Using Theorem 6.16, we obtain


d(x, H) = Ilx*ll-llx*(x) - cl = Ilx*ll-l[x*(X) - c]
::; Ilx*ll-l[X*(x) - supx*(K)] = d(x,H').
STRONG SEPARATION THEOREM 103

Also, for any y E K,

Ilx*ll-l[X*(X) - supx*(K)] ::; Ilx*ll-l[X*(X) - x*(y)] ::; Ilx - YII·

This implies d(x, H') ::; d(x, K). •


It is useful to know that the dual X* of an inner product space X is itself a
complete inner product space, even if X is not complete. This can be deduced from
the following lemma that shows that the parallelogram law holds in X* (see, e.g.,
Exercise 7 at the end of the chapter).
6.22 Parallelogram Law in X*. Let X be an inner product space. For every
x*, y* in X*,

(6.22.1) Ilx* + Y* 112 + Ilx* - y* 112 = 211x* 112 + 211Y* 112.


In particular, if Ilx*11 = Ily*11 = 1 and x* # y*, then 11~(x* + y*)11 < 1.
Proof. Choose xn, Yn in X with Ilxnll = 1 = IIYnl!' x*(xn) -+ Ilx*ll, and
Y*(Yn) -+ Ily*ll· By Theorem 6.7, Ilx*llxn -+ x* and IIY*IIYn -+ y*. Thus, using
Theorem 6.5 and the parallelogram law (Theorem 1.5) in X, we obtain

Ilx* + y*112 + Ilx* - Y*1I2 = lim[II(llx*llxn + Ily*IIYn)112 + 11(llx*llxn -lly*IIYn)112]


= lim[II(llx*llx n + Ily*IIYn)11 2 + 11(llx*llxn -IIY*IIYn)112]
= lim[211(llx*llxn )11 2 + 211(lly*IIYn)11 2]
= 211x*112 + 21IY*112. •

The main geometric result of this chapter is a "strong separation" principle that
implies (among other things) that (1) each point outside of a closed convex set
may be strictly separated from the set by a hyperplane; (2) every closed convex set
is the intersection of all the closed half-spaces that contain it; and (3) there is a
distance formula from a point to a convex set in terms of distances to separating
hyperplanes. It is one of the more useful theorems in the book.
6.23 Strong Separation Principle. Let K be a closed convex set in the inner
product space X and x E X \ K. Then there exists a unique x* E X* such that
Ilx* II = 1 and

(6.23.1) d(x, K) = x*(x) - supx*(K).

In particular,

(6.23.2) supx*(K) < x*(x).

Moreover, if x has a best approximation in K (e.g., if X or K is complete), then


x* has a representer (namely, d(x,K)-l[X - PK(x)]).
Proof. Since x et K, the distance d := d(x, K) is positive. To motivate the
general case, we first prove the special case when PK(X) exists. In this case, set
z = d-1(x - PK(x». By Theorem 4.1, SUPYEK(X - PK(x),y - PK(x») = 0, and so
SUPYEK(Z, Y - PK(x») = o. Thus Ilzll = 1, and

(x, z) - sup (y, z) = (x, z) - (PK(x), z) = (x - PK(x), z) = d.


yEK
104 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

It follows that (6.23.1) holds with x* = z.


In general, if x has no best approximation in K, choose a sequence {Yn} in K
such that Ilx - Ynll -+ d. Then {Yn} is a minimizing sequence, so by the proof
of Theorem 3.4(2), {Yn} is a Cauchy sequence. Setting Zn = d-1(x - Yn), we see
that {zn} is a Cauchy sequence and Ilznll -+ 1. By Theorem 6.8, the functional x*
defined on X by
x*(y)=lim(y,zn), yEX,
is in X* and Ilx* II = 1. For each y E K, we have
x*(y) = lim(y, zn) = lim[(y - Yn,zn) + (Yn - x,zn) + (x,zn)]
(6.23.3)

We next prove that

(6.23.4)
To this end, suppose (6.23.4) fails. By passing to a subsequence, we may~assume
that (Y- Yn, zn) ~ 5 > 0 for every n. Given any 0 < A < 1, the vector AY+ (1- A)Yn
is in K by convexity and

d2 <::: Ilx - [AY + (1 - A)Yn]112 = Ilx - Yn + A(Yn _ Y)112


= Ilx - Ynl1 2 + 2A(X Yn, Yn - y) + A211Yn YI12
<::: Ilx - Ynl1 2 - 2A5d + A2C,

where c:= sUPn llYn - Y112. Letting n -+ 00, we obtain


d2 <::: d2 - A[2d5 - AC].
For A > 0 sufficiently small the term in brackets is positive, and we obtain the
absurdity d 2 < d 2 . This contradiction verifies (6.23.4).
From (6.23.3) and (6.23.4), we see that x*(y) <::: -d + x*(x) for every Y E K.
Hence supx*(K) <::: -d + x*(x), and so

d <::: x*(x) - supx*(K) = inf x*(x - y) <::: inf Ilx - YII = d.


yEK yEK

Thus equality must hold throughout this string of inequalities, and (6.23.1) is ver-
ified.
To verify the uniqueness of x*, suppose both xl and xi; satisfy Ilxi II = 1 and
d = xi(x) - supxi(K) (i = 1,2). Letting x* = ~(xi + x z), we see that Ilx*11 <::: 1
and
x*(x) - supx*(K) = TIX~(X) + TIX;(X) - TISUp[X~(Y) +x;(y)]
yEK

~ Tlx~(X) + 2- 1x;(x) - Tl supx~(K) - Tl supx;(K)


=T1d+T1d=d.
On the other hand, Lemma 6.21 implies x*(x) - supx*(K) <::: dllx*11 <::: d. Hence
x*(x) - supx*(K) = d.
It follows from the last inequality that Ilx* II = 1. By Lemma 6.22, we deduce that
xi = X2· This completes the proof. •
STRONG SEPARATION THEOREM 105

Remarks. (1) The geometric interpretation of the strong separation theorem


is this. If x is any point outside a closed convex set K, there is a hyperplane H
(namely, H = {y E X I x* (y) = sup x* (K)}) that separates K and x such that the
distance from x to H is equal to the distance from x to K (see Figure 6.23.4).

d(x, K)
r-----------------
d(x, H)
x

H = {y E X I x*(y)= upx*( K»)

Figure 6.23.4

(2) Taking K = {O} in Theorem 6.23, we immediately deduce that for every
x E X, there exists x* E X* with Ilx*11 = 1 and x*(x) = Ilxll·
(3) There is a strengthening of the strong separation principle in the case where
the convex set is either a convex cone or a subspace. We state it next.
6.24 Strong Separation for Subspaces and Convex Cones. Let M (re-
spectively C) be a closed subspace (respectively closed convex cone) in the inner
product space X and x E X \ M (respectively x E X \ C). Then there exists a
functional x* E X* with Ilx*11 = 1, x*(x) > 0, and

x*(y) = 0 for all y E M (respectively x*(y) -:; 0 for all y E C).

In addition, if M (respectively C) is Chebyshev (e.g., if X is complete), then x*


has a representer in M.L (respectively in CO).
Proof. We prove the subspace case. The cone case is similar and is left as an
exercise. By the separation theorem (Theorem 6.23), there exists an x* E X* such
that Ilx*11 = 1 and supx*(M) < x*(x). If x*(y) =I 0 for some y E M, then by
replacing y with -y if necessary, we may assume x*(y) > O. But then the vector
Yn := ny is in M for every integer n, and x*(Yn) = nx*(y) -t 00, which contradicts
supx*(M) < 00. •

Using Theorems 6.23 and 6.24, Remark (1), and Lemma 6.21, we immediately
obtain the following two distance formulas.
6.25 Distance to a Convex Set. Let K be a closed convex set in the inner
product space X and x E X \ K. Then:
(1) (Analytic formulation)

(6.25.1) d(x,K) = max{x*(x) - supx*(K) I x* E X*, Ilx*11 = I},

and this maximum is attained for exactly one such functiona1.


106 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

(2) (Geometric formulation)

(6.25.2) d(x,K) = maxd(x, H),

where the maximum is taken over all hyperplanes H that separate K and x, and
this maximum is attained for exactly one such hyperplane.
Moreover, if PK(x) exists (e.g., if X is complete), then the maximum in (6.25.1)
may be further restricted to those functionals x* that have representers in X.
6.26 Distance to a Convex Cone and Subspace. Let M (respectively C)
be a closed subspace (respectively closed convex cone) in the inner product space
X and let x E X \ M (respectively x E X \ C). Then

(6.26.1)
d(x,M) = max{x*(x) I x* E X*, Ilx*11 = 1, x*(y) = 0 for all y E M}
(6.26.2)
(respectively d(x, C) = max{x*(x) I x* E X*, x*(y) ~ 0 for all y E C}).

Moreover, if PM(X) (respectively Pc(x) ) exists (e.g., if X is complete), then the


maximum in (6.26.1) (respectively in (6.26.2») may be further restricted to those
functionals x* that have representers in M 1. (respectively in CO).
In the next chapter these distance formulas will be further refined and applied.
Suppose it were somehow possible to obtain the separating hyperplane H that
maximizes the expression (6.25.2), Le., such that d(x, H) = d(x, K). How would
this help us to compute the best approximation PK(X)? The answer is that in this
case PK(x) = PH(x), where PH(x) is given explicitly by the formula (6.17.1). This
is the content of the next theorem.
6.27 Theorem. Let K be a closed convex set, x E X \ K, and let H =
{y E X I (y, z) = c} be a hyperplane that separates K and x; that is, Z E X \ {O},
c E lR, and
sup(y, z) ~ c ~ (x, z).
yEK

Then

(6.27.1)

Suppose also that PK(x) exists (e.g., if X is complete). Then the following state-
ments are equivalent:
(1) PK(x) = PH(X);
(2) PH(x) E K;
(3) d(x,H) = d(x,K);
(4) d(x, H) ~ d(x, K).
Also, if K is a convex cone, then each of these statements implies that H is a
subspace (i.e., c = 0).
Proof. The formula for PH(X) follows by Theorem 6.17.
The implications (1) =? (2) and (3) =? (4) are obvious.
(2) =? (3). If PH(x) E K, then by Lemma 6.21,

Ilx - PH(x)11 = d(x,H) ~ d(x,K) ~ Ilx - PH(x)ll·


BEST APPROXIMATION FROM HALF-SPACES 107

Hence equality must hold throughout, and (3) holds.


(4) ==? (1). If d(x, H) ;::: d(x, K), then by Theorem 6.16,

d(x, H) = Ilzll-1[(x, z) - c] ::; Ilzll-1[(x, z) - (PK(X), z)]


= Ilzll-1(x-PK(x),z)::; Ilx-PK(x)11 =d(x,K)
::; d(x, H).

Hence equality must hold throughout this string of inequalities. This implies that
Ilx - PK(x)11 = d(x, H) and (PK(X), z) = c. Thus PK(X) E Hand PK(x) = PH(x).
This proves (1), and hence all four statements are equivalent.
To prove the last statement of the theorem, suppose K is a convex cone and all
of the equivalent statements (1) through (4) hold. Now,

s := sup{ (y, z) lyE K} ::; c

by hypothesis. If s < c, then Theorem 6.16 along with Lemma 6.21 implies that
d(x, H) < d(x, H') ::; d(x, K), which contradicts (3). Thus s = c. Since 0 E K,
s ;::: O. If s > 0, choose y E K such that (y, z) > s/2. Since K is a convex cone, it
follows that Yo = 2y E K, and (Yo, z) > s, which is absurd. Thus s = o. •
This theorem suggests an "algorithm" for computing best approximations to x
from a convex Chebyshev set K and, simultaneously, the error d(x, K). It may be
briefly described as follows.
Algorithm.
(1) Obtain a hyperplane H = {y E X I (y,z) = c} that separates K and x,
and whose distance from x is maximal over all such hyperplanes; hence d(x, H) =
d(x,K).
(2) Then

d(x, K) = Ilzll-11(x, z) - cl and PK(X) = x -llzll-2[(x, z) - c]z.

Obviously, the success of this algorithm depends entirely on one's ability to


determine maximal separating hyperplanes.
For the remainder of this chapter, we will study the best approximation theory
from two important classes of sets: the half-spaces and the polyhedra.

Best Approximation from Half-Spaces


In this section we study the best approximation theory from half-spaces. We
will see that there is a close analogy with that of hyperplanes.
6.28 Definition. A half-space in an inner product space X is any set of the
form
H={YEXlx*(y)::;c} or H={YEXlx*(y);:::c},
where x* E X* \ {O} and c E R
In other words, a half-space is the set of all points that lie either on, or to one
side of, a hyperplane.
108 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

Remarks. (1) Every half-space is a nonempty closed convex set.


(2) Any half-space of the form H = {y E X I x* (y) 2: c} can obviously be written
in the form H = {y E X I xl(y) :s
cd, where xi = -x* and Cl = -c. Thus we
need only consider those half-spaces where the defining inequality is :S.
Half-spaces are fundamental for constructing all closed convex sets, as the fol-
lowing theorem attests.
6.29 Closed Convex Sets are Intersections of Half-Spaces. Let K be a
nonempty set in the inner product space X. Then K is closed and convex if and
only if K is the intersection of all the half-spaces that contain K.
Proof. The "if" part is clear, since the intersection of a collection of closed
convex sets is closed and convex. For the "only if" part, suppose K is closed and
convex. Let {Hj I j E J} denote the indexed collection of all half-spaces that
contain K. Then obviously, njEJH j =:J K. On the other hand, if x E X \ K, the
separation theorem (Theorem 6.23) implies that there exists x* E X* with Ilx* I = 1
and c := supx*(K) < x*(x). Thus the half-space H := {y E X I x*(y) :S c}
contains K but not x and hence is in {Hj I j E J}. This proves that x r:f. njEJHj ,
and so njEJHj C K. It follows that K = njEJHj. •
The theories of best approximation from half-spaces and from hyperplanes are
similar. Our first result gives a distance formula to half-spaces and should be
compared with the distance formula to hyperplanes (Theorem 6.16).
6.30 Distance to Half-Spaces. Let X be an inner product space, x* E X* \
{O}, c E JR, and 11.:= {y E X I x*(y) :S c}. Then for each x E X,

(6.30.1) d(x, H) = 11:*11 [x*(x) - cJ+.

Proof. If x E H, then d(x, H) = 0 and x*(x) :S c, so that (6.30.1) holds. If


x r:f. H, then x*(x) > c, and the hyperplane H := {y E X I x*(y) = c} separates
H from x. By Lemma 6.21, d(x, H) :S d(x, H). On the other hand, H =:J H implies
d(x, H) :S d(x, H). Thus d(x, H) = d(x, H), and using Theorem 6.16, we obtain

d(x, H) = 11:*lllx*(x) - cl = 11:*11 [x*(x) - c]+. •

6.31 Chebyschev Half-Spaces. Let X be an inner product space, x* E X* \


{O}, c E JR, H = {y E X I x*(y) = c}, and H = {y E X I x*(y) :S c}. Then the
following statements are equivalent:
(1) H is Chebyshev;
(2) H is Chebyshev;
(3) H is proximinal;
(4) Some x E X \ H has a best approximation in H;
(5) x* attains its norm;
(6) x* has a representer.
Moreover, if x* has the representer z E X (so that H = {y E X I (y, z) :S c}), then

(6.31.1)
BEST APPROXIMATION FROM POLYHEDRA 109

for every x E X.
Proof. The equivalence of (1), (5), and (6) is from Theorem 6.17. Also, the
implications (2) =? (3) =? (4) are obvious.
In the proof of Theorem 6.30 it was shown that for each x rt H, d(x, H) =
d(x,H). Hence if (1) holds, then for all x rt H, Theorems 6.17 and 6.27 imply that
1
P/i(x) = PH(X) = x - fzlP[(x,z) - c]z,

where z is the representer of x*. In case x E H, we have PH(x) = x, and (6.31.1)


holds again. This proves that (1) =? (2) and also verifies formula (6.31.1).
To complete the proof it suffices to show that (4) implies (5). Suppose some
x E X \ H has a best approximation Yo E H. Then, using the distance formula
(6.30.1), we obtain x*(Yo) <:::: c and

Ilx - yoll = 11:*11 [x*(x) - c] <:::: 11:*11 [x*(x) - x*(yo)] = 11:*I((x - Yo) <:::: Ilx Yoll·

Hence equality must hold throughout this string of inequalities, which implies
that Ilx*11 = x*((x - yo)/llx - yoll). This proves that x* attains its norm at
(x - yo)/llx - yoll and verifies (5). •

Best Approximation from Polyhedra

6.32 Definition. A polyhedron or a polyhedral set in an inner product


space is a nonempty set that is the intersection of a finite number of half-spaces.
Thus Q is a polyhedron in the inner product space X if and only if
Q = n~{y E X I xr(y) <:::: Ci} # 0,
for some x: E X* \ {O} and Ci E lR. (i = 1,2, ... , n).
We are going to characterize the polyhedra that are Chebyshev. But first we
need to record a few useful facts that are interesting in their own right.
6.33 Lemma. If a nonzero vector x is a positive linear combination of the
nonzero vectors Xl, X2, ... , x n , then x is a positive linear combination of a linearly
independent subset of {Xl, X2, ... , x n }.
Proof. Let x = 2::7 PiXi, where Pi > 0 for all i. If {Xl, X2, ... , x n } is linearly
independent, there is nothing to do. Thus we may assume that {Xl,X2, ... ,Xn }
is linearly dependent. That is, 2::70;ixi = 0 for some scalars O;i not all zero. In
fact, we may assume that O;i > 0 for some i. Set.A := min{pdO;i I O;i > O} and
Pi := Pi - .AO;i (i = 1,2, ... , n). Then .A > 0, Pi 2': 0 for all i, f.li = 0 for at least
one i, and x = 2::7 PiXi. This proves that x is a positive linear combination of at
most n - 1 of the Xi'S. If this subset of the Xi'S is linearly independent, we're done.
If not, we repeat the above argument to get x as a positive linear combination of
at most n - 2 of the x/So By continuing in this way, after at most n - 1 steps, we
must reach the conclusion of the lemma. •
We say that a convex cone C in the inner product space X is finitely generated
if it is the conical hull of a finite number of vectors. Thus C is finitely generated if

C = con {Xl, X2, ... , x n } = { ~ PiXi I Pi 2': o} ,


for some finite subset {Xl,X2, ... ,Xn } of X.
110 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

6.34 Finitely Generated Cones are Chebyshev. Every finitely generated


convex cone in an inner product space is a Chebyshev set.
Proof. Let {Xl, X2, ... ,xn } be a set of nonzero vectors in the inner product
space X, and let C = con {Xl, X2, .•. , x n }. Since C is a convex cone in the finite-
dimensional subspace span {Xl, X2, •.. , xn} of X, it suffices by Theorem 3.8 to show
that C is closed. Let Yk E C and Yk --+ y. We must show that Y E C. If Yk = 0
for infinitely many k, it follows that Y = 0 E C. Thus we may assume that
Yk # 0 for all k. By Lemma 6.33, each Yk is in the conical hull Ck of a linearly
independent subset of {Xl, X2, ... , x n }. Since there are only finitely many different
subsets C k , by passing to a subsequence {kl' k2' ... } of {1, 2, ... }, we may assume
that Co := C k , = Ck, = .... Thus each Yk j is in Co, where Co is the conical hull of
a linearly independent subset {Xi" Xi" ... , Xi=} of {Xl, X2, ... , Xn }. It follows that
m

Yk j = L AkjisXis (j = 1,2, ... )


8=1

for some Akji s :c:: O. Since Ykj --+ y, it f~llows by Theorem 3.7(4) that
m

Y = L AisXis
8=1

and Akji s --+ Ai s (8 = 1,2, ... , m). Since Akji s :c:: 0 for all sets of indices, Ai s :c:: 0
for all s. Thus Y E Co C C. •
6.35 LeInIlla. Let x*, xi, x2' ... ,x~ be linear functionals (but not necessarily
bounded) on the inner product space X. Then the following statements are equiv-
alent:
(1) x* Espan{xi,x2, ... ,X~};
(2) x*(x) = 0 whenever x;(x) = 0 for i = 1,2, ... ,n;
(3) n1 kerxi C ker x*.

Proof. The implications (1) ===} (2) o¢=} (3) are obvious.
(2) ===} (1). Assume (2) holds and let

M = {(xi(x),x~(x), ... ,x~(x)) I X EX}.

Then M is a subspace of 12(n), and we define F on jl,;J by

F((xi(x),x;(X), ... ,x~(x))) := x*(x), X EX.


[F is well-defined, since if

then xi(x - y) = 0 for all i, so that by hypothesis (2), x*(x - y) = 0, i.e., x*(x) =
x*(y).] Then F is a linear functional on M, which by Theorem 5.16 must be
bounded. By Theorem 6.9, F has an extension z* defined on all of 12(n). By
Theorem 6.10, z* has the represention
n
z*(y) = L a(i)y(i), Y E b(n),
BEST APPROXIMATION FROM POLYHEDRA 111

for some a E b(n). In particular, for every x E X,

x*(x) = F((xi(x),x;(x), ... ,x~(x))) = z*((xi(x),x;(x), ... ,x~(x)))


n
= La(i)xi(x).
I

Thus, x* = L~a(i)xi E span{xi,x2, ... ,x~}, and (1) holds. •


From this result, we can readily deduce the next "interpolation theorem" for
linearly independent linear functionals.

6.36 Interpolation Theorem. Let {xi, x 2, ... , x~} be a linearly independent


set of linear functionals on the inner product space X. Then there exists a set
{XI,X2, ... ,Xn } inX such that

(6.36.1) (i,j= 1,2, ... ,n).

In particular, for each set of n real scalars {CI, C2, ... , cn }, the element y = L~ qXi
satisfies

(6.36.2) (i=1,2, ... ,n).

Proof. We prove (6.36.1) by induction on n. For n = 1, choose any x E X


such that xr(x) =1= 0 and set Xl = [xr(x)]-lx. Then Xr(XI) = 1. Assume now
that (6.36.1) is valid for nand {xr, X2' ... , x~+1} is linearly independent on X. By
hypothesis, there exist elements YI, Y2, ... , Yn in X such that

(i,j = 1,2, ... ,n).

Next, note that there exists x E X such that xi(x) = 0 for i = 1,2, ... , nand
x~+1(x) =1= O. For if such an element failed to exist, then by Lemma 6.35, x~+1 E
span {xi, ... , x~}, which contradicts the linear independence of {xi, x z,· .. ,x~+1}.
Now set

for i = 1,2, ... , n. Clearly, xi (Xi) = (iij for i, j = 1,2, ... , n + 1. This completes
the induction. •

Now we can characterize the polyhedra that are Chebyshev.

6.37 Chebyshev Polyhedra. Let X be an inner product space, let xi E X* \


{OJ and Ci E lR for i = 1,2, ... , n, and let Q := n?{x E X I xi(x) ::::; Ci} =1= 0. Then
the following statements are equivalent:
(1) Q is Chebyshev;
(2) Q is proximinal;
(3) Each xi (i = 1,2, ... , n) attains its norm;
(4) Each xi (i = 1,2, . .. , n) has a representer.
112 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

Proof. The equivalence of (1) and (2) is just Theorem 2.4, while the equivalence
of (3) and (4) is from Theorem 6.12. Although the theorem is true as stated, to sim-
plify the proof we will hereafter assume that the set of functionals {xi, X2' ... , x~}
is linearly independent. By Theorem 6.36, there exists a linearly independent set
{Zl' Z2, ... ,zn} in X such that

(6.37.1) xi(Zj) = Oij (i,j = 1,2, ... , n).

Then the element Zo = L~ CjZj satisfies xi(zo) = Ci for all i, so Zo E Q and Q is a


nonempty polyhedron. Furthermore, observe that Q = Qo + zo, where
n
(6.37.2) Qo := n{x E X I xi(x) -:: a}.

(4) ==?(1). Suppose each xi has a representer Xi E X (i = 1,2, ... ,n). Then
n n
(6.37.3) Qo = n{x E X I (X,Xi) -:: o} = n{xf}.

Thus, using Theorems 4.5(3) and 4.6(2), we obtain

But con {Xl, X2,.'" xn} is a Chebyshev convex cone by Theorem 6.34, and therefore
its dual cone Qo is a Chebyshev convex cone by Theorem 5.6(1). Since Q is a
translate of Qo, Q must also be a Chebyshev set by Theorem 2.7(1)(v).
(1) ==? (4). Suppose that Q is a Chebyshev set. Fix any i E {1,2, ... ,n}. It
suffices to show that xi has a representer. We may assume Ilx; II = 1 for all j. Let
Yi = - L#i Zj. Then

(6.37.4)

In particular, Yi E Qo·
For the remainder of the proof, we will use the notation Bp(z) to denote the
open ball in X centered at Z with radius p. Choose any 0 < E < 1.
Claim 1. B,(Yi) nkerxi c Qo.
For let Y E B,(Yi) n ker xi. Then for all j # i, x;(y) = x;(y - Yi) + x; (Yi) <
E - 1 < O. Thus Y E Qo.
Next choose any x E B,/4(Yi) such that xi(x) > O. Then for all j # i,

(6.37.5)

Claim 2. PQo(x) E kerxi.


For if not, then xi (PQo (x)) < O. For each A E [0,1]' define z>-. := AX + (1 -
A)PQoex). Then xi(zo) < 0 < X:CZ1) implies that there exists 0 < AO < 1 such that
xi(z>-.o) = O. Then Z>-'o E Qo using (6.37.5), and
BEST APPROXIMATION FROM POLYHEDRA 113

which is impossible.
Claim 3. PQo(x) E B E/ 2 (Yi).
For Ilx - Yill < E/4 and Yi E Qo imply that Ilx - PQo(x)11 < E/4. Hence

which proves Claim 3.


By Claim 3 we can choose 0 < E' < E/2 such that BE,(PQO(x)) C B E/ 2 (Yi)' Set
x' := E'X + (1 - E')PQo(x). Then PQo(x') = PQo(x) by Exercise 18 at the end of
Chapter 2 (or Lemma 12.1 below) and x' E BE' (PQo (x')). Also, xi(x') = E'xi(x) >
o and, for j t i,
xj(x') = hj(x) + (1 - E')Xj(PQo(x)) < (1 - E')[Xj(PQo(x) - Yi) + Xj(Yi)]
< -(1 - E'? < O.
Claim 4. d(x',Qo)::; d(x',kerxi).
For let Y E kerx;:' If Y E BE(Yi), then Claim 1 implies Y E Qo, so d(x', Qo) =
Ilx' - PQo (x') II ::; Ilx' - YII· If Y ¢:. BE(Yi), then

Ily - x'il = IIY - [E'X + (1- E')PQO(X)] II


= IIY - Yi - [E'(X - Yi) + (1- E'){PQo(x) - Y;}]II
:::: Ily - Yill-IIE'(X - Yi) + (1 - E')(PQo(x) - Yi)11
:::: E - [E'llx - Yill + (1- E') IfPQo (x) - Yill]
:::: E - [E'(E/4) + (1 - E')(E/2)] > E - E/2
> Ilx - PQo(x)11 > Ilx' - PQo(x')11 = d(x', Qo),

and the claim is proved.


Next notice that ker xi is a hyperplane that separates x' and Qo (since xi( x') >
0:::: xi(Y) for all Y E Qo). Thus by Lemma 6.21, d(x',kerxi) ::; d(x',Qo). From
Claim 4, it follows that d(x',kerxi) = d(x',Qo). By Claim 2, PQo(x') = PQo(x) E
ker xi, so PQo (x') = Pkerx : (x'). Then x' E X \ ker xi has a best approximation in
ker xi. By Theorem 6.17, xi has a representer. •
We conclude this chapter by proving a characterization theorem for best approx-
imations from a polyhedron. It is convenient to isolate the key steps of the proof
as lemmas, since they have independent interest.

6.38 Lelllma. Let {C1, C 2 , ... ,Cm } be a collection of convex sets with 0 E
nrCi . Then

(6.38.1)

In other words, the operation of taking the conical hull commutes with intersection
whenever 0 is a common point to all the sets.
Proof. The inclusion con (nrCi ) C nrconci is obvious since the right side
is a convex cone that contains nrCi , and con (nrCi ) is the smallest such convex
cone.
114 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

Now let x E nrCOnci. By Theorem 4.4(4), we can write x = PiXi for some
Pi > 0 and Xi E C i for i = 1,2, ... , m. Then (1/ Pi)X = Xi E C i for each i. Letting
P = L:~ Pi and Ai = p;/ p, we see that Ai > 0, L:~ Ai = 1, and

~x = Pi (~x) = Aixi = AiXi + (1 - Ai) ·0 E Ci


P P Pi
for each i by convexity and the fact that 0 E Ci . Hence ~x E nrCi , and so

by Theorem 4.4(4). •
Let hEX \ {O}, c E JR, and define the half-space H by

(6.38.4) H:= {x E X I (x, h) s: c}.


It is easy to verify that the interior and boundary of H are given by

intH= {XEX I (X, h) <c}


and
bdH = {x EX I (x,h) = c}.
6.39 Lemma. Let X be an inner product space and H be the half-space defined
in (6.38.4). Then for every Xo E H,
if Xo E intH,
(6.39.1) can (H - xo) = con (H _ xo) = { X
H -Xo if Xo E bdH,
and
{O} if Xo E intH,
(6.39.2) (H -xot = {
con {h} if Xo E bd H.
In particular, both con (H - xo) and (H - xo)O are Chebyshev convex cones.
Proof. If Xo E int H, then 0 E int (H - xo), so con (H - xo) = X. If Xo E bd H,
then
(6.39.3) H - Xo = {x E X I (x, h) s: O} = {ht
is a closed convex cone. This verifies (6.39.1).
Using Theorem 4.5(3), we see that if Xo E int H, then

(H - xo)O = [con (H - xo)t = X o = {O}.


On the other hand, let Xo E bd H. Since con {h} is a Chebyshev convex cone by
Theorem 6.34, it follows by using (3) and (7) of Theorem 4.5 and (6.39.3) that

con{h} = (con {h}r = {h}OO = (H -xot.


This proves (6.39.2).
The last statement follows since H is a Chebyshev set by Theorem 6.31, so H -Xo
is also, and can (H - xo) is either X or H - Xo. Also, since (H - xo)O is either {O}
or can {h}, (H - xo)O is a convex Chebyshev cone by Theorem 6.34. •
BEST APPROXIMATION FROM POLYHEDRA 115

6.40 Dual Cone to a Polyhedron. Let X be an inner product space, and for
each i = 1,2, ... , m, let hi be a nonzero element of X, c; E ~, and
Hi := {x E X I (x, hi) ::; cd.
Then for each Xo E n;n Hi,
(6.40.1)

(nl Hi - xo) 0 = fl (Hi - xot = L


iEI(xo)
(Hi - xot = con {hi liE [(xo)},

where [(xo) := {i I (xo,h i ) = cd = {i I Xo E bdHd is the active index set for


Xo relative to n;n Hi.
Proof. We have

(~Hi-XOr [~(Hi-XO)r
=

= [con~(Hi -Xo)r (by Theorem 4.5(3))

= [~con (Hi - xo)] 0 (by Theorem 6.38)

L [con (Hi -
m
= xoW (by Theorem 4.6(4))

m
= L(Hi - xo)o (by Theorem 4.5(3))
1

= L (Hi - xo)O (by Theorem 6.39(2))


iEI(xo)

= L con {hi} (by Theorem 6.39(2))


iEI(xo)

= con {hi liE [(xo)}


= con {hi liE [(xo)} (by Theorem 6.34).
But the same reasoning that established the last four equalities shows that
m

L(Hi - xot = L (Hi - xot = con {hi liE [(xo)},


1 iEI(xo)

which is closed (by Theorem 6.34). •


Now we can characterize best approximations from a polyhedron.
6.41 Characterization of Best Approximations from a Polyhedron.
Let X be an inner product space, and for each i = 1,2, ... , m, let hi E X\ {O}, Ci E
~,Hi:= {x I (x,h i ) ::; cd, and Q:= n;nHi =1= 0. Iix E X, then
m

(6.41.1)
116 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

for any set of scalars Pi that satisfy the following three conditions:
(6.41.2) Pi2':O (i=I,2, ... ,m),
m
(6.41.3) (x, hi) - Ci - LPj(hj,hi ) ::; 0 (i = 1,2, ... ,m),
j=1

and
m
(6.41.4) p;[(x,hi)-Gi- LPj(hj,hi)l =0 (i= 1,2, ... ,m).
j=1

Consequently, ifx E X and Xo E Q, then Xo = PQ(x) if and only if

(6.41.5) Xo = x - L Pihi for some Pi 2': O.


iEI(xo)

Proof. Suppose the Pi satisfy the relations (6.41.2)-(6.41.4). Set Xo := x -


L:;r' Pihi.
Then (6.41.3) and (6.41.4) imply that
(6.41.6) (xo, hi) - Ci ::; 0 and pd (xo, hi) - Gil = 0 for i = 1,2, ... , m.
Now, (6.41.6) implies that Xo E nl'Hi and Pi = 0 whenever i tJ. I(xo). Thus
Xo = x - L:iEI(xo) Pihi. It follows from Theorem 6.40 that

x - Xo = L
iEI(xo)
Pihi E con {hi liE I(xo)} = (nl' Hi - xor = (Q - xor·
By Theorem 4.3, Xo = PQ(x).
The last statement is an easy consequence of the proof of the first part. •
It is worth noticing that this characterization gives rise to an explicit formula for
best approximations in the case where the elements hi that define the half-spaces
form an orthogonal set. We establish this next.
6.42 Corollary. Let the half-spaces Hi be defined as in Theorem 6.41, and let
Q= nl' Hi· If {hI, h 2 , .•. , h m } is an orthogonal set, then for each x EX,
m 1 +
(6.42.1) PQ(x) = x - ~ Ilhi l1 2 [(x, hi) - cil hi.

In particular,
m
(6.42.2) PQ(x) = (1- m)x + L PHi (x).

Proof. Using the orthogonality of the hi, it follows from Theorem 6.41 that for
each x E X,
m
(6.42.3)

for any set of scalars Pi that satisfy the following relations for each i = 1,2, ... , m:
(i) Pi 2': 0,
(ii) (x, hi) - Ci - pillhil12 ::; 0,
(iii) p;[(x, hi) - Ci - pillhil12l = O.
EXERCISES 117

From (i) and (ii), we deduce

Pi 2: Ilh1i 112 [(x, hi) - Ci


]+ for each i.

If Pi > Ilhi ll- 2[(x, hi) - e;] + for some i, then Pi > 0 and

which contradicts (iii). Thus we must have Pi = Ilhi ll-2 [(x, hi) - Ci]+ for each
i. Substituting these values for Pi back into (6.42.3), we obtain (6.42.1). Finally,
(6.42.2) follows by using formula (6.31.1). •
6.43 An Application. Let X denote the Euclidean plane £2(2) and let

Q = {y E X I y(2) - y(l) ::; 0 and y(l) + y(2) ::; O}.

Determine the best approximation PQ(x) to x from Q when (a) x = (1,2), (b)
X= (2,0), and (c) x = (-3,-1).
Note that Q may be rewritten in the form

Q = {y E X I (y, hI) ::; O} n {y E X I (y, h2) ::; O},


where hI = (-1,1) and h2 = (1,1). Then Q is a polyhedron and (h 1 ,h2) = 0, so
Corollary 6.42 implies that for each x EX,

PQ(x) = x - IIh1l l1 2[(x, hI) ]+ hI - IIh2112


1 [ +
(x, hz)] h2

= X- ~ [x(2) - X(1)thl - ~ [x(l) + x(2)]+h z.


From this formula it is easy to compute Pdx) for any x E X. In particular,
PQ((1,2») = (0,0), Pq((2,0») = (1, -1), and PQ((-3, -1») = (-2, -2).

Exercises

1. Let F be a linear functional on X. Show that

ker F:= {x E X I F(x) = O}

is closed if and only if F is continuous.


2. Give an alternative proof of the Fnkhet-Riesz representation theorem (The-
orem 6.10) by verifying the following steps. Given any x* E X*\{O}, let
M = ker x*. Then:
(i) M is a Chebyshev subspace.
(ii) There exists Xl E M.L with II xIiI = 1.
[Hint: Take any x E X \ M and set Xl = d(x, M)-I(X - PM(x».]
(iii) For any y E X, the element y - x*(Y)[X*(XI)J-IXI is in M.
(iv) X*(XI)XI is the representer of x*.
[Hint: Xl E M.L.]
118 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

3. Show that if M and N are subspaces of X, then X = M EEl N if and only if


each x E X has a unique representation in the form x = y + z, where y E M
and zEN.
4. A linear operator U : X -+ Y between inner product spaces is called a
linear isometry if IIU(x)11 = Ilxll for all x E X.
(a) Show that every linear isometry is one-to-one.
(b) Show that U is a linear isometry if and only if U "preserves" inner
products: (U(x), U(y)) = (x, y) for all x, y in X.
[Hint: Consider the polarization identity 4(x, y) = Ilx + Yl12 - Ilx - yI12.]
(c) Two inner product spaces X and Yare called equivalent if there exists
a linear isometry from X onto Y. Show that X and Yare equivalent if and
only if there exists a one-to-one mapping U from X onto Y that
(i) preserves the linear operations:

U(ax + fly) = aU(x) + flU(y) ,


(ii) preserves the inner product:

(U(x), U(y)) = (x, y);

(iii) preserves the norm:

IiU(x)11 = Ilxll·
Thus two equivalent inner product spaces are essentially the same. The
linear isometry amounts to nothing more than a relabeling of the elements.
(d) If two inner product spaces are equivalent and one is complete, so is the
other.
(e) Every Hilbert space is equivalent to its dual space.
[Hint: Define U : X -t X* by U(x) = x.]
(f) Any inner product space is equivalent to a dense subspace of its dual
space.
[Hint: See hint for part (e).] In particular, by identifying an inner product
space X with a dense subspace X of X* (via an isometry U from X onto
X), we conclude that every inner product space is a dense subspace of a
Hilbert space.
(g) Any two n-dimensional inner product spaces are equivalent.
[Hint: Let {Xl, X2, .•. , x n } and {Yl, Y2, ... , Yn} be orthonormal bases in X
and Y. Define U : X -t Y by U (1:=; aixi) = 1:=; aiyd
5. Let x* E X* \ {OJ.
(a) Show that x E X is the representer for x* if and only if x* attains its
norm at x/llxll.
(b) Show that x* attains its norm on at most one point.
(c) Show that the set of all functionals in X* that attain their norm is a
dense subspace in X*. What is the relationship between this set and X?
6. Let X be any incomplete inner product space.
(a) Show that there exists a functional x* E X* that does not have a
representer in X.
[Hint: Let {x n } be a Cauchy sequence in X that does not converge. Define
EXERCISES 119

x* as in (6.8.1).]
(b) Show that there exists a closed subspace in X that is not Chebyshev.
[Hint: Part (a) and Theorem 6.17.]
7. Show that the dual of an inner product space is a Hilbert space.
[Hint: Lemma 6.22 and Exercise 13 of Chapter 1.]
8. Let X be an inner product space. The following steps exhibit an alternative
approach to Exercise 7 for verifying that X* is a Hilbert space.
(a) Show that the sequence {x n } in X is Cauchy if and only if {Xn} is a
Cauchy sequence in X* .
(b) If x* E X*, there exist Xn E X such that Ilxnll = Ilx*11 for every nand
Ilx* - xnll-+ O.
(c) For every pair x*, y* in X*, let {x n } and {Yn} be any sequences in X
with Ilx* - xnll -+ 0 and Ily* - lin II -+ O. Define

(8.1) (x*, y*) := lim(xn, Yn).

(i) Show that this limit exists and is independent of the particular
sequences {x n } and {Yn} chosen.
(ii) Show that (8.1) defines an inner product on X* such that
(x*,x*) = Ilx*112; that is, the norm generated by this inner product is
the same as the operator norm on X*.
(iii) Show that X* is a Hilbert space.
9. Let x* E X*, Ilx* II = 1, and 0 < E < 1. If x, Y E X, Ilxll IIYII 1,
x*(x) > 1- E2/8, and x*(y) > 1- E2/8, then Ilx - yll < E.
[Hint: The parallelogram law (Theorem 1.5).]
10. Give an alternative proof of the fact that "x* E X* has a representer if (and
only if) x* attains its norm" along the following lines. Suppose x* attains
its norm at Xl, and let M = ker x*.
(i) For any X E X, show that PM(x) = x - x*(x)[llx*ll-l]xl'
(ii) Deduce Xl E M 1..
(iii) Take the inner product of PM (x) with Xl.
11. Show that an inner product space X is complete if and only if each x* E X*
attains its norm.
[Hint: Theorems 6.5, 6.10, 6.12, and Exercise 7.]
12. (a) If X is finite-dimensional, then each x* E X* attains its norm.
[Hint: The unit ball in X is compact.]
(b) Each finite-dimensional inner product space is a Hilbert space.
13. Consider the set in l2 defined by

Show that H is a Chebyshev hyperplane, and compute PH(x) and d(x, H)


for any X E b.
14. Which of the following hyperplanes in C 2 [-1, 1] are Chebyshev or antiprox-
iminal? For those that are Chebyshev, give a formula for PH(X).
(a) H = {y E C2[-1, 1]1 f~l yet) dt = 2}.
(b) H = {y E C2 [-1, 1]1 f~l yet) dt = fol yet) dt}.
120 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

(c) H = {y E C2 [-I, 1]1 J~l y(t) costdt = O}.


(d) H = {y E C2 [-I, 1]1 J~l ety(t) dt + Jo1 (1 + t)y(t) dt = O}.
15. (a) Find the distance from a point x E l2(3) to the plane

H = {y E l2(3) I ay(l) + by(2) + cy(3) = d} ,


where a, b, c, and d are any prescribed real numbers with a2 + b2 + c2 =I O.
(b) Find the best approximation in H to any x E l2(3).
16. (a) Find the distance from a point x E l2(3) to the half-space

H = {y E 12(3) I ay(l) + by(2) + cy(3) :::; d} ,

where a, b, c, and d are any prescribed real numbers with a2 + b2 + c2 =I O.


(b) Find the best approximation in H to any x E 12(3).
17. Let K be a convex Chebyshev set. Then K = njEJHj for some collection
of Chebyshev half-spaces {Hj I j E J}.
[Hint: Inspect the proof of Theorem 6.29.]
18. Let {x*,xi, ... ,x~} c X* satisfy x*(x) :2: 0 whenever xi(x) :2: 0 for all i.
Show that x* = L~ AiXi for some Ai :2: 0, that is, x* E con{xi,x2, ... ,x~}.
[Hint: Use Lemma 6.35 to deduce x* = L~ AiXi for some Ai E R Then use
Theorem 6.36 to deduce Ai :2: 0 for all i.]
19. Suppose that {x, Xl, ... ,xn } C X satisfies (x, y) :2: 0 whenever (Xi, y) :2: 0
for i = 1,2, ... , n. Show that x = L~ AiXi for some Ai :2: o.
[Hint: Exercise 18.]
20. (Farkas leInma) Let A be an m x n matrix, x E 12(n), and suppose
(x, y) :2: 0 whenever y E b(n) and Ay :2: o. Show that x is a nonnegative
linear combination of the rows of A.
[Hint: Exercise 19.]
21. Show that the functional, whose existence is guaranteed by the Hahn-
Banach extension theorem (Theorem 6.9), is unique.
22. Let Q = {y E C2[0, 1] I J~ y(t) dt :::; 0 and Jo1 ty(t) dt :::; I}. Show that if
x(t) = t 2 , then PQ(x)(t) = t 2 -~. [Hint: Theorem 6.41.]

For Exercises 23-29 below, we now assume that X is a cOlnplex inner


product space. Let C denote the set of complex numbers.

A linear functional on X is a function f : X -+ C such that

f(ax + (3y) = af(x) + (3f(y)


for all X,y E X and a,(3 E C.
The definition of f being bounded and the norm of f is just as in the real
case (see (6.1.2) and (6.1.3». Also, just as in the real case, f is bounded if
and only if it is continuous (Theorem 5.11). The dual of X is the set of all
bounded linear functionals on X and is denoted by X*.
23. The following results are valid as stated, just as in the real case:
(a) Theorem 6.3.
(b) Theorem 6.5.
EXERCISES 121

(c) Theorem 6.7.


(d) Theorem 6.8.
(e) Theorem 6.9.
(f) Theorem 6.10.
(g) Theorem 6.12.
(h) Theorem 6.15.
(i) Theorem 6.16.
(j) Theorem 6.17.
(k) Theorem 6.22.
The complex inner product space X may be regarded as a real inner
product space X IR as follows. Let X IR = X and define addition and multi-
plication by real scalars in XIR just as in X. Then, obviously, X IR is a real
linear space. Define

(x, Y)1R := Re(x, y) for all x, y E X IR ,

where Re(a) denotes the real part of the complex scalar a.


24. Verify that X IR is a real inner product space with inner product (-, ·)IR. Next
let x* E X*. That is, x* is a bounded linear functional on X. Define Rx*
on XIR by
Rx*(x) := Re[x*(x)], x E X IR .

25. For each x* E X*, Rx* is a bounded linear functional on X IR (i.e., Rx* E
X~) and IIRx*11 = Ilx*ll.
26. Show that the key lemma (Lemma 6.6) holds, provided that in (6.6.1) and
the sentence following this equation, x* (x) is replaced by Re x* (x).
27. Prove that x* E X* attains its norm if and only if Rx* attains its norm.
28. A real hyperplane in X, or a hyperplane in X IR , is any set of the form

H = {y E X I Rx*(y) = c}
(= {y E X I Rex*(y) = c}),

where x* E X* \ {O} and c E lEt


(a) Verify that Theorem 6.15 is valid for real hyperplanes, provided that
in statements (1) and (3), X is replaced by XIR, and in (1), span{xd is
replaced by "real span {xd," i.e., all real multiples of Xl.
(b) Prove that the distance formula to hyperplanes (Theorem 6.16) is also
valid for real hyperplanes. How must formula (6.16.1) be modified in this
case?
29. A real hyperplane H = {y E X I Rx*(y) = c} is said to separate a set K
and a point x if

supRx*(K) «; c «; Rx*(x),

where supRx*(K) = sup{Rex*(y) lyE K}. (Compare this with Defini-


tion 6.20.)
(a) What is the geometric interpretation of H separating K and x?
(b) Prove that Lemma 6.21 is valid for real hyperplanes.
122 BEST APPROXIMATION FROM HYPERPLANES AND HALF-SPACES

(c) Prove that the strong separation principle (Theorem 6.23) is valid, pro-
vided that x* is replaced by Rx*. In particular, relations (6.23.1) and
(6.23.2) are replaced by

d(x, K) = Rex*(x) - supRex*(K)

and
sup Rex*(K) < Rex*(x),
respectively.
(d) The distance formulas in Theorem 6.25 are valid for real hyperplanes.
In particular, (6.25.1) is replaced by

d(x, K) = max{Rex*(x) - supRex*(K) I x* E X*, Ilx*11 = I},

and (6.25.2) is replaced by

d(x, K) = maxd(x, H),

where the maximum is taken over all real hyperplanes H that separate K
and x.
(e) Prove that the analogue of Theorem 6.27 is valid for real hyperplanes.
In particular, (6.27.1) is replaced by

PH(x) = x -llzll-2[Re(x, z) - c]z.

(f) The algorithm described following Theorem 6.27 is valid for real hyper-
planes.
30. There are several results in this chapter that are valid in any (real) normed
linear space X.
(a) Show that Theorems 6.3,6.9,6.16, and 6.21 are valid in X.
(b) Show that Theorem 6.15 holds in X, provided that the last sentence in
part (4) is deleted.
(c) Statements (2), (3), and (4) of Theorem 6.17 are equivalent in X. Also,
show that if H = {y E X I x*(y) = c} is proximinal, then

PH(x) = x - {[ x*~;*11 c] z I z E X, Ilzll = 1, x*(z) = IIX*II}.

In particular, H is Chebyshev if and only if x* attains its norm at a unique


point z E X with Ilzll = l.
(d) Theorems 6.23 and 6.25, with the last statements concerning the repre-
senters of x* deleted, are valid in X.
3l. Exercises 1 and 18 are valid in any normed linear space.

Historical Notes
The key lemma (Lemma 6.6), which was fundamental to the main results of this
chapter, was proved by Ocneanu (1982) in response to a query of mine. While the
denseness of X in X* is well-known, the quantitative version given here (Theorem
6.7) seems new. Similarly, I have not found Theorem 6.8 in the literature either.
HISTORICAL NOTES 123

The Hahn-Banach extension theorem (Theorem 6.9) is actually a special case of


one of the major cornerstones of functional analysis, which goes by the same name,
and is valid in any (real) normed linear space. It was first proved by Hahn (1927;
p. 217). Banach (1929a; p. 212), (1929b; p. 226) proved not only the extension
theorem in normed linear spaces, but also the linear space (or "geometric") version,
which turned out to be the main tool for proving important separation theorems in
locally convex spaces as well. That these theorems also have analogues in complex
spaces as well as in real spaces is due to a clever trick of Bohnenblust and Sobczyk
(1938) and, independently, Soukhomlinoff (1938). It is perhaps of some academic
interest to note that, in contrast to the known proofs of the general Hahn-Banach
theorem, the proof of Theorem 6.9 given here does not use the axiom of choice.
The Frechet-Riesz Representation theorem (Theorem 6.10) in the space L 2 [0, 1]
was established independently by Frechet (1907a), (1907; p. 439) and Riesz (1907).
The result in full generality was given by Riesz (1934). Unlike the proof given
here, Riesz's original proof was modeled after Exercise 2 at the end of the chapter).
For a linear functional, the condition that norm-attaining is equivalent to having a
representer was noted in Deutsch (1982; p. 229). Formula (6.16.1) for the distance
to hyperplanes (which is valid in any normed linear space) is due to Ascoli (1932).
The strong separation principle (Theorem 6.23) is a quantitative version of a
separation theorem. Perhaps the most useful (qualitative) separation theorem,
valid in any locally convex linear topological space, is due to Eidelheit (1936) (see
also Dunford-Schwartz (1958; V. 2.8, p. 417). The distance to a convex set (6.25.1)
is a particular case of a duality result that is valid in any normed linear space (see
Nirenberg (1961; p. 39) or Deutsch and Maserick (1967». Theorem 6.27 can be
found in Deutsch, McCabe, and Phillips (1975). The distance formula to half-spaces
(6.30.1) is an easy consequence of the distance formula to hyperplanes (6.16.1).
That finitely generated cones are Chebyshev was observed by Deutsch (1982;
p. 230). Lemma 6.35, in any linear space, was established (by induction) by
Dieudonne (1942; p. 109). The characterization of Chebyshev polyhedra (Theo-
rem 6.37) was established by Deutsch (1982) in the case where all the functionals
are linearly independent by the proof given here, and by Amir, Deutsch, and Li
(2000) in the general case by a different proof. I have not seen Theorem 6.41
(the characterization of best approximations from a polyhedron) in the literature,
although it may well be in the "folk-theorem" category.
CHAPTER 7

ERROR OF APPROXIMATION

Distance to Convex Sets


In this chapter we will be interested in determining the error d(x, K) made in
approximating the element x by the elements of a convex set K. We have already
given an explicit formula for the distance d(x, K) in the last chapter (Theorem
6.25), and a strengthening of this distance formula in the particular case where the
convex set K is either a convex cone or a subspace (Theorem 6.26). Now we will
extract still further refinements, improvements, and applications of these formulas.
For example, there are useful formulas for d(x, M) in the cases where M is a sub-
space of finite dimension or finite codimension. The latter is used to solve Problem
4 of Chapter 1. The classical Weierstrass approximation theorem is established,
which implies that the subspace of all polynomials is dense in C2 [a, b]. This result
(along with the formula for the distance to a finite-dimensional subspace and a few
other technical facts) implies a beautiful generalization due to Muntz (Theorem
7.26).
Many of the results of this chapter rest on the distance formula (6.25.1). We
restate a slightly strengthened version here for ease of reference.
7.1 Distance to a Convex Set. Let K be a closed convex subset of X and
x E X\K. Then
(7.1.1) d(x, K) = max{x*(x) - supx*(K) I x* E X*, Ilx*11 = I},
and this maximum is attained for a unique functional Xo E X* with IlxC;11 = 1.
Moreover, if the best approximation PK(x) exists (e.g., if X is complete), then

(7.1.2) d(x,K) = max {(x,z) - sup(y,z) I z


yEK
EX, Ilzll = I},
and this maximum is attained for a unique point Zo E X with Ilzo II = 1. In addition,

(7.1.3) PK(x) =x - d(x,K)zo =x - [(X,zo) - sup(y,zo)] zoo


yEK
Proof. Using Theorem 6.25, it remains only to verify (7.1.3). Let
H = {y E X I (y, zo) = c},
where c = SUPYEK(Y' zo). Then H is a hyperplane such that
sup (y, zo) = c = (x, zo) - d(x, K) s: (x, zo).
yEK
Hence H separates K and x. By Theorem 6.16, d(x, H) = (x, zo) - c = d(x, K).
By Theorem 6.27,
PK(x) = PH(X) = x - [(x, zo) - c]zo = x - d(x, K)zo. •

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
126 ERROR OF APPROXIMATION

7.2 Remarks. (1) The geometric interpretation of (7.1.1) is that the distance
from x to K is the maximum of the distances from x to hyperplanes that separate
x and K. The geometric interpretation of (7.1.2) is that in the case where PK(X)
exists, we may further restrict the search for separating hyperplanes to those that
are also Chebyshev or, equivalently, to those whose functionals have representers.
(2) Formula (7.1.3) tells us that if we can find the element Zo E X, Ilzoll = 1, that
maximizes the right side of (7.1.2), then we have an explicit formula for computing
PK(X).
(3) Formulas (7.1.1) and (7.1.2) provide us with an easy way of obtaining lower
bounds for the error of approximation d(x, K). For if x* E X* and Ilx* II = 1, then

(7.2.1) x*(x) - supx*(K) S; d(x, K).

Upper bounds can, of course, always be obtained from the obvious inequality

d(x, K) S; Ilx - yll, y E K.

(4) Theorem 7.1 may be regarded as a "max-inf" theorem. For (7.1.1) may be
rewritten as

(7.2.2) d(x,K) = max { inf x*(x - y) I x* E X*,


yEK
Ilx*11 = 1}.
Moreover, the "inf" and the "max" in (7.2.2) may be interchanged:

(7.2.3) d(x, K) = inf max{x*(x - y) I x* E X*, Ilx*11 = 1}.


yEK

To see this, we note that for each y E K there exists x* E X* with Ilx* II = 1 and
x*(x - y) = IIx - yll (namely, x* = Ilx - yll-l(::C=Y)).
Thus
Ilx - yll = max{x*(x - y) I x* E X*, Ilx*11 = I},
and hence (7.2.3) follows.
(5) Formula (7.1.1) (respectively (7.1.2)) is also valid when x E K, provided that
the maximum is taken over all x* E X* with Ilx* I S; 1 (respectively z E X with
Ilzll S; 1) rather than just IIx*11 = 1 (respectively Ilzll = 1). (See Exercise 1 at the
end of the chapter.)
(6) The supremum in (7.1.1) (respectively (7.1.2)) cannot be replaced in general
by the maximum. (See Exercise 2 at the end of the chapter.)
(7) Clearly, the search for a maximum in (7.1.1) may be further restricted to
those norm-one functionals x* with x*(x) :::: supx*(K).
In the same way that Theorem 7.1 is an improvement to Theorem 6.25, we may
state the following improvement of Theorem 6.26.
DISTANCE TO CONVEX SETS 127

7.3 Distance to a Convex Cone and a Subspace. Let M (respectively C)


be a closed subspace (respectively closed convex cone) in X and let x E X \ M
(respectively x E X \ C). Then

(7.3.1) d(x,M) = max{x*(x) I x* E X*, Ilx*11 = 1, x*(y) = 0 for all y E M}

(7.3.2)
(respectively d(x, C) = max{ x* (x) I x* E X*, Ilx* I = 1, x*(y) :s; 0 for all y E C}),

and this maxmum is attained for a unique Xo E X*.


Moreover, if PM(X) (respectively Pc(x)) exists (e.g., if X is complete), then

(7.3.3) d(x,M) = max{(x,z) I z E Ml., Ilzll = I})

(7.3.4) (respectively d(x, C) = max{ (x, z) I z E Co, Ilzll = I},

and this maximum is attained for a unique Zo E Ml. (respectively Zo E CO) with
Ilzoll = 1. In addition,

(7.3.5)

(7.3.6) (respectively Pc(x) = x - (x, zo)zo).

Clearly, to apply Theorem 7.3 for determining the distance to convex Chebyshev
cones as well as determining best approximations from such sets requires knowledge
of the dual cones of these sets. In Chapter 10, we will make a systematic study
of dual cones and their applications in duality results such as Theorem 7.3. For
the moment, we will content ourselves with the following simple but illustrative
application of Theorem 7.3.

7.4 Example. Let C be the convex cone

C = {x E l2(2) IO:S; x(2):S; x(I)}.

It is easy to verify that its dual cone is given by

Co = {y E b(2) I y(I):S; 0 and y(l) +y(2):S; O}

(see Figure 7.4.1 below).


128 ERROR OF APPROXIMATION

Figure 7.4.1

Let x E 12(2) \ C. If x E Co, then by Theorem 7.3

d(x, C) = max{ (x, z) I z E Co, Ilzll = I} = Ilxll,


and the maximum is attained for Zo = x/llxll. Moreover,

Fc(x) = x - (x, zo)zo = O.

(These facts are also a consequence of Theorem 5.5(4).)


Thus we may assume that x rt C u Co. We have two cases to consider.
Case 1: xCI) > 0 and x(2) < O.
Then for each z E Co with Ilzll = 1, we get

(x, z) = x(I)z(I) + x(2)z(2) ~ x(2)z(2) ~ -x(2),

and equality holds if and only if z = Zo = (0, -1). Thus by Theorem 7.3,

d(x, C) = max{ (x, z) I z E Co, Ilzll = I} = -x(2)


and
Fc(x) = x + x(2)zo = (x(I), 0).
Case 2: xCI) + x(2) > 0 and x(2) > xCI).
Then for each z E Co with Ilzll = 1, we get

(x, z) = x(I)z(I) + x(2)z(2) = [xCI) - x(2)]z(I) + x(2)[z(I) + z(2)]


~ [xCI) - x(2)]z(1),
DISTANCE TO FINITE-DIMENSIONAL SUBSPACES 129

and equality holds if and only if z(l) + z(2) = 0 and z(l) < 0; i.e., z = Zo =
~(-l,l). Thus by Theorem 7.3,

1
d(x, C) = max{ (x, z) I z E Co, Ilzll = I} = J2[x(2) - x(l)]

and

1
Pc(x)=x- J2[x(2)-x(1)]zo= (12:[x(1)+x(2)], 2:[x(1)+x(2)]
1 ).
Distance to Finite-Dhnensional Subspaces
We turn now to the problem of determining the distance to a finite-dimensional
or finite-codimensional subspace. Recall that if {Yl, Y2,' .. ,Ym} is a subset of X, the
GraIll matrix of this set is the m x m matrix G(Yl' Y2,.'" Ym) whose (i,j)-entry
is (Yi, Yj)· The determinant of G(Yl' Y2,.·., Ym) is called the GraIll determinant
of {Yl, Y2,···, Ym} and is denoted by g(Yl, Y2,···, Ym). Thus

7.5 Lemma. Let {Xl, X2," . ,xn } eX. Then g(Xl' X2,' .. ,xn ) i= 0 if and only
if {Xl, X2,' .. ,xn } is linearly independent.
Proof. If {Xl, X2, ... ,Xn } C X is linearly dependent, say Xl = L:~ O;iXi, then by
adding -O;i times the ith row of g(Xl' X2,.' ., Xn) to the first row for i = 2,3, ... , n,
we obtain all zeros in the first row and hence g(Xl' X2,·'· ,x n ) = O.
Conversely, if g(Xl' X2, ... , Xn) = 0, then matrix algebra implies that some row of
G(Xl' X2,"" Xn) is a linear combination of the others. Without loss of generality,
we may assume that it is the first row:
n
((Xl, Xl), ... , (Xl, Xn)) = L O;i( (Xi, Xl), ... , (Xi, Xn))
i=2

Thus

implies that

In particular, Xl - L:~ O;iXi ~ M := span {Xl, X2, ... , xn} and also Xl - L:~ ();iXi E
M. That is,
n
Xl- L();iXi E MnM.L = {O}
2

or Xl = L:~ O;iXi· Hence {Xl, X2, ... , xn} is linearly dependent. •


130 ERROR OF APPROXIMATION

7.6 Distance to a Finite-Dimensional Subspace. Let


M = span{XI,X2,'" ,xn } be n-dimensional in X. Then for each x E X,

(7.6.1) d(x M)2= g(XI,X2, ... ,Xn,x)


, g(XI,X2, ... ,xn )
and

(7.6.2) d(x, M? = [[X[[2 - (3T G- I (3,

where G = G(XI,X2,"" x n ), G- I is the inverse matrix ofG, and

denotes the transpose of the column vector (3.


Moreover,

(7.6.3)

Proof. Fix x E X and let d = d(x, M). By Corollary 4.10, PM(x) = I:~ CtiXi,
where the Cti satisfy
n
(7.6.4) L Cti(Xi, Xj) - (x, Xj) = 0 (j = 1,2, ... , n).
I

Also, x - PM(x) ~ M by Theorem 4.9, so


n
d2 = (x - PM(X),X - PM (X)) = (x - PM(x),x) = (x, x) - L Cti(Xi, x),
I

or
n
(7.6.5) LCti(Xi,X) + [d2 - (x,x)] = O.

By introducing the constant Ctn +! := 1 as the coefficient of -(x, Xj) (j =


1,2, ... , n) and d 2 - (x, x), equations (7.6.4)-(7.6.5) represent a homogeneous linear
system of n + 1 equations in the n + 1 variables Ctl, Ct2, ... ,Ctn+l. Since Ctn+l = 1,
this system has a nontrivial solution, so that the determinant of the coefficient
matrix must be zero. Thus
(Xn, Xl) -(x, Xl)

(Xn' X2) -(x, X2)

=0.

(Xn,X n ) -(x, Xn)

(Xn, x) d2 - (X,X)
DISTANCE TO FINITE-DIMENSIONAL SUBSPACES 131

Since the determinant is a linear function of its last column, we obtain


(Xl,Xl) (Xn,Xl) 0 (Xl,Xl) (Xn,Xl) (X,Xl)

(Xl, Xn) (Xn, Xn) 0 (Xl, Xn) (Xn,x n ) (X, Xn)

(Xl, X) (Xn, X) d2 (Xl, X) (Xn, X) (x,x)


That is,
d 2g(Xl' X2,·.·, xn) = g(Xl, X2,··· , Xn, x),
and (7.6.1) follows.
Next observe that (7.6.4) may be rewritten in matrix notation as CtTG = (3T or
Ct T = (3T G- l . Using this and (7.6.5), we obtain
n
d2 = IIxI1 2- L Cti(Xi, x) = IIxI1 2- Ct T (3 = IIxI1 2- (3T G- l (3,
1
which establishes (7.6.2).
To verify (7.6.3), first suppose X E Ml.. Then (3T = (0,0, ... ,0), and the result
follows from (7.6.2). Thus we may assume that X if. Ml.. By (7.3.3), applied to
Ml. instead of M, and using the facts that Ml. is Chebyshev and Ml.l. = M (see,
e.g., Theorem 5.8), we obtain
d(x, Ml.) = max { (x, z) I z E M, Ilzll = I}

= max { (X, ~CtiXi) I Cti E~, II~CtiXill = 1 }


= max { ~Cti(X'Xi) I Cti E~, II~CtiXill = 1 }.
If II L~ CtiXi11 = 1, then by Schwarz's inequality (in ben)), we get

~Cti(X'Xi) ~ (~Ct;) ~ (~I(x,Xi)12) ~,


with equality holding if and only if there is a A > 0 such that Cti = A(X, Xi) for all
i. If equality holds, then since II L~ CtiXi I = 1, we obtain

or A = I L~(x,xi)xill-l. That is, Cti = I L~(x,xj)xjll-l(x,xi) for each i. It


follows that

d(x, Ml.) = 11~(X'Xi)Xirl ~ l(x,xi)1 2.


Using Theorem 5.8 (3), (7.6.3) is obtained. •
Our first application of the distance formula (Theorem 7.6) yields sharp bounds
on the Gram determinant g(Xl, X2, . .. , xn).
132 ERROR OF APPROXIMATION

7.7 Gram Determinant Inequality. Let {X1,X2, ... ,xn } be a set of nonzero
vectors in X. Then
(7.7.1 )
Moreover, equality holds on the left (respectively right) side of (7.7.1) if and only if
{XI,X2, ... ,xn } is linearly dependent (respectively orthogonal).
Proof. By Lemma 7.5, g(XI' X2, ... ,xn ) =;; 0 if and only if {Xl, X2, ... ,Xn } is
linearly dependent. We next show that if {Xl, X2,· .. , Xn} is linearly independent,
then g(XI' X2, . .. , xn) > O. The proof is by induction on n. For n = 1,
g(XI) = (Xl, Xl) = IIxll12 > O.
Assume that the result is true for some n ~ 1. By Theorem 7.6 with X = xn+l, we
see that

since Xn+l rt M = span {Xl, X2, . .. ,xn }.


It remains to verify the right side of (7.7.1). Clearly, if {Xl, X2, ... , xn} is or-
thogonal, then (Xi, Xj) = 0 for i =J j, so that
(Xl, Xl) 0 o
o (X2' X2) o
o o
and equality holds on the right side of (7.7.1).
It is left to prove that
(7.7.2)
and if equality holds, then {Xl, X2, ... ,Xn } must be orthogonal. We again proceed
by induction on n. For n = 1,
g(XI) = (XI,Xl) = Ilx1112,
and equality holds. But {xd is trivially an orthogonal set. Assume that (7.7.2) is
valid for some n ~ 1. Applying Theorem 7.6 with X = Xn+l, we obtain
g(XI' X2, ... , Xn , Xn+l) = g(XI' X2, ... ,xn)d(Xn+I' M)2
:':: Ilxd 21Ix2112 ... Ilxn 1121Ixn+l11 2 .
Thus the inequality in (7.7.2) holds. Moreover, if equality holds, then we must have
both
(7.7.3)
and
(7.7.4)
By the induction hypothesis, (7.7.3) implies that {XI,X2, ... ,X n } is orthogonal.
Further, (7.7.4) implies that PM(Xn+l) = O. Hence by Theorem 4.9, Xn+l E MJ.,
and thus {Xl, X2, ... , Xn , Xn+I} is orthogonal. •
If { Xl, X2, ... , Xn } is an orthonormal set, the formulas in Theorem 7.6 substan-
tially simplify.
FINITE-CODIMENSIONAL SUBSPACES 133

7.8 Corollary. Let {e1, e2, ... , en} be an orthonormal basis for M. Then for
eachxEX,
n
(7.8.1) PM(X) = ~)x,ei)ei

and
n
(7.8.2) d(x, M)2 = IIxl1 2- 2.:= 1(x, ei) 12.

Finite-Codimensional Subspaces
Recall that if M and N are subspaces of the inner product space X, we write
X = M EB N if each x E X has a unique representation in the form x = Y + z,
where Y E M and zEN. Equivalently, X = M EB N if and only if X = M + Nand
MnN={O}.
7.9 Definition. A closed subspace M of X is said to have codimension n,
written codim M = n, if there exists an n-dimensional subspace N such that X =
MEBN.
For example, by Theorem 6.15 the kernel of any nonzero bounded linear func-
tional is a closed subspace of codimension one. The following lemma shows that
this notion is well-defined.
7.10 Lemma. Let M be a closed subspace of X and let Nand L be subspaces
with dim N < (Xl and
X = M EB N = M EB L.
Then dimN = dimL.
Proof. If the result were false, we could assume that dim L > n := dim N.
Then L contains a linearly independent set {Y1, Y2,· .. , Ym}, where m > n. Write
Yi = Ui + Vi, where Ui E M and Vi EN. Since N is n-dimensional and m > n,
{V1,V2, ... ,Vm } must be linearly dependent. Thus there exist scalars (Xi, not all
zero, such that :L;" (XiVi = o. Then
m m m m

o oF 2.:= (XiYi = 2.:= (XiUi + 2.:= (Xi Vi = 2.:= (XiUi·


1 1 1 1

Since :L;" (XiYi ELand :L;" (XiUi E M, it follows that :L;" (XiYi E M n L = {O},
which is absurd. •
The following alternative description of the subspaces that have codimension n
will be useful to us.
7.11 Characterization of Finite-Codimensional Subspaces. Let M be a
closed subspace of X and n a positive integer. Then codimM = n if and only if

M = n{
n

i=l
x E XI x;(x) = O}
134 ERROR OF APPROXIMATION

for some linearly independent set {xi, ... , x~} in X* .


Proof. Let codim M = n. Then there exists an n-dimensional subspace N =
span {Xl, x2, .. . , Xn } such that X = M EB N. We proceed by induction on n. For
n = 1, Theorem 7.3 allows us to choose an xi E X* with Ilxi II = 1, xi 1M = 0, and
d(Xl,M) = Xl(Xl). Since M is closed and Xl rf:. M, d(Xl,M) > O. We deduce that
M = kerxi.
Now assume that the result is valid for n = m and let X = M EB N, where N =
span {Xl, ... ,Xm+l } is (m + I)-dimensional. Clearly, we have X = L EB N m, where
L = span {M, Xm+l } and N m = span {Xl, ... ,xm}. By the induction hypothesis,
L = nl"ker x; for some linearly independent set {xi, ... , x;'J in X*. But L =
M EB span{xm +1}, so by the case where n = 1, we have that M = kery* for some
y* E L* \ {O}. By the Hahn-Banach extension theorem (Theorem 6.9), there is an
x* E X* such that x* IL = y* and Ilx* II = Ily* II· Now x* rf:. span {xi, . .. ,x;;"}' since
x; IL = 0 for i = 1, ... , m and x* IL = y* of. o. Thus, setting X;;"+l = x*, we see that
{xi, ... ,X;;"+l} is linearly independent and

M = kery* = (kerx*) n L = nl"+l ker xi.


This completes the induction.
Conversely, suppose that for some linearly independent set {xi, ... , x~} in X*,
we have M = nJ'kerxi. We must show that codimM = n. The proof is by
induction. For n = 1, M = ker xi for some xi E X* \ {O}. Choose any Xl EX \ M
and let N = span{xI}. By Theorem 6.15, X = MEBN and hence M has codimension
one.
Next assume that the result is valid for n = m and let M = n;n+1 ker xi for some
linearly independent set {xi, ... ,x;;"+1} in X*. Letting Y = ker x;;"+l , we see that

Next we observe that {xi Iy, ... , x;;" Iy} is linearly independent in Y*. For if
L:;'" QiXi(y) = 0 for all y E Y, then by Lemma 6.35, L:;'" QiXi = QX;;"+1 for some
scalar Q. By linear independence of {xi, ... ,x;;"+1}' it follows that Qi = Q = 0 for
all i. By the induction hypothesis (applied to M in Y), there is an m-dimensional
subspace N m of Y, hence of X, such that Y = M EB N m . Moreover, by the case
where n = 1, we have that X = YEBNI , where dimNI = 1. Setting N = N m +NI'
it follows that N is an (m + I)-dimensional subspace. But

X = Y + NI = (M + N m ) + NI = M + N.
To complete the proof, it suffices to show that M n N = {O}. But if z E M n N,
then z = Ym + Yl, where Ym E N m and Yl E N l . Since

z ~ Ym = Yl E NI n (M + N m ) = Nl n Y = {O}
implies z = Ym E N m n M = {O}, the result follows. •
From this result we deduce, in particular, that the subspaces of co dimension one
are precisely the hyperplanes through the origin.
The subspaces of finite codimension that are Chebyshev have a simple descrip-
tion. They are precisely those for which the defining functionals have representers.
This fact, proved next, also generalizes Theorem 6.17.
FINITE-CODIMENSIONAL SUBSPACES 135

7.12 Characterization of Finite-Codimensional Chebyshev Subspaces.


Let M be a closed subspace of the inner product space X with codim M = n. Thus
n
(7.12.1) M = n{x E X I xT(x) = O}

for some linearly independent set {xi, X2, ... , x~} in X*. Then M is Chebyshev if
and only if each xi has a representer in X.
In other words, a closed subspace M of codimension n is Chebyshev if and only
if
n
M = n{x E X I (X,Xi) = O}
I

for some linearly independent set {Xl,X2, ... ,xn } in X.


Proof. Suppose M is Chebyshev. Then X = M EB M1. by Theorem 5.9.
Since codimM = n, it follows by Lemma 7.10 that dimM1. = n. Hence M1. =
span{xl, ... ,xn } for some linearly independent set {Xl, ... , X n }. Since M is Cheby-
shev, Theorem 5.8 (6) implies that M = M1.1.. That is,
n n

where Xi is the linear functional on X having Xi as its representer. By (7.12.1) and


(7.12.2), we have that
n n
(7.12.3) nkerxi = nkerXi.
I 1

In particular, for each j = 1,2, ... ,n, we have that ker xj =:J nfker Xi. By Lemma
6.35, xj E span{ Xl, X2, .. . ,Xn }, so xj = L~ QiXi for some scalars Qi. It follows
that L~ QiXi is the representer for xj. Hence each xj has a representer in X.
Conversely, suppose each xj has a representer Xj in X. Then (7.12.1) implies
that
n n
(7.12.4)

By Theorem 4.6(2),

M = (~spanxi) 1. = (span{xl,x2, ... ,Xn})1..

But the finite-dimensional subspace span{xl, X2, ... , xn} is Chebyshev by Theorem
3.8, and Theorem 5.8(1) implies that the subspace M = (span{ Xl, X2, ... ,Xn})1. is
also Chebyshev. •
Next we verify a formula for the distance to a finite-codimensional subspace. It
is a nice application of Theorem 5.8, where knowledge of M1. was used to compute
PM(x) and d(x, M).
136 ERROR OF APPROXIMATION

7.13 Distance to Finite-Codimensional Subspace. Let


{Xl, X2, ... , xn} be a linearly independent set in X and

M = n
n

1
{x E X 1 (x, Xi) = O}.

Then M is a Chebyshev subspace of codimension n, and for each x E X,


n
(7.13.1) PM(X) = x - L D:iXi,
1
where the scalars D:i satisfy the normal equations
n
(7.13.2) LD:i(Xi,Xj) = (x,Xj) (j=1,2, ... ,n),
i=l

t
and

(7.13.3) d(x,M)2 = D:i(X,Xi) = IIxI1 2 _ g(X1,X2, ... ,xn,x) = (3TG- 1(3,


1 g(X1, X2,···, xn)
where G = G(X1, X2, ... , xn) and (3T = «x, Xl), (x, X2)" .. , (x, xn))·
Moreover,
2:=~ I(X,Xi)1 2
if X tJ-M,
(7.13.4) d(x, M) = { ~ 2:=~ (x, Xi)X;!I
if X E M.
In particular, if {Xl, X2, ... , Xn} is an orthonormal set, then for every x EX,
n
(7.13.5) PM(x) = x - L(X,Xi)Xi,
1
and

(7.13.6)

Proof. By Theorem 7.12, M is a Chebyshev subspace of codimension nand


M.l = span{x1,x2, ... ,xn }. Applying Theorem 7.6 to the subspace M.l (instead
of M), we obtain that for every x E X,
n
(7.13.7) PM-L{x) = LD:iXi,

where the scalars D:i satisfy the equations (7.13.2), and


(7.13.8)
d(x, M.l)2 = g(X1,X2, . .. , Xn , x) = IIxI1 2 -
g(X1,X2, ... ,Xn)
t 1
D:i(X, Xi) = IIxI1 2 - (3T G- 1(3.

In addition, if x 'Ie M, then


(7.13.9) d(x,M.l)2 = IIxl12 _ (2:=~I(X'Xi)12):.
112:=1 (x, Xi)X;!l
But by Theorem 5.8, PM(x) = x -- PMJ.(X) and d(x, M)2 = IIxI1 2 - d(x, M.l)2.
Combining this with (7.13.7)-(7.13.9) completes the proof. •
FINITE-CODIMENSIONAL SUBSPACES 137

7.14 Exrunple. Let

M = { y E C2[O, IJ 111 yet) dt = 0 = 11 ty(t) dt }.

We will determine PM(x) and d(x,M) when x(t) = t 2.


First observe that we can rewrite M in the form

M = {y E C2 [O, 1]1 (y, Xi) = 0 (i = 1,2) },

where Xi(t) = ti-1. By Theorem 7.13, M is a Chebyshev subspace of co dimension


2; also,

(7.14.1)

and

(7.14.2)

where the O:i satisfy the equations

0:1(X1,X1) + 0:2 (X2, Xl) = (X,X1),


(7.14.3)
0:1(X1,X2) +0:2(X2,X2) = (X,X2).

1 1
0:1 + 2"0:2 = 3'

1 1 1
2"0: 1 + 30:2 = 4'
whose solution is 0:1 -i, 0:2 = 1. Substituting this back into (7.14.1) and
(7.14.2), we obtain

or
2 1
PM(X)(t) = t - t + 6'
and d(x,M) = V7j6.
Now we give an application of Theorem 7.13 to characterize best approximations
from translates of finite-co dimensional subspaces.
Recall that a translate of a subspace is called an affine set (or a linear variety
or a flat).
7.15 Best Approximation from Finite-Codimensional Affine Sets. Let
{Xl, X2, ... , xn} be linearly independent in X, {C1, C2, ... , cn } in JR, and

n{
n
v = yE X I (y, Xi) = cd·
i=l
138 ERROR OF APPROXIMATION

Then V is a Chebyshev affine set, and for each x EX,


n
(7.15.1) Pv(x) = X - I>XiXi,
1

where the scalars OCi satisfy the (modiEed) normal equations


n
(7.15.2) LOCi(Xi,Xj) = (x,Xj) - Cj (j=1,2, ... ,n),
i=l

and
n ] 1/2
(7.15.3) d(x, V) = [ ~ OCi( (x, Xi) - Ci)

Moreover,
~~[(X,Xi)-Ci]2
{ ~ ~~[(X,Xi) - if X tf. V,
(7.15.4) d(x, V) = Ci]Xill
if X E V.
In particular, if {Xl, X2, ... , Xn} is an orthonormal set, then for all x EX,
n
(7.15.5) Pv(x) = X - L[(X,Xi) - Ci]Xi

and
n ] 1/2
(7.15.6) d(x, V) = [ ~(X'Xi) - Ci)2

Proof. Since {X1,X2, ... ,Xn } is linearly independent, g(X1,X2, ... ,Xn ) i= 0 by
Lemma 7.5. This implies that there is an element v E span {Xl, X2, ... , xn} that
lies in V. In particular, V i= 0, and it is easy to verify that V = M + v, where
n
M = n{y E X I (y, Xi) = O}.
1

By Theorem 7.13, M is a Chebyshev subspace of codimension n. The remainder


of the proof follows immediately from Theorems 7.13 and 2.7. For example, V is
Chebyshev and
n n
Pv(x) = PM+v(x) = PM (X - v) +v = x - v - LOCiXi + V = X - LOCiXi,
1 1

where the OCi satisfy the equations


n
L Qi(Xi, Xj) = (x - v, Xj) = (x, Xj) - Cj (j = 1,2, ... , n). •
1

Using Theorem 7.15, we can now solve Problem 4 of Chapter 1.


THE WEIERSTRASS APPROXIMATION THEOREM 139

Problem 4. (A control problem) The position B of the shaft of a dc motor


driven by a variable current source u is governed by the differential equation

(4.1) B"(t) + B'(t) = u(t), B(O) = B' (0) = 0,


where u(t) is the field current at time t. Suppose that the boundary conditions are
given by

(4.2) B(1) = 1, B'(I) = 0,

and the energy is proportional to fo1 u 2(t)dt. Find the function u having minimum
energy in the class of all real continuous functions on [0,1] for which the system
(4.1) and (4.2) has a solution B.
In Chapter 2 we saw that this problem could be reformulated as follows: If
Xl (t) = et - 1, X2(t) = 1 - Xl (t), and
V = {u E C 2[0, 1]1 (U,Xl! = 0, (U,X2) = I},
find Pv(O).
By Theorem 7.15, Pv(O) = -(a1xl + a2x2), where the ai satisfy

al (Xl, Xl) + a2(X2, Xl) = 0,


ell (Xl, X2) + el2 (X2' X2) = -1.

After a bit of algebra, we obtain


e-l l+e
al = 3 - e and a2=---'
3-e
Hence
1-e (l+e)
Pv(O) = -3-Xl
-e
+ -3--
-e
X2 ,
or
1
Pv(O)(t) = --(1 + e - 2e t ).
3-e
The Weierstrass Approximation Theorem
In this section we give an elementary proof of one of the most important approxi-
mation theorems known: the classical Weierstrass (pronounced "VYE-er-shtrahss")
approximation theorem. It states that any real continuous function on a finite closed
interval can be uniformly approximated by polynomials. A simple corollary is that
the subspace of polynomials is dense in C2 [a, b].
The following simple inequality will be used a few times in the proof.
7.16 Bernoulli's Inequality. For any real number h 2: -1,

(7.16.1) (1 + h)n 2: 1 + nh (n = 1,2, ... ).

Proof. We proceed by induction on n. For n = 1, the result is obvious.


Assuming that (7.16.1) holds for some n 2: 1, we see that

(l+h)nH = (l+h)(1+h)n 2: (1+h)(1+nh) = 1+(n+1)h+nh2 2: 1+(n+1)h.


Thus the result is valid for n + 1. •
140 ERROR OF APPROXIMATION

7.17 Lemma. For each n E N, the function

is a polynomial on [0,1] with the properties that 0 qn «: «:


1, qn(t) ---+ 1 uniformly
on [0, c] for any 0 < c < ~, and qn(t) ---+ 0 uniformly on [d,l] for any ~ < d < 1.
Proof. Clearly, qn is a polynomial that is decreasing on [0,1], qn(O) = 1, and
qn(1) = O. Thus 0 qn«: «:
1. Fix any c E (O,~) and d E (~, 1).
«: «:
For each 0 t c, we use Bernoulli's inequality to obtain

Thus qn ---+ 1 uniformly on [0, c].


For any d t «: «:1, we obtain 0 «: qn(t) «: qn(d) and, using Bernoulli's inequality,
1
as n ---+ 00.
qn(d)

This proves that qn ---+ 0 uniformly on [d,l]. •


7.18 Lemma. Let s be the step function defined on [-1,1] by

0 if - 1 «: t < 0,
t '= { 1 if O«:t«:1.
s ().

Then for each E > 0 and p > 0, there exists a polynomial q such that 0 q(t) 1 «: «:
«: «: «:
for all It I 1, Is(t) - q(t)1 < E if p It I 1, and Is(t) - q(t)1 1 if It I < p. «:
Proof. Let qn be the polynomials defined in Lemma 7.17. Define Pn on [-1, 1]
by Pn (t) = qn e;-t). It follows from Lemma 7.17 that 0 Pn «: «:
1, Pn ---+ 1 uniformly
on [p, 1], and Pn ---+ 0 uniformly on [-1, - pl. Thus for n sufficiently large,

Is(t) - Pn(t) I < E on [-1, -p] U [p, 1]

and
Is(t) - Pn(t) I «: 1 for all t.
Thus q = Pn works for n sufficiently large. •
7.19 Weierstrass Approximation Theorem. Let J be a real continuous
function on [a, b]. Then for each E > 0, there exists a polynomial P such that

(7.19.1) sup If(t) - p(t)1 < E.


tE[a,b]

Proof. Assume first that [a, b] = [0,1]. By replacing f with J - J(O), we may
assume J(O) = O. We may also assume E < 1. Since J is continuous on [0,1]' it is
uniformly continuous by Theorem 5.4. Hence we may choose a partition

0= to < t1 < t2 < ... < tm < tm+l = 1

such that If(t) - f(t')1 < E/3 whenever t,t' E [tj _ 1,tj ] (j = 1,2,oo.,m + 1).
Choose any p > 0 such that 2p < minlSiSm+l(ti - ti-l). Let Ej = J(tj) - f(tj-l)
THE WEIERSTRASS APPROXIMATION THEOREM 141

(j = 1,2, ... , m + 1), get) = L~+1 EjS(t - t j ), and pet) = L~+1 Ejq(t - tj ), where
q is the polynomial whose existence is guaranteed by Lemma 7.18 with c replaced
by E/(3m).
Then p is a polynomial and IEj I < E/3 for all j. If t E [t i - 1 , ti), then set - tj)
°
for all j :s: i - I and set - tj) = if j :c:: i. Hence
= 1

IJ(t) - g(t)1 = IJ(t) - ~1 EjS(t - tj) I = IJ(t) - ~ Ej I = IJ(t) - J(t i - 1 )1 < i'
Also, IJ(tm+l) - g(tm+l) I = IJCtm+ 1 ) - J(tm+l) I = 0.
If t E (tk - p, tk + p) for some k, then, using Lemma 7.18, we obtain

m+l
Ip(t) - g(t)1 = L Ej[q(t - tj) - set - tj)]
j=1

:s: L [Ejllq(t - tj) - set - tj)1 + IEkl· 1


jik

<" !:. . _E_ + !:. = E2 +!:. < 2E.


- ~ 3 3m 3 9 3 3

If t tf. (tk - p, tk - p) for any k, then a slight modification of the above argument
shows that
Ip(t) - g(t)1 < _.
E E(m + 1) < -.
E
3 3m 3
Thus, for any t E [0,1],

IJ(t) - p(t)1 :s: IJ(t) - g(t)1 + Ig(t) - p(t)1 < c.


This proves (7.19.1) when [a, b] = [0,1]. In the general case, if J is continuous on
[a, b], define F on [0,1] by F(t) = J((b - a)t + a). Then F is continuous on [0,1],
so by the first part of the proof, given E > 0, there exists a polynomial P such that
IF(t) - P(t)1 < E for all t E [0,1]. Hence IJ((b - a)t + a) - P(t)1 < E for t E [0,1].
Making the change of variable u = (b - a)t + a, we get that

IJ(u) - p(u)1 < E for all u E [a, b],

where p(u) = P(~=:) is a polynomial in u. •


7.20 Corollary. The polynomials are dense in C 2 [a, b].
Proof. Given J E C 2 [a, b] and E > 0, Theorem 7.19 allows us to choose a
polynomial p such that
E
IJ(t)-p(t)l< (b-a)l/2 foralltE[a,b].

lb lb
Then
IIJ - pl12 = IJ(t) - p(tWdt < b ~ a dt = E2. •

An inner product space X is called separable if it contains a countable dense


subset.
142 ERROR OF APPROXIMATION

7.21 Proposition.
(1) Every finite-dimensional inner product space is separable.
(2) £2(r) is separable if and only if r is countable.
(3) C2 [a, bj is separable.
In particular, £2(n) and £2 are separable, but £2(~) is not separable.
Proof. (1) Let X be a finite-dimensional inner product space. Choose an
orthonormal basis {e1,e2, ... ,en } of X. Then, by (4.10.3) or Corollary 7.8, each
x E X has the representation x = L~(X, ei)ei.
Claim: D := {L~ Tiei I Ti E Q} is a countable dense set in X, where Q denotes
the set of rational numbers.
First recall that Q is a countable dense set in R. Also, there is a one-to-one
correspondence of D with Qn := Q x Q x ... x Q via the mapping L~ Tiei >-+
(TI' T2, ... ,Tn). Since Qn is a finite product of countable sets, it is countable.
Hence D is countable.
If x E X and E > 0, choose rational numbers Ti (i = 1,2, ... , n) such that
I(x,ei) - Til < Eln. Then

so D is dense in X.
(2) For each "f E r, define e-y by e-y(i) = 0 if i E r \ h} and e'Y("() = 1. Then
{e'Y I "f E r} is an orthonormal set in £2(r). In particular, Ile'Y - e'Y,1I = V2 if
"f ~ "f'. If D is a dense set in £2 (r), then, for every "f E r, there exists Y'Y E D such
that Ile'Y - Y-yll < V2/2. If "f ~ "f', we have

Ily'Y - y"!' I = Ily'Y - e'Y + e'Y - e-y' + e'Y' - Y'Y' I


:::: Ile'Y - e'Y' 1I-II(y'Y - e'Y) + (e'Y' - Y'Y' )11
:::: v'2 - (1Iy'Y - e'Yll + Ile'Y' - y-y,II) > O.
That is, Y'Y ~ Y'Y'. It follows that D contains at least as many elements as r. Hence
if r is uncountable, so is D. Thus £2 (r) is not separable.
Note that every x E £2(r) may obviously be written in the form x =
L'YEI'(X, e-y)e'Y' since the sum on the right clearly has the same value as x (namely,
x("() = (x, e'Y)) at each point of r.
If r is finite, then £2(r) is finite-dimensional and hence separable by (1). If
r is count ably infinite, say r = {"fn I n = 1, 2, ... }, then for every x E £2 (r),
we have x = limn -+ oo L~=l (x, e'Y,)e'Yi. It follows that the set D := Ul" Dn, where
Dn := {L~ Tie'Yi I Ti E Q}, is dense in £2(r). Since each Dn is countable (by the
same proof as in (1)) and D is the union of a countable number of countable sets,
D must also be countable. Hence £2 (r) is separable.
(3) To prove (3), we will just outline the steps and leave to the reader the
verification of the details. First, the set of all polynomials P is dense in C 2 [a, bj
by Corollary 7.20. Then observe that the set of all polynomials with rational
coefficients, PIQ, is dense in P, hence dense in C2 [a, bj. Finally, the set PIQ is
countable. Thus C 2 [a, bj is separable. •
MUNTZ'S THEOREM 143

Muntz's Theorem
For the remainder of the chapter we will study an interesting generalization of
Corollary 7.20 due to Muntz. Recall that Corollary 7.20 implies that the space of
polynomials
P = span {I, t, t 2 , ... , tn, .. . }
forms a dense subspace in C 2 [0, 1]. A natural and interesting question is whether
there are smaller subspaces of P that are also dense in C 2 [0, 1].
For example, what can we say about the subspace

where rn denotes the nth prime number, or the subspace

(Miintz's theorem will imply that the former is dense in C 2 [0, 1], but the latter is
not.)
Fix any sequence of real numbers {rn} (not necessarily integers) with

o :::; ri < r2 < ... < r n < ...


and rn -t 00, and let

What are necessary and sufficient conditions on the exponents {rn} to ensure that
M is dense in C 2 [0, 1]7 The elegant answer to this question is the theorem of
Miintz, which states that M is dense in C 2 [0, 1] if and only if 2:~(l/rn) = 00.
One intriguing aspect of this result is that it connects two seemingly unrelated
properties. Namely, it connects the topological property of denseness with the
number-theoretic property of the divergence of the sum of reciprocal exponents.
To simplify the proof of Miintz's theorem, we divide the steps into four lemmas
each of which has independent interest.
7.22 Cauchy Determinant Formula. Fix an integern 2=: 1, and let ai, bi, (i =
1,2, ... , n) be 2n real numbers such that ai + bj i= 0 for all i and j. Define

1 1 1
ai + bi ai + b2 ai +bn
1 1 1
a2 + bi a2 +b2 a2 + bn
Dn :=

1 1 1
an + bi an +b2 an +bn
Then

(7.22.1)
144 ERROR OF APPROXIMATION

Proof. The proof is by induction. For n = 1, Dl = 1/(al + b1 ). This is also


the value of the the right side of (7.22.1) when n = 1 (since a product over the
empty set of indices is 1 by definition). Next assume that (7.22.1) is valid when n
is replaced by n - l. Then subtract the last row of Dn from each of the other rows
and remove common factors. This yields

1 1 1
al + b1 al + b2 al + bn
D _ I17:11(an -ai)
n - I1?=1 (an + bj ) 1 1 1
an-l + b1 an-l + b2 an-l + bn
1 1 1

Now subtract the last column from each of the others and remove common factors.
This yields

1 1
1
al + b1 al + bn- 1
D _ I17:!(an - ai)(bn - bi )
n- I1?=1 (an + bj ) I17:11(ai + bn ) 1 1
1
an-1 + b1 an-l + bn- 1
0 0 1

I1~ll(an - ai)(bn - bi ) D = I1l<i<j<n(aj - ai)(bj - bi )


I1?=1(a n +bj )I17=11(ai+ bn) n-l I1~j=l(ai+bj) ,

which verifies (7.22.1). This completes the induction and the proof. •
Let {r n} be a sequence of numbers satisfying

(7.22.2) o :s; rl < r2 < ... < rn < . ", lim rn = 00,
n->oo

and define for each n ;::: 1,

(7.22.3)

and

(7.22.4)

7.23 Distance from t m to Mn. Let Mn be the subspace defined as in (7.22.3).


Then for any integer m ;::: 0,

(7.23.1) d(tm M ) = 1 lIn h - ml


, n -12m + 1 j=l rj + m + 1 .
MUNTZ'S THEOREM 145

Proof. By Theorem 7.6, we have that

(7.23.2)

Since

it follows that
1 1 1

1 1 1
Tn + Tl + 1 Tn+ T2 + 1
Applying Proposition 7.22 (with ai = bi = Ti + ~), we obtain

9
(t Tl , tT2 , ... , tTn) = rrl<i<j<n (Tj - Ti)2
rrn ( ) .
i,j=l Ti Tj + +
1

Similarly,

(t Tl tT2 tTn t m ) = rrl<i<j<n(Tj - Ti)2 rr~=l (m - Tk)2


9 , , ... " (
2m + 1) rrni,j=l ( Ti + Tj + 1) rrnk=l ( m + Tk + 1)2 .
Substituting these expressions into (7.23.2), cancelling the common factors, and
taking the square root, we obtain (7.23.1). •
The next lemma relates the divergence of an infinite product with the divergence
of an infinite series.
7.24 Lemma. Suppose 0 < an i= 1 for every n and an --+ O. Then
n
(7.24.1) (i.e., II (1 - ai) --+ 0)

if and only jf
00 n
(7.24.2) L a i = 00 (i.e., Lai --+ (0).

Proof. By omitting finitely many of the ai's, we may assume that an < 1 for
all n. Using L'Hospital's rule,

lim log(l - t) = lim ~ = -l.


t-+O t t-+O 1 - t
146 ERROR OF APPROXIMATION

Thus for each 0 < E < 1, there exists an integer N such that

or
-(1 + E)a n < log(l - an) < (-1 + E)a n ,
for all n ::::: N. This implies that 'L~ ai -t 00 if and only if 'L~ 10g(1 - ai) -t -00.
But log fE(1 - ai) = 'L~ 10g(1 - ai) -t -00 if and only if I1~(1 - ai) -t O. This
completes the proof. •
Our last lemma shows that the denseness of M is equivalent to being able to
approximate the monomial t m for each integer m ::::: O.
7.25 Lemma. Let the exponents {Tn} and the subspace M of O2 [0, 1] be given
as in (7.22.2) and (7.22.4), and let Mn be the subspace defined as in (7.22.3). Then
M is dense in O2 [0,1] if and only if

(7.25.1) lim d(tm,Mn) = 0


n--+oo

for each integer m ::::: 0 with m rf {Tl' T2," ., Tn,· .. }.


Proof. Let M be dense in O2 [0,1], m a nonnegative integer, and x(t) = tm.
Given any E > 0, choose an element y E M such that Ilx - yll < E. Since M =
Ul" Mn and Mn C M n+1 for all n, y E Mn for n sufficiently large. Hence for n
large,
d(x, Mn) ::::; Ilx - yll < E.
This verifies (7.25.1).
Now suppose (7.25.1) holds for each integer m ::::: 0 with m =1= Ti for every i. If
m = Ti for some i, then t m E Mn for all n ::::: i, so (7.25.1) holds for this case
also. That is, (7.25.1) holds for every nonnegative integer m. Next let p be any
polynomial. Then pet) = 'L~ aiti for some scalars ai, and using Theorem 2.7 (2),
we see that
(7.25.2)

d(p,Mn) = d (taiti'Mn) ::::; td(aiti,Mn) = t laild(ti,Mn) -t 0


0 0 0

as n -t 00 by (7.25.1). Finally, let x E O2 [0,1] be arbitrary and E > O. By Corollary


7.20 there exists a polynomial p such that Ilx - pll < E/2. By (7.25.2) there exists
an integer N such that d(p, Mn) < E/2 for all n ::::: N. Using Theorem 5.3, it follows
that for n ::::: N,

d(x, M) ::::; d(x, Mn) ::::; Ilx - pll + d(p, Mn) < E.
This proves that M is dense in O2 [0,1]. •
MUNTZ'S THEOREM 147

7.26 Mootz's Theorem. Let {Tn} be a sequence of real numbers with


o : : ; Tl < T2 < ... < Tn -+ 00,
let Mn be defined by

and let
UM - span{tTl , tT2 , ... , t Tn , . .. }.
00

M ..-
- n -
1
Then the following statements are equivalent:
(1) M is dense in C 2[O, 1];
(2) limn~ood(tm,Mn) = 0 for each nonnegative integer m (j: {Tl,T2, ... };
(3) limn~oo d(tm, Mn) = 0 for some nonnegative integer m (j: {Tl, T2, ... };
(4) 2::::"=2 (If Tn) = 00.
Proof. The equivalence of (1) and (2) is just Lemma 7.25, while the implication
(2) =? (3) is obvious.
(3) =? (4). Suppose liilln~ood(tm,Mn) = 0 for some nonnegative integer m
not in {Tl' T2, ... }. Now, for each integer n 2: 1,

(7.26.1) fr
1
h - ml
Tj+m+l
= IT 11 -
1
ajl,

where aj = (2m + l)f(Tj + m + 1). Note that 0 < aj f= 1 for every j and aj -+ O.
By Lemma 7.23 and (7.26.1), we see that

(7.26.2) lim rr 11- ajl


n~oo
= O.
By Lemma 7.24, (7.26.2) is equivalent to
00

(7.26.3) Laj = 00.


j=1

That is,
00 1
~~~ ~ =00.
L...
j=1 J
T ·+ m +l

For j sufficiently large, Tj > m + 1, so that


1 1 1
(7.26.5) - < <-.
2Tj Tj +m+ 1 Tj
By the comparison test for infinite series, it follows that (7.26.4) holds if and only
if( 4) holds.
(4) =? (2). Suppose that 2:7=21fTj = 00. Then for every nonnegative integer
m not in {Tl' T2, .. . }, (7.26.5) holds for j sufficiently large. By retracing the steps
of the proof of the implication (3) =? (4), we deduce that

lim rrn ITj - ml = O.


n .... oo.
J=1
TJo+m +1
By Lemma 7.23, this implies limn.... oo d(tm, Mn) = O. This proves (2) and completes
the proof. •
148 ERROR OF APPROXIMATION

7.27 ReIIlark. While the Weierstrass approximation theorem (Corollary 7.20)


implies that the polynomials are dense in C 2 [a, b] for any interval [a, b], Muntz's
theorem is not generally valid in C 2 [a, b]. For example, consider the subspace of
C2 [-1, 1] defined by

M := span {1, t 2 , t 4 , ... , t 2n , ... }.

In particular, all functions in M are even. Also, L:~ 1/(2n) = 00. However, taking
x(t) = t and any p E M, we see that p is an even function so that

Hence d(x, M)2 2': ~ so that M is not dense in C2[-1, 1].


7.28 Exam.ples. (1) The subspace of all polynomials

P = span { 1, t, t 2 , ... , tn, ... }

is dense in C 2 [0, 1], since L:~ l/n = 00. Thus we recover Corollary 7.20 again.
However, keep in mind that Corollary 7.20 was used in the proof of the Muntz
theorem. It would be of some interest to give a direct proof of Muntz's theorem
that does not use the Weierstrass approximation theorem.
(2) span {t 2 , t 4 , . .. ,t2n , ... } is dense in C 2 [0,1], since L:~ 1/(2n) = 00.
(3) M := span {1, t 2 , t 4 , t S , . .. ,t2n , . .. } is not dense in C 2 [0,1] since L:~ 1/(2n)
= 1 < 00. In fact, by Theorem 7.26, if m is any positive integer such that m # 2n
for any integer n, then t m cannot be approximated arbitrarily well by elements of
M.

Exercises

1. Verify that formula (7.1.1) (respectively (7.1.2)) is also valid when x E K,


provided that the maximum is taken over all x* E X* with Ilx* I :s: 1
(respectively z E X with Ilzll :s: 1) rather than just Ilx*11 = 1 (respectively
Ilzll = 1).
2. Let X = [2(2) and

K = {y E X I y(2) 2': e-y(l)}.


In other words, K is the "epigraph" of the function t f-t e- t .
(a) Show that K is convex and closed.
°
(b) For z = (0, -1), show that SUPYEK(Y, z) = 0, yet (y, z) < for all y E K.
(c) Show that the "supremum" (respectively "infimum") in (7.1.1) or (7.1.2)
EXERCISES 149

(respectively (7.2.2)) cannot be replaced by "maximum" (respectively "min-


imum"). [Hint: Part (b).]
3. Let {Xl, X2, ... , xn} be a basis for the n-dimensional subspace M. Show
that that expression

g(XI,X2, ... ,xn,x)


g(XI' X2, ... , xn)
depends only on X and not the particular basis chosen for M.
4. Give an alternative proof of Lemma 7.10 by showing that L contains a set
of n linearly independent vectors if and only if N does.
5. Let {Xt,x2"" ,x~} be a linearly independent set in X* and

M = n{ n

I
xEX I x;Cx) = o}.

Verify the following distance formula:

for every x E X. How does this generalize formula (7.13.3)?


6. Find the best approximation to
(a) x(t) = et
and
(b) x(t) = sint
from the set

where XI(t) = 1 and X2(t) = t.


7. Let {xi, X2, ... , x~} be linearly independent in X*, {CI' C2, ... , cn } C JR,
and
v= n{
n

I
y E X I x;(y) = cd·
Show that:
(a) V is a nonempty closed affine set.
(b) V = M +v, where

M = n{ n

I
yEX I x; (y) = 0}

and v E V is arbitrary.
(c) V is Chebyshev if and only if each xi has a representer in X.
[Hint: Theorem 7.12.]
8. Let {Xl, X2, ... , xn} be linearly independent in X, {CI' C2, ... , cn } C JR, and
n
150 ERROR OF APPROXIMATION

(a) Prove that the element of smallest norm in V is given by Vo = E~ Q;iXi,


where the scalars Q;i are chosen such that

(j = 1,2, ... ,n).

(b) Show also that

9. Let M be a subspace of the inner product space X.


(a) Show that M is dense if and only if x* = 0 whenever x* E X* and
x*(y) = 0 for all y E M. Now suppose that X is a Hilbert space.
(b) Show that M is dense in X if and only if MJ. = {O}. That is, (x, y) = 0
for all y E M implies x = O.
(c) Give an example to show that if X is not complete, then (b) is false.
[Hint: Let x* E X* have no representer and set M = {x E X I x*(x) = O}.
If MJ. i= {O}, some point off M has a best approximation in M. Deduce
that M is Chebyshev, which contradicts x* not having a representer.]
10. Consider the subspace

of C2 [-I, 1].
(a) Prove that if x is an even function in C 2 [-I, 1], then x can be approxi-
mated arbitrarily well by elements in M. [Hint: Reduce this to a problem
in C2 [0, 1].]
(b) Prove that if x is an odd function (not identically zero), then x can not
be approximated arbitrarily well by elements of M. [Hint: Look at Remark
7.27.]
11. Which ofthe following subspaces are dense in C2 [0, I]? Justify your answers.
(a) span {t, t 3 , t 5 , .•. , t 2n - \ ... }.
(b) span {t a , t a +! , t a +2 , ... , t a +n , ... }, where Q; = 10100. Are there any
values of Q; > 0 that change your answer?
(c) span{t,t4 ,t9 , .•. ,tn ·, ... }.
(d) span {t, t...i2, t..J3, ... , t..fii, ... }.
(e) span {t P1 , tP., •.. , t Pn ••• }, where Pn is the nth prime number.
[Hint: You may use the prime number theorem, which implies that
limn .... oo(nlogn)/Pn = 1.]
12. Unlike the proof of the Weierstrass approximation theorem that we gave in
Theorem 7.19, the "usual" proof uses Bernstein polynomials and is based
on the following steps.
(a) Verify the "binomial theorem":
HISTORICAL NOTES 151

where 8 and t are any real numbers.


e
(b) For any x E 2 [0, 1], the Bernstein polynomial of degree at most n
for x is defined by

Verify that Bnx E P n , Bneo = eo for n 2'" 0, Bne1 = e1 for n 2'" 1, and
Bne2 = e2 + (eI/n) (eo - e1) for n 2'" 2, where ek(t) = tk.
e
(c) For any x E 2 [0, 1] and E > 0, show that for n sufficiently large,

IBnx(t) - x(t)1 < E for all t E [0,1].

[Hint: Since x uniformly continuous, there exists 15 = I5(E) > 0 such that
Ix(t) - x(8)1 < E/2 whenever It - 81 < 15. Choose any n > Ilxll oo /EI5 2 , where
Ilxll oo := maxa:s;t9Ix(t)l, and write

Ix(t) - Bnx(t)1 = I~ [X(t) - x (~)] G)tk (l- W- k \


::; ~)t) + ~)t),
1 2

where

~)t) : = L \x(t) - x (~) \ (~)tk(l- tt- k ,


1 {kll~-tl<o}

L(t) : = L \x(t) - x (~) \ G)t k (l- t)n-k.


2 {kll~-tl2:o}

Verify that 2.:1 (t) < E/2 and 2.:2(t) < E/2 for any t E [0,1]. To verify the
second inequality, it may help to observe that

and the latter sum may be evaluated using part (b).]


e
(d) For any x E 2 [0, 1], Bnx(t) -+ x(t) uniformly on [0,1].
e
(e) For any x E 2 [a, b], define f(8) := x((b-a)8+a) for 8 E [0,1] and apply
part (d) to f. From this we deduce that there is a sequence of polynomials
that converges uniformly to x on [a, b].

Historical Notes
The first statement of Theorem 7.1, in any normed linear space, can be found
in Nirenberg (1961; p. 39). The distance formula (7.6.1) in terms of Gram deter-
minants was given by Gram (1879) (see also Gram (1883)). The characterization
of finite-codimensional Chebyshev subspaces (Theorem 7.12) was established by
Deutsch (1982; Theorem 3.3).
152 ERROR OF APPROXIMATION

The important and fundamental Weierstrass approximation theorem (Theorem


7.19, which we abbreviate WAT) was first proved by Weierstrass (1885) using a
singular integral. He also deduced, as a consequence of this, that every 21r-periodic
continuous function on the real line can be uniformly approximated by trigonomet-
ric polynomials. This is sometimes called the "second Weierstrass approximation
theorem." Conversely, it can be shown that the second Weierstrass approximation
theorem implies the WAT, so that they are logically equivalent. Runge (1885),
(1885a) had all the ingredients for a proof of the WAT at his disposal. He showed
that every continuous function on [a, b] can be uniformly approximated arbitrarily
well by rational functions (Runge (1885a)). But he had already established the
tools in his 1885 paper that he could replace these rational functions by polynomi-
als! Thus, either by overlooking this fact or not thinking it was important enough
to include, Runge missed out on the opportunity to independently establish the
WAT in the same year as Weierstrass.
Picard (1891) gave an alternative proof of the WAT, and a generalization to a
continuous function of n variables. Lerch (1892) (see also (1903)) showed that the
WAT could be deduced from Dirichlet's proof of a theorem of Fourier.
Lebesgue (1898) gave a proof of the WAT by reducing it to first uniformlyap-
proximating the function f(t) = It I on the interval [-y0, V2] by polynomials. Then
any piecewise linear continuous function on [0, 1] can be uniformly approximated
by polynomials, and finally, any continuous function on [0,1] can be uniformly ap-
proximated arbitrarily well by polynomials. Fejer (1900) gave a proof of the WAT
by showing that the averages, or Cesaro sums, of the partial sums of the Fourier
series of a continuous 21r-periodic function on the real line converge uniformly to the
function. Mittag-LefHer (1900) gave a proof of the WAT that was similar to that
of Runge and Lebesgue in the sense that the key step was reducing the problem to
approximating piecewise linear functions. Landau (1908) gave a proof of the WAT
based on singular integrals. Inspired by Landau's proof, de la Vallee-Poussin (1908)
gave a proof of the second Weierstrass approximation theorem based on singular
integrals.
Jackson (1911) showed that if f is a 21r-periodic Lipschitz continuous function,
with Lipschitz constant A, then the error

d(f, Tn):= inf max If(t) - T(t)1


TETn tErn;

°
is bounded above by a constant Kn that depends only on A, and Kn --t as n --t 00.
(In (1938), Achieser and Krein showed that Kn = 1rA/(2(n+ 1)) is the best possible
constant.)
Bernstein (1912), motivated by a probabilistic interpretation, gave a fairly ele-
mentary proof of the WAT by giving an explicit construction of the polynomials
Bn = Bn(f) such that Bn converges to f uniformly on [0,1]. However, in practice,
the Bernstein polynomials are rarely used to approximate functions, since their
convergence is generally very slow. Indeed, the Bernstein polynomials Bn(f) for
the function f(t) = t 2 converge to f like lin (see, e.g., Davis (1963; pp. 116)).
Korovkin (1953) (see also Korovkin (1960)) noted that the crux of Bernstein's
proof was in verifying that (Bn) is a sequence of positive linear operators with the
property that (Bn!) converges to f uniformly on [0,1] whenever f(t) = 1, t, or t 2 .
This led Korovkin to establish a few elegant generalizations, which, in turn, has
spawned a whole area of so-called theorems of Korovkin type. Bohman (1952) had
HISTORICAL NOTES 153

proved similar theorems a year earlier. But it was not until fairly recently that
Bohman's contribution became widely known. Now most authors correctly refer to
these results as "Bohman-Korovkin" theorems. One of the most elementary proofs
of the WAT was given by Kuhn (1964). This is the proof given in this book (7.16-
7.19) based on first approximating step functions. A nice history of the classical
proofs of the WAT can be found in the (unpublished) expository masters thesis
of Cakon (1987). More recent, and more accessible, is the excellent exposition on
Weierstrass's work in approximation theory by Pinkus (2000).
Normally, one would expect that a proof of the WAT could be attained through
Lagrange interpolating polynomials. However, Faber (1914) showed that no matter
how the nodes for interpolation are prescribed, there is a continuous function f
whose interpolating polynomials fail to converge uniformly to f on [a, b]. Despite
this, I<ejer (1930) gave a proof of the WAT using Hermite interpolation. Specifically,
the interpolating polynomials whose derivatives are all zero at the nodes do converge
uniformly to the continuous function.
Perhaps the most important and far-reaching generalization of the WAT is the
result of Stone (1937) (see also Stone (1948)), which now goes by the name of the
Stone·-Weierstrass theorem. A fine exposition of this theorem and its extensions
was given by Prolla (1993). Virtually all proofs of the Stone-Weierstrass theorem
use the WAT as a key step, or at least the special case of the WAT concerned
with the uniform approximation of the function f(t) = It I on the interval [-1,1]
by polynomials. However, Brosowski and Deutsch (1981) have given an elementary
proof of the Stone-Weierstrass theorem that does not use the WAT (nor uses the
special case mentioned above) and uses only the basic facts that a closed subset
of a compact set is compact, a positive continuous function on a compact set has
a positive infimum, and the elementary Bernoulli inequality (Theorem 7.16). In
particular, the WAT appears as a genuine corollary of this proof. Ransford (1984)
used this idea to give an elementary proof of Bishop's generalization of the Stone-
Weierstrass theorem.
Muntz's Theorem (Theorem 7.26) is due to Muntz (1914), with some subsequent
extensions (to complex exponents) and simplifications by Szasz (1916). In contrast
to the proof given here, an elementary proof of Muntz's theorem that does not use
the WAT was given by Burckel and Saeki (1983).
CHAPTER 8

GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

Linear Operator Equations


In this chapter we consider the fundamental problem of solving the linear oper-
ator equation

(8.0.1) Ax=y,
where A : X -+ Y is a bounded linear operator between the inner product spaces
X and Y, and y is a given element of Y. If X = l2(n) and Y = l2(m), then (8.0.1)
reduces to a system of m linear equations in n unknowns. Even in this special case,
however, we know that there are three possible outcomes of (8.0.1): no solution, a
unique solution, or infinitely many solutions.
In the case that there is no solution to (8.0.1), it still makes sense to try to satisfy
(8.0.1) "as best as possible." This suggests determining Xo E X so as to minimize
the expression IIAx - yll. That is, find Xo E X such that

(8.0.2) IIAxo - yll = inf IIAx - yll·


xEX

Any Xo EX satifying (8.0.2) is called a generalized solution to (8.0.1). Of course,


every solution to (8.0.1) in the ordinary sense (i.e., satisfying Axo = y) is also a
generalized solution.
We will see that if X and Yare complete and the range of A is closed, there
always exists a generalized solution. In general, however, there are infinitely many
generalized solutions. For computational purposes, we shall be interested in the
generalized solution having the smallest norm. This particular solution is unique,
and it gives rise to the notion of the "generalized inverse" of A.
To obtain useful characterizations of generalized inverses, it is convenient to
introduce the notion of the "adjoint" of a bounded linear map from X to Y. We
begin by first recalling the definition of the the inverse of a linear operator, and
establishing some fundamental properties of the inverse.
For brevity throughout this chapter, we shall denote the set of all bounded linear
operators from X into Y by B(X, Y). As was shown in Chapter 5 (Theorem 5.11),
boundedness is equivalent to continuity for linear operators. Recall also that the
norm of a bounded linear operator A E B(X, Y) is given by

(8.0.3) IIAII = sup { lilAI Xli" I X E X\{O}} = sup IIAxll = sup IIAxl1,
x IIxll9 Ilxll=l

and

(8.0.4) II Ax II .: :; IIAllllxll for all x E X.

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
156 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

Here, and in the sequel, we will often write Ax instead of A(x) whenever A is a
linear operator.
There is a useful alternative expression of the norm of an operator A E SeX, Y).
Namely,

I All = sup { I(Ax,y)1


Ilxllllyll I x E X\{O}, y E Y\{O}
}
(8.0.5) = sup{ I(Ax, y)11 x EX, Ilxll :::; 1, y E Y, Ilyll:::; I}
= sup{ I(Ax, y)11 x EX, Ilxll = 1, y E Y, Ilyll = I}.

To verifY this, note that by Schwarz's inequality,

I(Ax,y)1 < IIAxllllyl1 < IIAII


Ilxllllyll - Ilxllllyll - .
Hence
I(Ax,y)1 I } <
sup { Ilxllllyll x E X\{O}, Y\{O} _ IIAII·
On the other hand, choose Xn E X, Ilxnll = 1, such that IIAx n ll-+ IIAII. Then

implies that

IIAII :::; sup{ I(Ax, y) I I x EX, Ilxll = 1, y E Y, Ilyll = 1}


:::; sup{ I(Ax, y)11 x E X, Ilxll:::; 1, y E Y, lIyll:::; I}
I(Ax,y)1 I }
:::; sup { Ilxllllyll x E X\{O}, y E Y\{O} :::; IIAII·

This verifies (8.0.5).


Also, if A E SeX, Y) and B E S(Y, Z), then the composition BoA: X -+ Z
will often be written as juxtaposition: BA:= BoA, so that BAx = B(Ax) for all
xEX.
8.1 Lemnla. If A E SeX, Y) and BE S(Y, Z), then BA E SeX, Z) and

(8.1.1) IIBAII:::; IIBIIIIAII·

Proof. To verify the linearity of BA is a simple exercise using the linearity of


A and B. For boundedness, note that for each x EX,

IIBAxl1 = IIB(Ax)11 :::; IIBllllAxl1 :::; IIBIIIIAllllxll·

Hence BA is bounded and (8.1.1) holds. •


LINEAR OPERATOR EQUATIONS 157

8.2 Theorem. With the usual pointwise operations of addition and scalar mul-
tiplication, B(X, Y) is a normed linear space. Moreover, B(X, Y) is complete if Y
is complete.
Proof. If A, Bin B(X, Y), Xl, x2 in X, and 001, 002 in JR., we see that

(A + B)(alxl + a2x2) = A(alxl + a2x2) + B(alxl + a2x2)


= alAxl + a2Ax2 + alBxl + a2 Bx2
= 001 (AXI + BXl) + 002 (AX2 + BX2)
= 001 (A + B)Xl + 002 (A + B)X2.
Thus A +B is linear. Moreover, for any X E X,

II(A+B)xll = IIAx+Bxll :::; IIAxl1 + IIBxl1 :::; IIAllllxl1 + IIBllllxl1 = (IIAII + IIBII)llxll·
This proves that A + B is bounded and IIA + BII :::; IIAII + IIBII.
Similarly, it is easy to verify that aA is a bounded linear operator if A is, and

IlaA11 = lalllAli·
Clearly, IIAII 2> 0 and IIAII = 0 if and only if Ax = 0 for all x, i.e., A = O. This
proves that B(X, Y) is a normed linear space.
Finally, suppose that Y is complete. Let {An} be a Cauchy sequence in B(X, Y).
Then for each E > 0 there exists N = N(E) EN such that

(8.2.1 ) IIAn - Amll < E for all n, m 2> N.

For each x EX, it follows that

(8.2.2)

for all n, m 2> N. Thus {Anx} is a Cauchy sequence in Y for each x EX. Since Y
is complete, there exists a point A(x) E Y such that An(x) --+ A(x) for each x E X.
To complete the proof, it suffices to show that A E B(X, Y) and IIAn - All --+ O.
Now, for any Xl,X2 EX, 001,002 E JR., we have (using Theorem 1.12),

A(alxl + a2x2) = limAn(alxl + a2x2) = lim[alAnxl + a2AnX2]


= alA(xl) + a2 A (x2).
Thus A : X --+ Y is linear.
Passing to the limit in (8.2.2) as m --+ 00, we see that

(8.2.3) IIAnx - Axil:::; Ellxll


for all n 2> N, x E X. This implies that A - AN E B(X, Y), and hence A =
(A - AN) + AN E B(X, Y), and

(8.2.4) IIAn - All:::; E for all n 2> N . •


158 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

8.3 Dual Spaces are Complete. If X is any inner product space (complete
or not), then X' is complete.
Proof. Note that X* = B(X, JR.) and JR. is complete, so we can apply Theorem
8.2 . •
When X and Yare both finite-dimensional, there is a natural way of representing
linear operators in B(X, Y) by matrices.
8.4 Definition. Let {XI,X2, ... ,Xn } (respectively {YI,Y2, ... ,Ym}) be an or-
thonormal basis for X (respectively Y) and A E B(X, Y). The matrix of A relative
to these bases is the m x n matrix

(8.4.1)

where

(8.4.2) aij = (Axj, Yi) (i = 1,2, ... , m; j = 1,2, ... , n).

Now, each vector x E X may be written uniquely in the form x = L;


(x, Xi)Xi
by Corollary 4.10. Let us "identify" x with its (column) vector of coefficients

(8.4.3)

in l2(n). More precisely, the mapping x N ((x, Xl), (x, X2), .. . , (x, xn))T from X to
l2(n) is linear, one-to-one, onto, and preserves both the inner product and norm.
In the language of Exercise 4 at the end of Chapter 6, the spaces X and l2(n) are
equivalent.
Similarly, each Y E Y may be identified with the vector

((y, Yl), (y, Y2), ... , (y, Ym) f

in lz(m). In particular, Ax may be identified with

((Ax, YI), (Ax, Y2)," ., (Ax, Ym) f·

n n
(Ax,Yi) = :L)x,xj)(Axj,Yi) = I:>ij(X,Xj).
j=l j=l

Thus Ax can be identified with the m-tuple

((Ax, Yl),(Ax, Y2), ... , (Ax, Ym))T


LINEAR OPERATOR EQUATIONS 159

But this m-tuple is just the formal matrix product

a ln ] [(X,XI)]
a2n (x, X2)
(8.4.4)
amn (x,Xn)

For this reason, we often identify the operator A with its matrix (8.4.1), x with
the n-tuple (8.4.3), and Ax with (8.4.4). In short, we are identifying X (respec-
tively Y) with b(n) (respectively 12(m)) and the basis {Xl, X2, ... , xn} (respectively
{Yl, X2, ... , Ym}) with the canonical orthonormal basis {el' e2, ... , en} in 12(n) (re-
spectively {el' e2, ... , em} in 12(m)), where ei(j) = Oij for all i, j.
To determine the matrix of a given operator A, it is helpful to notice that AXi
can be identified with the ith column of the matrix for A:

ali
a2i
AXi= [ ....
1 T
=(ali,a2i,.·.,ami).
am,

This follows by substituting Xi for x in (8.4.4).


8.5 Example. Let {XI,X2,X3} (respectively {YI,Y2}) be an orthonormal basis
for a 3-dimensional (respectively 2-dimensional) inner product space X (respec-
tively Y). The mapping A : X -* Y is defined by

Determine the matrix of A relative to these bases.


We have AXI = YI + 3Y2 = (1,3)T, AX2 = -YI + 2Y2 = (-1,2)y, and AX3 =
Y2 = (0, l)T. Thus the matrix for A is

A=[~ -;1 ~].


8.6 Definition. The identity operator on X is the mapping I = Ix : X -* X
defined by
Ix =X, x E X.
When there is not likely to be any confusion as to the space to which we are
referring, the space subscript on the identity operator will usually be omitted.
Note that I E B(X, X) and 11111 = l. Moreover, if X is n-dimensional and
{Xl,X2, ... , xn} is any orthonormal basis in X, then the matrix of I relative to this
basis is the n x n matrix with l's on the diagonal and O's elsewhere:

o
1

o
That is, the (i, j) entry is Oij.
160 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

8.7 Definitions. IfF is any mapping from X into Y, then the range of F is
the set
R(F) = F(X) := {F(x) I x EX},
and the null space of F is the set

N(F):= {x E X I F(x) = o}.

If A E 8(X, Y), then its range and null space are linear subspaces of X and Y,
respectively. Moreover, N(A) is closed, since A is continuous. However, the range
of A need not be closed.
8.8 Example. (A bounded linear operator with nonclosed range) Define A :
12 -+ l2 by
Ax = (x(1), !x(2), !x(3), ... , !.x(n), . .. ) .
2 3 n
Then A E 8(l2' l2) and II All = 1. Clearly, R(A) contains all sequences with finitely
many nonzero components. Hence R(A) = l2. However, the vector

y= (1,~,~, ... ,~, ... )


is in l2\R(A), and thus R(A) is not closed.
Recall that a mapping F : X -+ Y is called injective (or an injection or one-
to-one) if F(Xl) = F(X2) implies Xl = X2. F is called surjective (or a surjection
or onto) if R(F) = Y. Finally, F is called bijective (or a bijection) if it is both
injective and surjective.
It is easy to see that a linear mapping F is injective if and only if N(F) = {o}.
(This follows from the fact that F(xI) = F(X2) B F(Xl - X2) = 0.)
8.9 Lemma. Let F be any function from X into Y. The following statements
are equivalent:
(1) F is bijective;
(2) There exists a (unique) map G : Y -+ X such that

(8.9.1) FoG = Iy and Go F = Ix.


Moreover, if F linear, then so is G.
Proof. (1) =? (2). Suppose F is bijective. Then for each y E Y there is an
x E X such that F(x) = y. Since F is injective, x is unique. Thus we can define a
function G : Y -+ X by setting

G(y) =x B F(x) = y.

Clearly,
x = G(y) = G(F(x)) = Go F(x)
for all x E X implies that Ix = Go F.
Similarly,
y = F(x) = F(G(y)) = F 0 G(y)
for all y E Y implies that FoG = /y. This proves (8.9.1).
LINEAR OPERATOR EQUATIONS 161

Next we show that G is unique. If H : Y -+ X is any function that satisfies

HoF=Ix and FoH=Iy,

then
H = H 0 Iy = H 0 (F 0 G) = (H 0 F) 0 G = Ix 0 G = G,
which proves uniqueness.
(2) 0=} (1). Suppose there is a function G : Y -+ X satisfying (8.9.1). Let
F(xd = F(X2). Then

Xl = lx(Xl) = Go F(Xl) = G(F(Xl)) = G(F(X2)) = Go F(X2) = lx(X2) = X2·

Thus F is injective.
Now let y E Y. Then

Y = Iy(y) = F 0 G(y) = F(G(y)).

Taking x = G(y) E X shows that F(x) = y. That is, F is surjective, hence bijective.
Finally, suppose F is linear. We will show that G is also linear. Let Yi E Y,
ai E R, Xi = G(Yi) (i = 1,2), and X = G(alYl + a2Y2). Then

F(x) = F[G(alYl + a2Y2)] = alYl + a2Y2


= alF(G(Yl)) + a 2F(G(Y2))
= alF(xl) + a2F(x2) = F(alxl + a2x2).
Since F is one-to-one, x = alXl + a2X2. That is,

Hence G is linear. •
For any function F : X -+ Y that is bijective, the unique function G from Y to
X satifying (8.5.1) is called the inverse of F and is denoted by F- l . Thus

F 0 F- l = Iy and F- l 0 F = Ix.

Also, by Lemma 8.9 (applied to F- l instead of F), we see that F- l : Y -+ X is


bijective, and
(F-l)-l = F.

Moreover, if F is also linear, then so is F- l .

8.10 Example. Let X be an n-dimensional Hilbert space, let {Xl, X2,"" xn}
be an orthonormal basis for X, let AI, A2,' .. ,An be n real numbers, and let the
mapping A : X -+ X be defined by
n
A(x) = LAi(X,Xi)Xi, X E X.

Then:
162 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

(a) A E 8(X,X) and the matrix of A relative to {X1,X2, ... ,xn } is

[: o
(b) IIAII = maxi IAil·
(c) A has an inverse if and only if Ai of- 0 for all i.
(d) If Ai of- 0 for all i, then A-I: X --+ X is given by

n
A- 1 (x) = LA;l(X,Xi)Xi, x E X,

and the matrix for A-I relative to {Xl, X2, ... , xn} is

(e) IIA- 1 11 = maxi II/Ail.


It is clear that A is linear, and since X is finite-dimensional, A must be bounded
by Theorem 5.16. Also,

This verifies part (a).


To verify (b), note that

, IAil 2 L l(x,xiW
:s; max = , IAil 211xI12
max
1

using the orthogonality of the x/so Thus IIAxl1 :s; maxi IAilllxll for all x, and
hence IIAII :s; maxi IAil. To see that equality holds, choose an index io such that
IAio I = maxi IAil· Then with x = Xio' we see that Ilxll = 1 and IIAxl1 = IIAioxio II =
IAio I = maxi IAi I· This proves (b).
If Ai = 0 for some i, say Al = 0, then
n n
AXI = L Ai (Xl, Xi)Xi = L Ai 61i Xi = 0
1 2

implies that N(A) of- {O} and A is not injective. Hence A has no inverse.
LINEAR OPERATOR EQUATIONS 163

Conversely, if Ai i= 0 for all i, consider the mapping B : X -+ X defined by


n
B(x) = LA;l(X,Xi)Xi, x E X.

By parts (a) and (b), B is a bounded linear map with IIBII = maxi IA;ll. Also, for
each x E X,

n n n

= LA;l(X,Xi)Axi = LA;l(X,Xi) LAj(Xi,Xj)Xj


i=l i=l j=l
n n
= LA;l(X,Xi)AiXi = L(X,Xi)Xi =x.
i=l 1

Thus AB = I. Similarly, BA = I. By Lemma 8.9, A has an inverse and B = A-I.


This proves (c), and parts (d) and (e) follow immediately.
A natural question suggested by Lemma 8.9 is the following. If A is a bounded
linear operator from X to Y that is bijective, must A-I also be bounded?
We will see that the answer is affirmative when both X and Yare complete, but
not in general.
8.11 Definition. An operator A E B(X, Y) is bounded below if there is a
constant p > 0 such that

(8.11.1) IIAxl1 2': plixil for all x E X,

equivalently, if p:= inf{IIAxlll x E X, Ilxll = I} > O. In this case, we also say that
A is bounded below by p.
8.12 LeIllIlla. Let A E B(X, Y).
(1) If A is bounded below, then A is injective.
(2) A has an inverse that is bounded if and only if A is both surjective and
bounded below.
Proof. (1) Suppose A is bounded below by p > o. IfAxl = AX2, then

implies that Xl = X2. Thus A is injective.


(2) Suppose A has an inverse that is bounded. Then A is bijective, and for each
xEX,
Ilxll = IIA-lAxll ~ IIA-lIIIIAxll·
If X i= {O}, then A-I i= 0 implies
1
IIAxl1 2': IIA -lllllx li for all x E X.

Thus A is bounded below (by IIA-lll- l ). If X = {O}, then A is trivially bounded


below (by any p > 0).
164 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

Conversely, suppose A is both surjective and bounded below by p > o. By (1),


A is injective. By Lemma 8.9, A -1 is also linear.
It remains to verify that A-1 is bounded. For any y E Y, there is an x E X such
that Ax = y. Then

This proves that A -1 is bounded and, in fact, IIA -111 :s; p-1. •

The Uniform Boundedness and Open Mapping Theorems


Thus to determine whether A has a bounded inverse requires, in particular,
checking whether A is bounded below. This may be difficult to verify in practice.
However, in the case where both X and Yare complete, there is an easily verifiable
criterion; namely, A is bijective. To prove this, we need a few facts that are of
independent interest.
Recall that a subset D of X is called dense in X if for each x E X and E > 0,
there exists y E D such that Ilx - yll < E. In other words, B(x, E) n D cF 0.
Equivalently, D n U cF 0 for every nonempty open set U in X.
8.13 Baire's Theorem. Let X be a complete subset of an inner product space.
(1) If {Dn} is a sequence of dense open sets in X, then nj'" Dn is dense in X.
(2) If {Sn} is a sequence of sets in X with X = Uj"'Sn, then Sn contains an
interior point for some n.
Proof. (1) Let D = nj'" Dn and let U be a nonempty open set in X. We
are to show that U n D cF 0. Since D1 is dense, U n D1 cF 0. Since U n D1 is
open, it follows that there is an open ball B(X1,E1) c UnD 1, where 0 < E1:S; l.
Since B[X1' Ed2] C B(X1' E1), it follows that un D1 contains the closed ball B1 :=
B[X1' Ed2]. Similarly, since B(X1' Ed2) nD2 is nonempty and open, it must contain
a closed ball B2 := B[X2' E2], where 0 < E2 :s; ~. Continuing in this way, we obtain
for each integer n a closed ball Bn := B[xn, En] contained in B n - 1 n D n , where
0< En :s; ~. For every n ~ m, we see that Xn E B m , and hence

It follows that {xn} is a Cauchy sequence. Since X is complete, there is x E X


such that Xn --+ x. For each m, the set Bm is closed and contains Xn for all n ~ m.
Hence x E Bm C Dm for every m. It follows that x E U n D.
(2) Since X is complete, it is closed, and hence X = Uj'" Sn- If the result were
false, then Sn would have no interior point for each n. Hence Dn := X\Sn would
be a dense open set for each n. But

which contradicts part (1). •


One important consequence of Baire's theorem is the following theorem, which
shows that if a collection of operators is pointwise bounded, it must actually be
uniformly bounded.
THE UNIFORM BOUNDED NESS AND OPEN MAPPING THEOREMS 165

8.14 Uniform Boundedness Theorem. Suppose X is a Hilbert space, Y is


an inner product space, and A c B(X, Y). If

(8.14.1) sup{IIAxll I A E A} < 00 for each x E X,

then

(8.14.2) sup{IIAIII A E A} < 00.

Proof. For each n E N, set

Sn:= {x E X IllAxl1 s n for each A E A}.


Then Sn is closed, and X = Ul."'Sn by (8.14.1). By Baire's theorem (Theorem
8.13), some Sn has an interior point. Thus there exist N E N, Xo E SN, and E > 0
such that B(xo, E) C SN. Thus for all A E A, IIAxl1 s N for all x E B(xo, E). In
particular, IIAxol1 S N, and if Ilx - xoll < E, then

IIA(x - xo)11 ::; IIAxl1 + IIAxol1 ::; 2N.

That is, IIAzl1 ::; 2N whenever Ilzll < E. Thus, IIAwl1 ::; 2N/E whenever Ilwll < 1.
This proves that IIAII S 2N/E for all A E A, and hence (8.14.2) holds. •
8.15 Corollary. Let X be a Hilbert space, Y an inner product space, and
An E B(X, Y) for n = 1,2, .... If

(8.15.1 ) A(x) := limAnx

exists for each x E X, then A E B(X, Y),

(8.15.2) sup IIAnl! < 00,


n

and

(8.15.3) IIAII ::; liminf IIAnll·


n

Proof. Since {Anx} converges for each x, the collection A = {An I n E N} is


pointwise bounded: sUPn IIAnxl1 < 00 for each x E X. By the uniform boundedness
theorem (Theorem 8.14), (8.15.2) holds. Further, the mapping A defined by (8.15.1)
is clearly linear, since all the An are, and

IIA(x)11 = lim IIAnxl1 = liminf IIAnxl1 ::; liminf IIAnllllxl1


n n n

for all x E X. Thus A is bounded and (8.15.3) holds. •


166 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

8.16 Lemma. Let X and Y be inner product spaces and A E B(X, Y). Then
the following statements are equivalent:
(1) A(U) is open in Y for each open set U in X;
(2) There exists 8 > 0 such that

(8.16.1) A(B(X)) ::J 8B(Y),


where B(X) (respectively B(Y)) denotes the open unit ball in X (respectively Y);
and
(3) There exists p > 0 such that for each y E Y, there is an x E X with Ax = y
and Ilxll :::; pllyll·
Proof. (1) =? (2). Suppose (1) holds. Then A(B(X)) is open. Since 0 = A(O) E
A(B(X)), there exists 8 > 0 such that

8B(Y) = B(O, 8) c A(B(X)).


(2) =? (3). If (2) holds, then
B(Y) c 8- 1 A(B(X)) = A(8- 1B(X)).

Fix any>. E (0,1). Then for every y E Y\ {O}, >'Ilyll-l y E B(Y). Thus there exists
Xo E 8- 1B(X) such that Axo = >'Ilyll-l y. Then y = A(llyll>' - l XO ), and setting
x = Ilyll>.-l XO , we see that Ax = yand Ilxll = Ilyll>.-lllxoll :::; 8- 1>.-11Iyll. Taking
p= (8).)-1, (3) follows.
(3) =? (1). Assume that (3) holds and let U be open in X. We must show that
A(U) is open in Y. Let Yo E A(U) and choose Xo E U such that Yo = Axo. Since
U is open, there exists 10 > 0 such that

Xo + lOB (X) = B(xo, E) C U.


Then
Yo + EA(B(X)) = A(xo + EB(X)) c A(U).
But by our hypothesis, for each y E B(Y), there exists x E X such that Ax = y
and Ilxll < p. That is, B(Y) C A(pB(X)) = pA(B(X)). Thus

B(yo, Ep-1) = Yo + Ep-1 B(Y)) c Yo + EA(B(X)) c A(U),

and hence Yo is an interior point of A(U). Since Yo was arbitrary, A(U) is open .
• 8.17 The Open Mapping Theorem. Let X and Y be Hilbert spaces. If A is
a bounded linear operator from X onto Y, then A(U) is open in Y for every open
set U in X. Hence all the statements of Lemma 8.16 hold.
Proof. Let U = B(X) and V = B(Y). By Lemma 8.16, it suffices to prove that
there exists 8 > 0 such that

(8.17.1) A(U) ::J 8V.


Given y E Y, there exists x E X such that Ax = y. Let k be any integer such
that Ilxll < k. It follows that y E A(kU). This proves that

Y k
= U =l A(kU).
THE UNIFORM BOUNDED NESS AND OPEN MAPPING THEOREMS 167

°
Since Y is complete, it follows from Baire's theorem that A(kU) contains an interior
point Yo for some kEN. Hence there is an TJ > such that

(8.17.2) W := B(yo, TJ) C A(kU).

For any y E Y with Ilyll < TJ, we have that y + Yo E W. From (8.17.2), we can
choose sequences {x~}, {x~} in kU such that

Ax~ -+ Yo and Ax~ -+ y + Yo·


Setting Xn = x~ - x~, we see that Ilxnll < 2k and AXn -+ y. This proves that for
each y E Y with Ilyll < TJ and for each E> 0, there exists x E X with

(8.17.3) Ilxll < 2k and IIAx - yll < E.

Claim: If 8 = TJ/(4k) , then for each y E Y and E> 0, there exists x E X with

1
(8.17.4) Ilxll :::; Jllyll and IIAx - yll < E.

To prove the claim, let y E y\{o}. Then y' = TJ(21Iyll)-ly has the property that
Ily'll < TJ. By (8.17.3), there exists x' E X with Ilx'll < 2k and IIAx' - Y'll < E',
where E' = TJ(21Iyll)-IE. Then x = 21IyIITJ- I X' satisfies Ilxll < 4kTJ- I llyll = 8- l llyll
and
IIAx - yll = 21IyIITJ- I IIAx' - Y'll < 2I1yIITJ- I E' = E.
This verifies (8.17.4). If y = 0, (8.17.4) is trivially true with x = 0.
It remains to prove (8.17.1). Fix any y E 8V and E > 0. By (8.17.4) there exists
Xl E X with IlxI11 < 1 and

(8.17.5)

Applying the claim with y replaced by y - AXI, we obtain X2 E X with

(8.17.6)

Continuing in this fashion, we obtain a sequence {x n } in X with

(8.17.7)

and
1
(8.17.8) Ily - AXI - AX2 - ... - Axnll < 2n & (n = 1,2, ... ).

Setting 8n = L~ Xi, we see from (8.17.7) that for n > m,

1 1 1 1
118n - 8m ll = Ilxm+1 + X m +2 + ... + xnll :::; 2mE+ 2m+1 E+ ... + 2n- 1E < 2m- 1E.
168 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

It follows that {sn} is a Cauchy sequence. Since X is complete, there exists an


x E X such that Sn - t x. Using (8.17.7), we get

Thus

Since A is continuous, ASn - t Ax. Passing to the limit in (8.17.8), we deduce that
Ax=y.
We have shown that A((l+E)U) :::> OV, and hence A(U) :::> (l+E)-IOV, for every
E > O. But oV = U<>o(l + E)-IOV, so that (8.17.1) holds. •

The Closed Range and Bounded Inverse Theorems

8.18 Closed Range Theorem. Suppose X and Yare Hilbert spaces and
A E B(X, Y) \ {a}. Then the following statements are equivalent:
(1) A has closed range;
(2) There exists p > 0 such that IIAxl1 ?: pllxll for all x E N(A)J.;
(3) p:= inf{IIAxlll x E N(A)J., Ilxll = 1} > O.
Proof. The equivalence of (2) and (3) is clear.
(1) *(2). Suppose A has closed range. Let Xo := N(A)J., Yo := R(A), and
Ao := Alxo' Note that Xo and Yo are closed subspaces of Hilbert spaces and hence
are themselves Hilbert spaces. Also, Ao : Xo - t Yo is clearly linear and surjective.
Finally, Ao is injective, since if AOXI = AOX2 for xl, X2 E X o , then A(XI - X2) = 0
implies Xl - X2 E N(A) n Xo = {a}. That is, Xl = X2.
Applying the open mapping theorem (Theorem 8.17) to Ao E B(Xo, Yo), we
obtain a 0 > 0 such that

(8.18.1) A(B(Xo)) :::> oB(Yo)·

Now, AOI : Yo -t Xo exists and is linear by Lemma 8.9. We deduce from (8.18.1)
that

or
I
Ao [B(Yo)] C 81 B (Xo).
It follows that

I I I I
IIAo II = sup IIAo (yo)11 .:::; 8 sup Ilyoll = 8'
YoEB(Yo) YoEB(Yo)

and hence AOI is bounded.


By Lemma 8.12, Ao is bounded below: there exists p > 0 such that IIAoxo11 ?:
pllxo II for all Xo E Xo. In other words, (2) holds.
THE CLOSED GRAPH THEOREM 169

(2) =?(1). Suppose (2) holds. To show that R(A) is closed, let Yn E R(A) and
Yn -+ y. Then Yn = AXn for some Xn E X. Write Xn = Un + Vn, where Un E N(A)
and Vn E N(A)..L. Then

as n, m -+ 00. Thus {v n } is a Cauchy sequence in N(A)..L. By completeness, there


exists v E N(A)..L such that Vn -+ v. Hence Yn = AXn = AVn -+ Av, which implies
that Y = Av E R(A), and thus R(A) is closed. •
8.19 Bounded Inverse Theorem. Let X and Y be Hilbert spaces and A E
S(X, Y). Then A has a bounded inverse if and only if A is bijective.
Proof. The necessity is immediate. For the sufficiency, assume that A is bijec-
tive. Then N(A) = {O} implies N(A)..L = X. Also, A is surjective implies that
R(A) = Y is closed. By the closed range theorem (Theorem 8.18), A is bounded
below. By Lemma 8.12, A has a bounded inverse. •

The Closed Graph Theorem


As a final application of the open mapping theorem, we prove the "closed graph"
theorem and an important consequence of it that will be used in the next two
chapters.
If X and Yare inner product spaces, let X x Y denote the set of all ordered
pairs (x, y) with x E X and Y E Y. Addition and scalar multiplication are defined
"componentwise" in X x Y; that is,

(8.19.1)

and

(8.19.2) a(x, y) := (ax, ay).

With these operations, it is not difficult to verify that X x Y is a linear space. In


fact, one can define an inner product in X x Y by

(8.19.3)

Of course, the first (respectively second) inner product on the right denotes the
inner product in the space X (respectively Y).
8.20 Proposition. X x Y is an inner product space. It is complete if and only
if both X and Yare complete.
We leave the simple verification of this proposition as an exercise. Note that the
norm in X x Y is given by

II(x,y)II:= V((x,y), (x,y)) = Vllxl1 2+ IIYI12.


In particular, convergence in X x Y is componentwise:

(8.20.1) II(xn,Yn) - (x, Y)II-+ 0 {o} Ilxn - xii -+ 0 and llYn - YII -+ O.
170 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

8.21 Definition. The graph of a mapping F : X -+ Y is the set

Q(F) := {(x,F(x» E X x Y Ix EX}.

F is said to have a closed graph if its graph is closed as a subset of X x Y.


Equivalently, using (8.20.1), it is easy to see that F has a closed graph if Xn -+ x
and F(x n ) -+ y implies that y = F(x).
It is obvious that every continuous mapping has a closed graph. The converse is
true when both X and Yare complete, but not in general.
8.22 Exrunple (A discontinuous linear map with a closed graph). Let
X denote the inner product space

X = {x E odO, 1]1 x is a polynomial and x(O) = O}.


That is, X is the space of all polynomials on [0, 1] without constant term. Define
F:X-+02 [0,1] by
F(x) = x', x E X,
where the prime denotes differentiation. Clearly, F is linear. We will show that F
has a closed graph but is not continuous.
To see that F is not continuous, let xn(t) = J2n + 1 t n (n = 1,2, ... ). Then
Ilx n I = 1 and IIF(x n ) II > n. Thus F is unbounded, hence discontinuous by Theorem
5.1l.
To see that F has a closed graph, let Xn -+ x and F(x n ) -+ y. We must show that
y = F(x). For any t E [0,1], the fundamental theorem of calculus and Schwarz's
inequality imply

It follows that with h(t) = J; y(s)ds, we have h' = y and

Since x~ = F(x n ) -+ y, it follows that Xn -+ h. But Xn -+ x. Hence x = h, and so

F(x) = x' = h' = y.

However, if both X and Yare complete, then a linear map with a closed graph
must be continuous. This is the content of our next theorem.
8.23 Closed Graph Theorem. Let X and Y be Hilbert spaces and let A :
X -+ Y be a linear map that has a closed graph. Then A is continuous.
Proof. First note that Q(A), the graph of A, is a closed linear subspace of
X x Y, hence must be complete, since X x Y is. Define Q1 and Q2 on Q(A) by

and
Q2((x,Ax» = Ax, (x, Ax) E Q(A).
ADJOINT OF A LINEAR OPERATOR 171

Then Ql and Q2 are linear on 9(A),

IIQl((X, Ax))11 = IIxll ~ II (x, Ax) II,

and
IIQ2((x,Ax))1I = IIAxll ~ I (x, Ax)lI·
This implies that both Ql and Q2 are bounded and IIQili ~ 1 (i = 1,2). That is,
Ql E B(9(A),X) and Q2 E B(9(A), Y).
Note that Ql is bijective onto X. By the bounded inverse theorem (Theorem
8.19), Ql has a bounded inverse Q11 E B(X, 9(A)). Clearly,

for each x E X. By Lemma 8.1, A = Q2 0 Q 11 is continuous. •


The following application of the closed graph theorem will be used in the next
two chapters.
8.24 Proposition. Let U and V be closed subspaces in the Hilbert space X
such that U + V is closed and Un V = {OJ. Then the mapping Q: U + V -+ U
defined by
Q(U + v) = u, for every U E U and v E V,
is a continuous linear mapping.
Proof. Let Xo = U + V. Since un V = {OJ, this is a direct sum: Xo = U EB V.
Moreover, since Xo and U are closed subspaces in X, they are also Hilbert spaces.
It is easy to see that Q : Xo -+ U is well-defined, linear, Q(u) = u for all u E U,
and Q(v) = 0 for all v E V.
Next we verify that Q has a closed graph. To see this, let Un + Vn -+ x and
Q(un + vn) -+ y. We must show that y = Q(x). Since U + V is closed, x = u + v
for some u E U and v E V. Also, Un = Q(u n + v n) -+ Y E U since U is closed.
Hence Vn = (un + v n) - Un -+ X - Y = u + v - y. Since Vn E V for all n and V is
closed, u + v - Y E V. Thus

Q(x) = Q(u + v) = Q(u + v - y) + Q(y) = Q(y) = y.


This proves that Q has a closed graph and hence, by the closed graph theorem
(Theorem 8.23), Q is continuous. •

Adjoint of a Linear Operator


Now we turn to the study of the "adjoint" of a bounded linear operator.
8.25 The Adjoint Operator. Let X and Y be inner product spaces and
A E B(X,Y).
(1) If there exists a mapping B : Y -+ X that satisfies

(8.25.1) (Ax, y) = (x, B(y))

for all x E X and y E Y, then there is only one such mapping. It is called the
adjoint of A and is denoted by A * .
172 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

(2) If X is complete, then A* always exists.


(3) If A* exists, then A* is a bounded linear operator from Y into X, i.e.,
A* E B(Y, X),

(8.25.2) IIA*II = IIAII,

and

(8.25.3) IIAI12 = IIA*AII = IIAA*II·

In particular, if X is a Hilbert space, the adjoint of A is a bounded linear mapping


A* : Y -+ X that is denned by

(8.25.4) (Ax, y) = (x, A*y) for all x E X, Y E Y.

Proof. (1) Suppose C : Y -+ X satisfies (8.25.1) as B does. Then

(x, C(y)) = (x, B(y))

for all x E X and y E Y. Hence, for each y E Y,

(x, C(y) - B(y) = 0

for all x E X. It follows that C(y) - B(y) = 0, or C(y) = B(y). Since y was
arbitrary, C = B.
(2) Now suppose X is complete. Fix any y E Y. Define fy on X by

(8.25.5) fy(x) = (Ax, y), x E X.

It is obvious that fy is linear, since A is. Moreover,

lfy(x)1 ~ IIAxllllyl1 ~ IIAllllxllllyll·


Hence fy is bounded and Ilfy I ~ IIAllllyll· That is, fy E X*. Since X is complete,
Theorem 6.10 implies that fy has a unique representer Zy EX. That is,

(8.25.6) fy(x) = (x, Zy), x E X.

Define B : Y -+ X by B(y) = Zy. Then (8.25.5) and (8.25.6) imply

(8.25.7) (Ax, y) = (x, B(y).

By part (1), B = A*.


(3) For all Yl, Y2 E Y and (Xl, (X2 E lR, (8.25.1) implies

(x, A* (XlYl + (X2Y2) = (Ax, (XlYl + (X2Y2) = (Xl (Ax, Yl) + (X2(Ax, Y2)
= (Xl (x, A*(Yl)) + (X2(X, A*(Y2)) = (x, (XIA*(Yl) + (X2A*(Y2))
for all x E X. Thus
ADJOINT OF A LINEAR OPERATOR 173

and hence A * is linear.


Furthermore, for all y E Y, we have that x = A*y E X, so that

IIA*YIl2 =(A*y,A*y) = (x,A*y) = (Ax,y)


:sIIAxlillyll :S IIAllllxllllYIl = IIAIIIIA*YIlIlYII·
Cancelling IIA*YII, we deduce that
IIA*YII :S IIAlillyll, y E Y.
Thus A* is bounded and IIA*II :S IIAII. Then, for all x E X,

IIAxll2 = (Ax, Ax) = (x,A*(Ax)):S IIxIlIlA*(Ax)lI:s IIxIlIlA*IIIIAxll·

This implies IIAxll :S IIA*lIlIxll for all x, so that IIAII :S IIA*II· Hence IIA*II = IIAII·
Using Lemma 8.1, we obtain

IIA* All :S IIA*IIIIAIl = IIAII2.


Conversely, for any x E X with IIxll = 1,

IIAxll 2= (Ax, Ax) = (A* Ax,x) :S IIA* Axil :S IIA* All·


Hence
IIAII2 = sup IIAxIl 2 :s IIA* All :S IIAII2,
IIxll=l
and this proves the first equality of (8.25.3). The second equality is proved similarly.
The last statement of the theorem follows from (1) and (2). •
A mapping A E B(X, X) is called self-adjoint if A* exists and A* = A.
8.26 Examples. (1) (The zero mapping) Let 0 : X -+ Y denote the zero
mapping: O(x) = 0 for all x. Clearly, 0* is the zero mapping from Y to X. If
X = Y, then 0 is self-adjoint.
(2) (The identity mapping) If I is the identity map on X, then for all x, y EX,

(Ix, y) = (x, y) = (x,Iy).

Thus 1* = I and I is self-adjoint.


(3) (The metric projection) Let M be a Chebyshev subspace of X. Then by
Theorem 5.13, (PM(X), y) = (x, PM(y)) for all x, y E X. Hence PM = PM and PM
is self-adjoint.
(4) (Adjoint of a matrix) Let A E B(b(n), l2(m)) have the matrix

au
[
A= ~::
am 1

relative to a given orthonormal basis {Xl, X2,.··, x n } in b(n) and {Y1, Y2,· .. , Ym}
in l2(m). Thus

(8.26.1) aij = (AXj,Yi) (i = 1,2, ... ,m; j = 1,2, ... ,n).


174 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

Since l2(n) is finite-dimensional, it is complete (Theorem 3.7), so that A* exists by


Lemma 8.25. Also, using (8.26.1), the matrix of A* : l2(m) -+ l2(n) is given by
(8.26.2)

::',::~::~ O~:jF:' ~:P'~~':f :h:' [! ';' ~r':,:~:: IT rItmn,po"


It follows that if l2(n) = l2(m) (Le., if n = m), then A is self-adjoint if and only
if the matrix of A is syrrunetric, Le., the matrix of A is equal to its transpose.
(5) (Integral operator) Consider the operator defined in Proposition 5.17.
Thus A : C2 [a, b] -+ C 2 [a, b] is given by

(8.26.3) (Ax)(s) = lb k(s, t)x(t) dt, s E [a, b],

where k is a continuous function on [a, b] x [a, b]. We saw there that A is in


B(C2 [a, b], C2 [a, b]). Now by interchanging the order of integration, we get

(Ax, y) = lb [l b k(s, t)x(t) dt] y(s) ds = lb [l b


x(t) k(s, t)y(s) dS] dt

= (x, B(y)),
where

(8.26.4) B(y)(t):= lb k(s, t)y(s) ds, t E [a, b].

It follows from Theorem 8.25 that A* exists and A* = B. Comparing (8.26.3)


and (8.26.4), we see that A is self-adjoint if (and actually only if) k is symmetric;
that is,
(8.26.5) k(s, t) = k(t, s) for all s, t E [a, b].
Next we show that without the completeness hypothesis, the adjoint of an oper-
ator may not exist.
8.27 Theorem. Let X and Y be inner product spaces with Y =I {O}. Then A*
exists for each A E B(X, Y) if and only if X is complete.
Proof. The sufficiency was already established in Theorem 8.25.
For the necessity, suppose A* exists for every A E B(X, Y). If X were not
complete, then by the remark following Theorem 6.10 (see also Exercise 6 at the
end of Chapter 6), there would exist x* E X* that has no representer. Fixing any
Yl E Y\ {O}, define A on X by
Ax = X*(X)Yl, X E X.
Then A E B(X, Y) and for each x E X and Y E Y,
(x, A*y) = (Ax, y) = (X*(X)Yl, y) = x*(X)(Yl, y).
Putting Y = Yl yields (X,A*Yl) = x*(x)IIYIil 2 , or
x*(x) = (x,A*(IIYlll- 2 Yl)), x E X.
Thus x* has the representer A*(IIYlll- 2 Yl), which is a contradiction. •
ADJOINT OF A LINEAR OPERATOR 175

Remark. During the course of the proof we essentially proved that if A = x* E


X* := B(X,lR), then A* exists if and only if A has a representer. Moreover, if A
has the representer Xl, then A* : lR -+ X is given by

A*Y=YXI, yER

More generally, we can determine precisely which bounded linear operators from
X into a finite-dimensional inner product space have adjoints, and we leave the
proof as an exercise (see Exercise 15 at the end of the chapter).
8.28 Representation Theorem for Linear Operators.
Let X be an inner product space, Y a finite-dimensional inner product space
with an orthonormal basis {YI, Y2, ... , Yn}, and A E B(X, Y). Then:
(1) There exist xi E X* (i = 1,2, ... ,n) such that
n
(8.28.1) Ax = I: xi (X)Yi for all X E X.
I

(2) A* exists if and only if each xi has a representer in X.


(3) If xi has the representer Xi EX (i = 1,2, ... , n), then A* E B(Y, X) is given
by
n
(8.28.2) A*y = I: (y, Yi)Xi for all Y E Y.

In particular, if X is a Hilbert space, then there exist Xl, X2, •.. ,Xn in X such that
n
(8.28.3) Ax = I:(x, Xi)Yi for all x E X,

and (8.28.2) holds.


The adjoint has other useful properties that are listed in the next few results.
8.29 Lemma. Let X and Y be inner product spaces, A, B E B(X, Y), and
a: E R If A* and B* exist, so do (A + B)*, (a:A) * , and A** := (A*)*. Moreover,
(1) (A + B)* = A* + B*,
(2) (a:A)* = a:A*,
(3) A** = .A.
This result is an easy consequence of the definitions involved and is left as an
exercise.
8.30 Lemma. Let X, Y, and Z be inner product spaces, A E B(Y, Z), and
BE B(X, Y). If A* and B* exist, so does (AB)*, and

(8.30.1) (AB)* = B* A*.

Proof. We already saw in Lemma 8.1 that AB E B(X, Z). Hence

(ABx,z) = (A(Bx),z) = (Bx,A*z) = (x,B*(A*z) = (x,B*A*z)

for all x EX, z E Z. Thus (AB)* = B* A*. •


176 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

8.31 Theorem. Let X and Y be Hilbert spaces and A E H(X, Y). Then A has
a bounded inverse if and only if A* does. In this case,

(8.31.1)

Proof. If A has a bounded inverse, then, using Lemma 8.30,

Similarly, 1= A*(A-I)*. It follows from Lemma 8.9 that (A*)-I = (A-I)*. Thus
A* has a bounded inverse given by (8.31.1).
Conversely, suppose A* has a bounded inverse. By the first part of the proof,
A** has a bounded inverse. But by Lemma 8.29, A** = A. •
If A E H(X,X), it is not true in general that IIAnl1 = IIAlln for each n E N.
But this will be the case when A is self-adjoint or, more generally, whenever A is
normal, i.e., whenever A commutes with its adjoint: AA* = A* A.
8.32 Lemma. Let A E H(X, X) and suppose A* exists (e.g., if X is complete).
Then:
(1) (An)* exists for each n E Nand

(8.32.1)

In particular, if A is self-adjoint, so is An.


(2) If A is normal, then

(8.32.2) IIAnl1 = IIAlln for each n E N.

In particular, (8.32.2) holds if A is self-adjoint.


Proof. (1) When n = 1, the result is trivially true. Assume that the result
holds for some n ~ 1. Then for all x, y EX,

(An+Ix, y) = (A(Anx), y) = (Anx, A*y) = (x, (An)* A*y)


= (x, (A*)n A*y) = (x, (A*)n+I y).

This shows that (An+l)* exists and is equal to (A*)n+l. Thus, by induction, the
result is valid for all n.
(2) Assume first that A is self-adjoint. Then by Theorem 8.25, IIA211 = IIAI12.
Since A2 is self-adjoint by part (1), it follows that IIA411 = IIA2112 = IIAI14. Contin-
uing inductively in this fashion, we deduce that

(8.32.3)

whenever n = 2m, mEN.


Now suppose that A is normal, but not necessarily self-adjoint. A simple in-
duction shows that A (respectively A*) commutes with any integer power of A*
(respectively A). It follows that any power of A (respectively A*) commutes with
any power of A* (respectively A). Then, using (1), we obtain
GENERALIZED SOLUTIONS TO OPERATOR EQUATIONS 177

That is, An is normal. By Theorem 8.25,

(8.32.4)

for any n E N. But AA* is self-adjoint, so that by (8.32.3) (applied to AA* instead
of A), we deduce

(8.32.5)

whenever n = 2m, mEN. Combining (8.32.4) and (8.32.5), we see that

(8.32.6)

whenever n = 2m , mEN. For a general n, choose mEN such that r:= 2m-n 2': o.
Since (8.32.6) is valid with n replaced by n + r = 2m , we have, using Lemma 8.1,
that

Hence IIAlln ::; IIAnll. Since the reverse inequality is always valid by Lemma 8.1, it
follows that IIAnl! = IIAlln. •
There are interesting duality relationships that hold between the null spaces and
ranges of A and A * .
8.33 Lemma. Let X and Y be inner product spaces, A E B(X, Y), and suppose
A* exists (e.g., if X is complete). Then:
(1) N(A) = R(A*)..l and N(A*) = R(A)..l.
(2) If X (respectively Y) is complete, then N(A)..l R(A*) (respectively
N(A*)..l = R(A)).
(3) N(A* A) = N(A).
Proof. (1) x E N(A) {=} Ax = 0 {=} (Ax, y) = 0 for all y E Y {=} (x, A*y) = 0
for all y E Y {=} x E R(A*)..l. Thus N(A) = R(A*)..l.
Since A* exists, Lemma 8.29 implies A** = A. Substituting A* for A in the first
part yields N(A*) = R(A)..l.
(2) Assume that X is complete. By (1), N(A) = R(A*)..l. Since R(A*) is a
closed subspace of X, it is Chebyshev by Theorem 3.5 and

N(A)..l = R(A*)..l..l = R(A*)..l..l = R(A*)


using Theorem 4.5. Similarly, N(A*)..l = R(A) if Y is complete.
(3) Clearly, N(A) c N(A* A). If x E N(A* A), then A* Ax = 0 implies

0= (A* Ax,x) = (Ax, Ax) = IIAxI1 2 .


Thus Ax = 0 or x E N(A). •

Generalized Solutions to Operator Equations


Now we can characterize the generalized solutions to the equation

(8.33.1) Ax=y.
178 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

8.34 Characterization of Generalized Solutions. Let X and Y be inner


product spaces, A E 8(X, Y), and suppose A* exists (e.g., if X is complete). For
given elements Xo E X and y E Y, the following statements are equivalent:
(1) Xo is a generalized solution to Ax = y; that is, IIAxo - yll :::: IIAx - yll for all
XEX;
(2) Axo = PR(A)(Y);
(3) A* Axo = A*y.
Proof. (1) and (2) are obviously equivalent.
(2) {=> (3). Using the characterization theorem (Theorem 4.9) with M = R(A),
we see that Axo = PR(A)(Y) {=> y - Axo E R(A)l. {=> Y - Axo E N(A*) (using
Lemma 8.33) {=> A*(y - Axo) = 0 {=> (3) holds. •
Note that this theorem characterizes the generalized solutions of (8.33.1) as the
ordinary (or exact) solutions to the linear equation

(8.34.1) A* Ax = A*y.

One consequence of this equivalence is that there are available methods for solving
equations such as (8.34.1). In particular, if the range of either A or A* is finite-
dimensional, then (8.34.1) reduces to solving a finite linear system of equations,
where the number of unknowns is equal to the number of equations.
However, this does not answer the question of whether generalized solutions even
exist, and if they do, whether they are unique.
For each y E Y, let G(y) denote the set of all generalized solutions to (8.33.1).
By Theorem 8.34,
G(y) = {x E X I A* Ax = A*y}.
Using Theorem 8.34 again, we see that generalized solutions exist, i.e., G(y) =1= 0,
if the range of A is Chebyshev in Y. This will be the case, for example, when
R(A) is closed and Y is complete. But even when R(A) is Chebyshev, G(y) may
contain more than one point. For example, if Xo E G(y), then Axo = PR(A) (y).
But A(xo + x) = PR(A)(Y) is also valid for every x E N(A). That is, Xo + x E G(y)
for each x E N(A). Hence if N(A) =1= {O}, that is, if A is not injective, then G(y)
contains infinitely many points whenever it contains one. In fact, G(y) contains at
most one point if and only if N(A) = {O}, i.e., A is injective.
We summarize these remarks in the following theorem.
8.35 Set of Generalized Solutions. Let X and Y be inner product spaces,
A E 8(X, Y), and suppose A* exists (e.g., if X is complete). For each y E Y, let
G(y) denote the set of all generalized solutions to

(8.35.1) Ax=y.

Then:
(1) G(y) = {x E X I A* Ax = A*y}.
(2) G(y) = N(A) + x(y) for any x(y) E G(y). In particular, G(y) is a closed
affine set in X.
(3) G(y) =1= 0 for all y E Y if and only ifR(A) is a Chebyshev subspace in Y.
This is the case, for example, when Y is complete and R(A) is closed.
(4) The following statements are equivalent:
GENERALIZED INVERSE 179

(a) G(y) contains at most one point for every y E Y;


(b) G(y) is a singleton for some y E Y;
(c) A is injective.
(5) G(y) is a single point for all y E Y if and only ifR(A) is Chebyshev and A
is injective.
(6) If X and Yare complete and R(A) is closed, then G(y) is a (nonempty)
Chebyshev set for each y E Y.
Proof. Everything has essentially already been verified with the exception of
(2) and (6). From (1), we deduce that if x(y) E G(y), then

G(y) = N(A* A) + x(y).


But N(A* A) = N(A) from Lemma 8.33, and this verifies (2).
To verify (6), note that G(y) is a closed convex set in X by (2), which by (3) is
nonempty for each y E Y if Y is complete and R(A) is closed. Since X is complete,
G(y) is Chebyshev by Theorem 3.5. •

Generalized Inverse
In practical applications it is often necessary to compute some generalized solu-
tion. One natural candidate that has been the object of wide study is the minimal
norm generalized solution.
8.36 Definition. Let X and Y be Hilbert spaces, A E B(X, Y) and suppose
R(A) is closed. Then (by Theorem 8.35), for each y E Y, the set G(y) of all
generalized solutions to

(8.36.1) Ax=y

has a unique element A-(y) := PC(y) (0) of minimal norm. This element is called the
minimal norm generalized solution to (8.36.1), and the mapping A- : Y --+ X
thus defined is called the generalized inverse of A.
We will see later that the generalized inverse is a bounded linear mapping. In
certain simple cases, the generalized inverse is related to ordinary inverses as follows.
8.37 Lemma. (1) If A-l exists, then A- = A-l.
(2) If (A* A)-l exists, then A- = (A* A)-l A*.
We leave the simple proof as an exercise.
The following characterization of the minimal norm generalized solution is funda-
mental for its practical computation, for the computation of the generalized inverse,
and for showing that the generalized inverse is a bounded linear mapping.
8.38 Characterizing the Minimal Norm Generalized Solution. Let X
and Y be Hilbert spaces and suppose A E B(X, Y) has closed range. Then, for
everyy E Y,

A-(y) = G(y) nN(Al = {x E N(A).l. I A* Ax = A*y}


= {x E N(A).l. I Ax = PR(A)Y}.
180 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

Proof. From Theorem 8.35, we have that G(y) = N(A) + x(y) for any x(y) E
G(y). Since A - (y) = PC(y) (0), it follows from Exercise 2 at the end of Chapter 4
that A-(y) E N(A)..L. Thus x = A-(y) if and only if x E G(y) and x E N(A).L. But
x E G(y) if and only if A* Ax = A*y (by Theorem 8.35) if and only if Ax = PR(A)Y
(by Theorem 8.34). •
Just as in the above theorem, it will be seen that a critical assumption for
several important facts concerning the generalized inverse is that the range of a
certain operator be closed. The next result shows that this is equivalent to the
range of its adjoint being closed.
8.39 Lemma. Let X and Y be Hilbert spaces and A E B(X, Y). Then R(A)
is closed if and only ifR(A*) is closed.
Proof. If R(A) is closed, the closed range theorem (Theorem 8.18) implies that
there exists 8 > 0 such that
(8.39.1) IIAxll:::: 811xll for all x E N(A).L.
Since N(A)..L is a closed subspace in X, it is Chebyshev by Theorem 3.5, and hence
IIAxl1 = IIA(PN(A).LX + PN(A)X) I = IIAPN(A).Lxll :::: 8IfPN(A).Lxll
for all x E X by (8.39.1).
---- ---- ----..L..L
If y E R(A*), then Theorem 4.5 and Lemma 8.33 imply R(A*) = R(A*) =
R(A*)..L.L = N(A)..L, so that y E N(A)..L. Define f on R(A) by
(8.39.2) f(Ax) := (x, y) for all x E X.
Then f is well-defined, since ifAxl = A.T2, then Xl - X2 E N(A), and hence
0= (Xl - X2, y) = (Xl, y) - (X2' y).
Moreover, f is linear and bounded; the latter fact follows because
If(Ax)1 = l(x,y)1 = I\PN(A).LX + PN(A)X,y)1 = I(PN(A).LX,y)1
1
::; IfPN(A).L(x)llllyll::; JllAxllllYl1
implies Ilfll ::; Ily11/8.
By the Frechet-Riesz representation theorem (Theorem 6.10), f has a representer
Axo E R(A). That is,
(8.39.3) f(Ax) = (Ax,Axo) for all x EX.
From (8.39.2)-(8.39.3), we obtain
(x, y) = (Ax, Axo) = (x, A* Axo)
for all x E X. This implies that y = A* Axo E R(A*). Thus R(A*) C R(A*) and
R(A*) is closed.
Conversely, if R(A*) is closed, the above argument shows that R(A**) is closed.
But A** = A by Lemma 8.29. •
For the next several results we will be working under the same hypotheses.
Namely, the hypothesis
(8.39.4) X and Y are Hilbert spaces and A E B(X, Y) has closed range.
GENERALIZED INVERSE 181

8.40 Lemma. If hypothesis (8.39.4) holds, then

(8.40.1) R(A- A) = R(A-) = R(A* A) = R(A*) = N(A)1-.


In particular, the ranges of A -, A-A, and A * A are all closed.
Proof. By Lemma 8.39, R(A*) is closed. By Lemma 8.33,

N(A)1- = R(A*) = R(A*),

which proves the last equality of (8.40.1).


Clearly, R(A* A) c R(A*). Conversely, if x E R(A*), then x = A*y for some
Y E Y. Thus, using Lemma 8.33,

x = A*(PJV(A*)Y + PJV(A*)J.Y) = A* PJV(A*)J.Y = A* PR(A)Y E R(A* A).

This proves R(A*) = R(A* A).


For each Y E Y, using Theorem 8.38 and Lemmas 8.33 and 8.39, A-y E N(A)1- =
R(A*). Thus R(A-) c R(A*). For the reverse inclusion, let x E R(A*) = N(A)1-
and set Y = Ax. Then A* Ax = A*y. By Theorem 8.38, x = A-y E R(A-). Thus
R(A*) = R(A-).
Finally, observe that R(A- A) c R(A-) is obvious. If x E R(A-), then x = A-y
for some y E Y. But by Theorem 8.38, A-y = A- PR(A)Y E R(A-A). Thus
R(A- A) = R(A-). •

Now we are in a position to prove that the generalized inverse is a bounded linear
operator.
8.41 Generalized Inverse is Linear and Bounded. If A E B(X, Y) and
hypothesis (8.39.4) holds, then A- E B(Y, X).
Proof. Let Yi E Y, D:i E lR (i = 1,2). By Theorem 8.38, A-(Yi) E N(A)1-nG(Yi)
and Xo := D:1A-(Yl) + D:2A-(Y2) E N(A)1-, since N(A)-L is a subspace. Further,
since A - (Yi) E G(Yi), we have

A*Axo = A*A[D:1A-(Yl) +D:2A-(Y2)] = D:1A*A[A-(Yl)] +D:2 A *A[A-(Y2)]


= D:IA*Yl + D:2A*Y2 = A*(alYl + a2Y2).

and hence A-is linear.


By Lemma 8.40, R(A-) = R(A*) = N(A)1-. It follows from the closed range
theorem (Theorem 8.18) that IIA(A-y)11 ;::: pllA-yll for all Y E Y, and some p > O.
Since AA-y = PR(A)Y by Theorem 8.38, we obtain

Hence A- is bounded and IIA-II <::: lip. •

There are alternative ways of describing the generalized inverse. An explicit


formula, which is often used later, is the following.
182 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

8.42 Theorem. If hypothesis (8.39.4) holds, then

(8.42.1 )

Proof. We first show that the expression on the right side of (8.42.1) is well-
defined, i.e., that the restriction of A to N(A).l has a bounded inverse on R(A).
Let Ao := AIN(A).L. Then Ao is a bounded linear operator from the Hilbert space
N(A).l (in X) to the Hilbert space R(A) (in Y). Moreover, Ao is bijective. To see
that A is injective, suppose AOX1 = AOX2 for some Xi E N(A).l. Then Xl - X2 E
N(A).l and A(X1 - X2) = AO(X1 - X2) = 0, so Xl - X2 E N(A).l nN(A) = {OJ, or
Xl = X2· To see that Ao is surjective, let y E R(A). Then there exists X E X such
that y = Ax. Then

Thus Ao is surjective. By the bounded inverse theorem (Theorem 8.19), we see


that Ao has a bounded inverse, i.e., A01 E B(R(A),N(A).l). Let

B := A01 PR(A).

We will show that B = A-. To this end, first note that B E B(Y,N(A).l) C
B(Y, X). Fix any y E Y. Since the range of B is contained in N(A).l, we have
By E N(A).l. Also,

ABy = AoBy = AoAo1 PR(A)Y = PR(A)Y.

By Theorem 8.38, By = A-y. Since y E Y was arbitrary, B = A-. •

Less explicit descriptions of the generalized inverse of an operator, which are


nevertheless useful in the applications, are collected in the next theorem.
8.43 Characterization of Generalized Inverse. Let hypothesis (8.39.4) hold
and let B E B(Y,X). Then the following four statements are equivalent:
(l)B=A-;
(2) (i) BAx = X for all X E N(A).l,
(ii) By = 0 for all y E R(A).l;
(3) (i) AB = PR(A),
(ii) BA = PR(B);
(4) (i) (AB)* = AB,
(ii) (BA)* = BA,
(iii) ABA = A,
(iv) BAB = B.
Proof. (1) =? (2). Assume (1) holds. Using Theorem 8.42, we can write B =
(AIN(AV)-l PR(A). Thus for each x E N(A).l, we have

BAx = (AIN(A).L )-1 PR(A)Ax = (AIN(A).L )-1 Ax = x.

If y E R(A).l, then PR(A)Y = 0, and so By = O. Thus (2) holds.


GENERALIZED INVERSE 183

(2) =} (3). Assume (2) holds. Then, for each y E Y, we have


ABy = AB(PR(A)Y + PR(A)"-Y) = ABPR(A)Y'
Choose x E X such that Ax = PR(A)Y' Since x = PN(A)X + PN(A)"-X, we see that
we can write PR(A)Y = AXI, where Xl = PN(A)"-X E N(A)l.. Hence
ABy = ABPR(A)Y = ABAxI = AXI = PR(A)Y,
where (2)(i) was used for the third equality. Since Y was arbitrary, it follows that
AB = PR(A); that is, (3)(i) holds.
Similarly, for each x EX,
BAx = BA(PN(A)X + PN(A)"-X) = BAPN(A)"-X = PN(A)"-X = PR(A*)X,
where we used Lemmas 8.33 and 8.39 for the last equality. Hence BA = PR(A*)'
It follows that R(A*) C R(B). On the other hand, if x E R(B), then x = By for
some Y E Y and
x = B(PR(A)Y + PR(A)"-Y) = BPR(A)Y'
We then have PR(A)Y = Az for some z E X. Thus,
x = BAz = BA(PN(A)Z + PN(A)"-Z) = BAPN(A)"-Z
= PN(A)"-Z E N(A)l. = R(A*)
using Lemmas 8.33 and 8.39. This proves R(B) C R(A*) and hence R(B) = R(A).
If follows that BA = PR(B)'
(3) =} (4). Assume (3) holds. Then since projections are self-adjoint by Example
8.26(3), it follows that AB and BA are self-adjoint. Also, ABA = PR(A)A = A
and BAB = PR(B)B = B. Thus (4) holds.
(4) =} (1). Assume (4) holds. Then AB is self-adjoint and (AB? = (ABA)B =
AB, so AB is idempotent. It follows (see Exercise 13 at the end of the chapter)
that AB is the orthogonal projection onto its range:
(8.43.1) AB = PR(AB)'
We claim that R(AB) = R(A). To see this, first note that R(AB) C R(A) is
obvious. Conversely, since A = ABA, we see that R(A) = R(ABA) c R(AB),
which proves the claim.
If follows from this claim and (8.43.1) that AB = PR(A)' By symmetry, BA =
PR(B)' Next we observe that R(B) = R(A*). To see this, note that B = BAB =
(BA)* B = A* B* B, and so R(B) C R(A*). On the other hand, A = ABA implies
that A* = (BA)* A* = BAA* and hence R(A*) C R(B). This proves R(B) =
R(A*). Combining these results we obtain
AB = PR(A) and BA = PR(A*) = PN(A)"--
If Y E R(A), then Y = Ax for some x E X, so Y = A(PN(A)X + PN(A)_LX)
APN(A)"-X, and
A-y = (AIN(A)"-)-l PR(A)Y = (AIN(A)"-)-l APN(A)"-X = PN(A)"-X = BAx = By.
On the other hand, if Y E R(A)l., then PR(A)Y = 0 and
A-y = (AIN(A)"-)-l PR(A)Y = 0= BPR(A)Y = BABy = By.
This shows that A- and B are equal on R(A) UR(A)l.. Since Y = R(A) EEJR(A)l.,
it follows that B = A-, and thus (1) holds. •
It is now fairly simple to compute some examples of generalized inverses based
on Theorem 8.43.
184 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

8.44 Examples.
(1) The generalized inverse of 0 E B(X, Y) is 0 E B(Y, X).
(2) If M is a closed subspace of the Hilbert space X, then PM = PM. In
other words, the generalized inverse of an othogonal projection is the orthogonal
projection itself.
(3) If A E B(£2(n),£2(n)) is the "diagonal" operator defined by

for some scalars (Xi E JR, then A- E B(£2(n),£2(n)) is given by

In other words, if the matrix for A is the n x n diagonal matrix

then the matrix for A-is the n x n diagonal matrix

where
if (Xi of 0,
if(Xi = O.
(4) If A is the linear functional defined on £2 (n) by
n
Ax = (x,a) = Lx(i)a(i), x E £2 (n),

for some a E £2(n) \ {O}, then A- E B(JR'£2(n)) is given by

A- 1 A* 1 JR
y = IIal1 2 y = IIal1 2 ay, y E .

We note that using Theorem 8.43, (1) is clear and (2) follows, since R(PM ) = M
and PMPM = pir PM· Similarly, examples (3) and (4) can be easily verified
using Theorem 8.43.
For the actual computation of the generalized inverse, the next result is often
useful, since when either X or Y is finite-dimensional, it reduces the computation
of the generalized inverse of A to that of the generalized inverse of a matrix.
GENERALIZED INVERSE 185

8.45 Theorem. If hypothesis (8.39.4) holds, then


(8.45.1) A- = (A* A)- A* = A*(AA*)-.
Proof. We first observe that by Lemma 8.40,
(8.45.2) R(A*) = R(A* A).
For each y E Y, we deduce from Theorem 8.43(1) (applied to A* A rather than
A) and (8.45.2) that
A* A(A* A)- A*y = P'R.(A*A) (A*y) = P'R.(A*)(A*y) = A*y.
From Theorem 8.34, it follows that Xo = (A* A)- A*y is a generalized solution to the
equation Ax = y. Moreover, using Lemma 8.40, the fact that A* A is self-adjoint,
and (8.45.2), we see that
Xo E R((A* A)-) = R((A* A)*) = R(A* A) = R(A*) = N(A)l..
By Theorem 8.38, Xo = A-y. Thus
(A* A)-A*y = A-y.
Since this holds for every y E Y, the first equality in (8.45.1) holds. The second
equality is proved similarly (see Exercise 23 at the end of the chapter). •
8.46 Application. Let X be a Hilbert space, {Xl,X2,'" ,xn } C X, and define
A: X -t l2(n) by
(8.46.1) Ax=(X,Xl),(X,X2), ... ,(x,xn », xEX.
By Theorem 8.28, A* : l2(n) -t X is given by
n
(8.46.2) A*y = Ly(i)Xi, Y E l2(n).
1

We want to determine the generalized inverse of A.


By Theorem 8.45, A- = A*(AA*)-. Letting B = AA*, the problem reduces to
determining B-. But B: ben) --+ l2(n) is given by

(8.46.3) n
= Ly(i)( (Xi, Xl), (Xi, X2),' .. , (Xi, Xn».
i=l

Then the matrix of B relative to the canonical orthonormal basis {ei' e2, ... , en}
in l2(n), where ei(j) = 8ij , is given by

bij =(Bej, ei) = /


\k=l
t ej(k)«xk, Xl)' (Xk' X2), .. . ,(Xk, x n»), ei)

n
=L ej(k)«( (Xk, Xl)' (Xk, X2), ... , (Xk, x n », ei)
k=l
n
= Lej(k)(Xk,Xi) = (Xj,Xi) = (Xi,Xj).
k=l
186 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

That is, the matrix of B is the Gram matrix B = G(Xl' X2," ., xn). Thus
(8.46.4)

and the problem reduces to finding the generalized inverse of the n x n symmetric
Gram matrix B.
In particular, if {Xl, X2, ... , xn} is linearly independent, then B has an inverse
by Theorem 7.7, so that
(8.46.5)
Moreover, if {Xl, X2, ... , xn} is actually orthonormal, then B = I, so that
(8.46.6)
We conclude this chapter by showing how Theorem 8.38 can be used to compute
the generalized inverse of a matrix.
8.47 Exrunple. Let us compute the generalized inverse of the matrix

A= [-1 1]
1 -1'

(That is, A is the operator in B([2(2), [2(2)) given by

Ax = [-x(l) + x(2)]el + [x(l) - x(2)]e2 = [-x(l) + x(2)](el - e2),


where ei(j) = Dij for i,j = 1,2.)
Fix any y E [2(2). We want to determine x = A-(y) using Theorem 8.38.
According to that theorem, we seek the element x in N(A)-L that also satisfies
(8.47.1) A*Ax = A*y.
By Lemma 8.40, N(A)-L = R(A*) = R(A), since A* = A. Thus we must determine
the x E R(A) satisfying (8.47.1). But

R(A) = {o:(el - e2) I 0: E lR} = span{ el - e2}.


Thus x E R(A) B x = o:(el - e2) for some 0: E R Further, if x also satisfies
(8.47.1), then A 2x = Ay. Now,

A 2x = A(Ax) = A[-20:(el - e2)] = -20:A(el - e2) = 40:(el - e2)


and
Ay = [-y(l) + y(2)](el - e2).
Hence A2X = Ay B 40: = -y(l) +y(2), or 0: = H-y(l) +y(2)]. That is, x = A-(y)
B x = H-y(l) + Y(2)](el - e2)' This proves that the generalized inverse of A is
given by

(8.47.2)

It follows that the matrix of A-is given by

A- = [-tt _\].
EXERCISES 187

Exercises

1. Assume that X and Yare inner product spaces with X n-dimensional, and
let A: X -+ Y be linear.
(a) If A is injective, then R(A) is n-dimensional.
(b) If A has an inverse, then Y is n-dimensional.
2. Suppose A, An are in SeX, Y), x, Xn in X, An -+ A, and Xn -+ x. Show
that Anxn -+ Ax.
3. Let {eI, e2, .. . , en} be the canonical orthonormal basis in £2(n) , i.e., ei(j) =
Oij for i,j = 1,2, ... , n. Given any subset {Xl, X2, ... ,xn } ofl'2(n), consider
the operator A on £2 (n), defined by
n
Ax = 2:)x, xi)ei for all x E £2(n).
1

(a) Show that A is a bounded linear operator from £2(n) into itself.
(b) Show that A is injective if and only if {Xl, X2, ... ,xn } is linearly indepen-
dent.
(c) Show that A is surjective if and only if {Xl, X2, ... , xn} is linearly inde-
pendent. [Hint: The Gram determinant g(XI, X2, ... , xn) =1= 0 if and only if
{Xl,X2, ... ,xn } is linearly independent.]
(d) Show that A is bijective if and only if {Xl, X2, ... ,xn } is linearly indepen-
dent.
(e) Show that the matrix of A relative to the basis {eI, e2, ... , en} is given
by [aij]f,j=l' where aij = (ej,xi) = Xi(j)·
(f) If Xi = I:~=l ej (i = 1,2, ... , n), what is the matrix of A?
4. If A E B(X, Y) has a bounded inverse, show that

IIA-lil = ( inf IIAXII)-l


11",11=1
5. Suppose X is an inner product space, X E X, and {xn} in X satisfies

X*(Xn) -+ x*(x) for all x* E X*.

Show that {xn} is bounded: SUPn Ilxnll < 00, and Ilxll ::; liminfn Ilxnll.
[Hint: For n E N, define An : X* 1-7 lR by An(x*) = x*(xn), X* E X*.
By Corollary 8.15, sUPn IIAnll < 00. But Ilxnll = II An II using Theorem 6.25
with K = {O}.]
6. Let X be a Hilbert space and let {en I n = 1,2, ... } be an orthonormal
sequence in X. If I:~ (x, en) converges for every X E X, show that I:~ en
converges. [Hint: Consider the sequence of functionals {x~} defined on X
by x~(x) = I:~=l (x, ek) for every x E X. Use the uniform boundedness
theorem (Theorem 8.14).]
7. Define A on b by

Ax = (x(I), x(2) - x(I), x(3) - x(2), ... ,x(n) - x(n - 1), ... ).
Show that
(a) A E B(l2' l2) and IIAII ::; 2.
188 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

(b) A is injective.
(c) A is not surjective.
(d) Y E R(A) if and only if

~ [tY(i)r < =.
8. Let X be a Hilbert space, A E B(X, X), and IIAII < 1. Show that I - A has
a bounded inverse given by

(I - A)-I = lim(I + A + A2 + ... + An),


n

where the limit is in the norm of the space B(X, X). [Hint: Let Bn =
I + A + A2 + ... + An, n = 1,2, .... Show that limn An = 0, B := limEn
exists, and (I - A)Bn = I - An+1 = Bn(I - A).]
9. Let X be a Hilbert space, A E B(X, X), and III - All < 1. Show that A has
a bounded inverse given by

A-I = lim[J + (I - A) + (I - A)2 + ... + (I - A)n].


n

[Hint: Problem 8.]


10. Verify Proposition 8.20. That is, prove that X x Y is an inner product space;
moreover, X x Y is complete if and only if both X and Yare complete.
[Hint: {(xn,Yn)} is Cauchy in X x Y if and only if {xn} is Cauchy in X
and {Yn} is Cauchy in Y.]
11. Prove Lemma 8.29. That is, if both A, B E B(X, Y) have adjoints, then
A + B, aA, and A* also have adjoints, which are given by
(a) (A+B)*=A*+B*,
(b) (aA)* = aA*, and
(c) A** = A.
12. If A E B(X, Y) has an adjoint, then
(a) AA* and A* A are both self-adjoint,
(b) AA* is "nonnegative" on Y, i.e., (AA*y, y) ::,. 0 for all Y E Y,
(c) A* A is nonnegative on X.
13. Let X be a Hilbert space and P E B(X, X). Show that P is idempotent
and self-adjoint (i.e., p 2 = P, and P* = P) if and only if P = PM for some
closed subspace M in X.
14. Let X be a Hilbert space and M and N closed subspaces. With the help of
Exercise 13, prove:
(a) PMPN is an orthogon;:tl projection ¢} PM and PN commute: PMPN =
PN PM. In this case, PM PN = PMnN .
(b) PM + P N is an orthogonal projection ¢} PNPM = O. In this case,
PM +PN = PM +N .
(c) If PM and P N commute, then PM + PN - PM PN is an orthogonal
projection.
(d) If PM and P N commute, then PM + PN - 2PM PN is an orthogonal
projection.
15. Let A E B(X, Y) and suppose A* exists.
(a) Show that R(A) is dense in Y if and only if A* is injective. [Hint:
EXERCISES 189

Lemma 8.33.]
(b) Show that R(A*) is dense in X if and only if A is injective.
(c) If R( A) is closed, then A (respectively A *) is injective if and only if A *
(respectively A) is surjective.
16. Prove Proposition 8.28. That is, let X be an inner product space, Y an
n-dimensional inner product space with orthonormal basis {y!, Y2,···, Yn},
and A E B(X, Y). Show that
(a) There exist xi E X* (i = 1,2, ... , n) such that
n
Ax = Lx7(x)Yi for all x E X.
1

(b) A* exists if and only if each xi has a representer in X.


(c) If xi has the representer Xi E X (i = 1,2, ... ,n), then A * is given by
n
A*y = L(Y,Yilxi for all Y E Y.
i=l

17. Give an example of A E B(X, X) such that IIA211 < IIAI12. Hence show that
Lemma 8.32(2) is false if A is not normal.
18. If P and Q are metric projections onto closed subspaces of a Hilbert space,
prove that R(PQP) = R(PQ). [Hint: Lemma 8.33.]
19. Prove Lemma 8.37. That is, verify
(a) If A- l exists, then A- = A- l .
(b) If (A*A)-l exists, then A- = (A*A)-lA*.
20. If A E B(X, Y) and B E B(Y, Z) each have a bounded inverse, then BA E
B(X, Z) has a bounded inverse given by

21. Show that if A E B(X, Y) has closed range, then (A*)- = (A-)*. In other
words, the generalized inverse of the adjoint is the adjoint of the generalized
inverse.
22. Compute the generalized inverse of each of the two matrices

[1 43 5]
2 6 and
[1~ 2]
: .

[Hint: Exercise 21 will reduce the amount of work.]


23. Prove the second equality in Theorem 8.45. That is, if X and Yare Hilbert
spaces and A E B(X, Y) has closed range, then A- = A*(AA*)-. [Hint:
Inspect the proof of the first equality in (8.45.1) and use Theorem 8.43(3)
applied to AA* rather than A.]
24. Use Theorem 8.43 to verify Example 8.44(3). That is, if A E B(£2(n), £2(n))
is defined by
190 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

for some scalars Qi E lR, then A- E B( (12 (n), (12 (n)) is given by

where
if Qi =f 0,
if Qi = O.
25. Use Theorem 8.43 to verify Example 8.44(4). That is, if A is the bounded
linear functional on (12 (n) defined by
n
Ax = (x,a) = I:x(i)a(i), x E (l2(n),

for some a E (l2(n) \ {O}, then A- E B(lR, (12 (n)) is given by

26. If A E B(X, Y) and a E lR, show that (QA)- = iA- if Q =f 0, and


(QA)- = 0 if Q = O.

The next three exercises show that it is sometimes convenient to work in a


"product space" setting for certain problems that do not initially appear to
be in the form of a usual problem of best approximation in I·filbert space.

27. Let X be a Hilbert space and Xi E X for i = 1,2, ... , n. Show that x mini-
mizes 2.:~ Ilx - xil1 2 if and only if x = ~ 2.:~ Xi. In other words, the vector
that minimizes the sum of the squares of its distances from a finite number
of points is just the average of these points. [Hint: Consider the product
space xn := X x X x .. , x X (with n factors X) consisting of all n-tuples
of vectors (XI,X2,'" ,xn ), where Xi E X for all i, with the inner prod-
uct and norm defined by ((XI,X2,· .. ,Xn),(YI,Y2,.··,Yn)) := 2.:~(Xi'Yi)
and II(XI, X2,' .. , x n ) 112 = 2.:~ IIXi 112. Show that xn is a Hilbert space (see
Exercise 10). Let D := {(x,x, ... ,x) I x E X} denote the "diagonal" sub-
space in xn. Observe that (xo,xo, ... ,xo) ED is the best approximation
to (Xl,X2, ... ,Xn ) from D if and only if Xo minimizes 2.:~ Ilx - xi11 2. Fi-
nally, apply the characterization of best approximations (Theorem 4.9 in
the space xn) to deduce that (xo, Xo, ... , xo) is the best approximation to
(Xl, X2, .. . , x n ) from D if and only if Xo = ~ 2.:~ xd
28. More generally than Exercise 27, let A E B(X, Y) and Yi E Y (i = 1, ... , n).
Show that x minimizes 2.:~ IIAx - Yil1 2 if and only if x is a generalized
solution to Ax = ~ 2.:~ Yi. [Hint: Consider the map A : X --t yn defined
by A(x) := (Ax, Ax, .. . , Ax). Apply known results to A instead of A.]
29. More generally than Exercise 28, let Ai E B(X, Y) and Yi E Y (i =
1,2, ... ,n). Show that x E X minimizes 2.:~ IIAiX - Yil1 2 if and only if
x is a solution of the equation (2.:~ AT Ai)X = 2.:~ ATYi.
HISTORICAL NOTES 191

Historical Notes

Baire (1899) proved (part (1) of) Baire's theorem 8.13 in the particular case
where X = £2(n). (The case n = 1 had been proved two years earlier by Osgood
(1897).) But the result and proof immediately generalize when £2(n) is replaced
by any complete metric space. This was done explicitly by Kuratowski (1930) and
Banach (1930).
The uniform boundedness theorem (Theorem 8.14) was proved when X is a
Banach space and Y is the scalar field by Hildebrandt (1923; p. 311), but exactly
the same proof is valid when Y is any Banach space. This result is often incorrectly
attributed to Banach or Banach and Steinhaus, although certain special cases were
known earlier (see Dunford and Schwartz (1958; pp. 80-81». Banach and Steinhaus
(1927) showed that the uniform boundedness theorem is still valid when X and
Yare Banach spaces and (8.14.1) holds only for x in a second-category subset
of X. For an elegant proof of the uniform boundedness theorem, based only on
completeness and avoiding the notion of category, see Hausdorff (1932; p. 304).
Banach (1929a) and (1929b) first proved the bounded inverse theorem (Theorem
8.19) for the general case where X and Yare Banach spaces. For another proof of
Theorem 8.19, see Schauder (1930). Banach (1932; Theorem 7, p. 41) also proved
the closed graph theorem (Theorem 8.23) in the general case where both X and Y
are Banach spaces.
The notion of the adjoint of a bounded linear operator originated with matrix
theory and with the theories of differential and integral equations. In the Lp spaces
(p> 1) and £2, Riesz (1910; p. 478), (1913; p. 85) made use of this notion. Banach
(1929a; p. 235) introduced the adjoint operator in a general Banach space, and
proved Theorem 8.31 and Lemma 8.33(2) in this context. The closed range theorem
(Theorem 8.18) is due to Hestenes (1961; Theorem 3.3).
What we have called a "generalized solution" is also called a "least squares
solution" or "extremal solution" by some authors. The notion of a generalized
inverse goes back at least to Fredholm (1903), where a particular generalized inverse
(called a "pseudoinverse") of an integral operator was considered. (A nice outline of
the history behind generalized inverses can be found in the Introduction of the book
by Ben-Israel and Greville (1974).) Generalized inverses of differential operators
were implicit in the study of generalized Green's functions by Hilbert (1904). For
a good historical survey of this subject up to 1968, see Reid's book (1968). We
should mention that some authors use other notation for generalized inverses, for
example, At and A + .
In a short abstract, Moore (1920) (see also Moore (1935) for more complete
details) defined a unique generalized inverse (called a "general reciprocal") for every
rectangular matrix. Moore's definition, in the more general context of bounded
linear operators with closed range, is the following: If A E 8(X, Y) has closed
range, then A-is the unique operator in 8(Y, X) that satisfies AA - = PR(A) and
A- A = PR(A-)' That is, owing to Theorem 8.43, Moore's definition of generalized
inverse agrees with the "variational" definition (Definition 8.36) given here. We
preferred Definition 8.36 since it seemed to us more natural and geometrically
intuitive, while Moore's "definition" (as well as others mentioned below) actually
required first proving that the object even existed and was unique! Of course, with
Definition 8.36, it may not have been a priori obvious that the generalized inverse
was linear. We note that Definition 8.36 of generalized inverse was also preferred
192 GENERALIZED SOLUTIONS OF LINEAR EQUATIONS

in the book of Groetsch (1977).


A number of authors, evidently unaware of Moore's work, also considered gen-
eralized inverses in different contexts, for example, Siegel (1937) for matrices, and
Tseng (1933) (see also (1949a), (1949b), (1949c), and (1956)), Murray and von
Neumann (1936), and Atkinson (1951), (1953) for operators. Bjerhammar (1951)
defined the generalized inverse of a rectangular matrix, and Penrose (1955) refined
Bjerhammar's notion, which, in the more general context of operators, may be
stated as follows. If A E B(X, Y) has closed range, then A- is the unique operator
in B(Y,X) that satisfies the following four properties: (1) AA- = (AA-)*, (2)
A-A = (A- A)*, (3) AA- A = A, and (4) A- AA- = A-. Rado (1956) pointed
out the equivalence of the definitions of Moore and Penrose, and this is one of the
reasons why this generalized inverse is often called the "Moore-Penrose" inverse.
That is, Rado (1956) proved the equivalence of statements (2) and (3) in Theo-
rem 8.43. For an operator A E B(X, Y) that has closed range, Desoer and Whalen
(1963) defined the generalized inverse of A to be the unique operator A- E B(Y, X)
that satisfies (1) A- Ax = x for all x E N(A).l. and (2) A-y = 0 for all y E R(A).l..
By Theorem 8.43, the definitions of Moore, Penrose, and Desoer and Whalen are
all equivalent to Definition 8.36.
Anderson and Duffin (1969; Theorem 8) have proved the following interest-
ing result: If M and N are closed subspaces of a Hilbert space, then PMnN =
2PM(PM + P N )- PN .
While the Moore-Penrose inverse is generally agreed to be the generalized inverse
par excellence, not all definitions of generalized inverses that have been studied are
equivalent. For example, Rao (1962) defined a "g-inverse" of an m x n matrix A to
be an n x m matrix G that satisfies only the third property of Penrose's definition,
namely, AGA = A. (See also the book of Rao and Mitra (1971), where this is more
fully developed.)
The main reason for the "closed range" hypothesis (8.39.4) for several of the
fundamental results of this chapter is to ensure the existence of generalized solutions,
and hence the existence of generalized inverses. To see this, first note that the range
of a matrix operator, being finite-dimensional, is always closed, so the closed range
hypothesis is automatically satisfied. However, for more general operators between
Hilbert spaces, this is not always the case. If R(A) were not closed in the Hilbert
space Y, then R(A) would not be Chebyshev, and hence there would exist some
y E Y that failed to have a best approximation in R(A). It follows by Theorem
8.35(3) that the set of generalized solutions G(y) to the equation Ax = y is empty!
In particular, the generalized inverse could not be defined at y. In the case where
the range of the operator is not closed, the generalized inverse turns out to be an
unbounded linear operator (see Groetsch (1977; Chapter III)).
There have been many books and countless papers devoted to the theory and
computation of generalized inverses. For example, a small sample of books includes
Boullion and Odell (1971), Rao and Mitra (1971), Ben-Israel and Greville (1974),
Nashed (1976), and Groetsch (1977). In particular, Groetsch (1977; Chapter II)
has described various methods for actually computing the generalized inverse of
operators. Finally, we should note that a very general abstract theory of generalized
inverses for linear operators between two linear topological spaces has been studied
by Nashed and Votruba ([19741]' [1974II]).
CHAPTER 9

THE METHOD OF ALTERNATING PROJECTIONS

The Case of Two Subspaces


In this chapter we will describe a theoretically powerful method for computing
best approximations from a closed convex set K that is the intersection of a finite
number of closed convex sets, K = nr
K i . This method is an iterative algorithm
that reduces the problem to finding best approximations from the individual sets
K i . The efficacy of the method thus depends on whether the given set K can
be represented as the intersection of a finite number of sets Ki from which it is
"easy" to compute best approximations. This will be the case, for example, when
the Ki are either half-spaces, hyperplanes, finite-dimensional subspaces, or certain
cones. Several applications will be made to a variety of problems including solving
linear equations, solving linear inequalities, computing the best isotone and best
convex regression functions, and solving the general shape-preserving interpolation
problem.
The rate of convergence of the method of alternating projections will also be
studied, at least when all the K/s are subspaces or, more generally, affine sets.
This rate will be described in terms of the angles between the various subspaces
involved.
In its simplest abstract formulation (namely, the intersection of two subspaces),
the method of alternating projections originated with von Neumann in 1933. Be-
cause the theory is complete, simpler, and more elegant in the two-subspace case,
we shall study this case first in some detail to motivate the more general results to
come later.
We begin with two simple observations. The first characterizes the bounded lin-
ear operators that are orthogonal projections, and the second characterizes exactly
when the product of two orthogonal projections is again an orthogonal projection.
9.1 Lemma. Let X be an inner product space and A E B(X, X). Then A is
idempotent and self-adjoint (i.e., A2 = A and A* = A) if and only if A = PM for
some Chebyshev subspace M. (In fact, M = R(A), the range of A.)
Proof. If A = PM, then by Theorem 5.13, A = A* and A2 = A.
Conversely, suppose A = A* and A2 = A. Let M = R(A). If y E M, y = Ax for
some x E X, which implies y = Ax = A 2x = A(Ax) = Ay. That is, Ay = y for all
y EM. For any x E X and y EM,
(x - Ax, y) = (x, y) - (Ax, y) = (x, y) - (x, Ay) (since A* = A)
= (x,y) - (x,y) = o.
Thus x - Ax E M.L. By Theorem 4.9, Ax = PMx. Since x was arbitrary, A = PM .
• Next we characterize when orthogonal projections commute.

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
194 THE METHOD OF ALTERNATING PROJECTIONS

9.2 Lemma. Let M and N be Chebyshev subspaces of the inner product space
X. Then the following statements are equivalent:
(1) PM and PN "commute"; i.e., PMPN = PNPM ;
(2) PN(M) eM;
(3) PM(N) eN;
(4) PMPN = PMnN ;
(5) PM P N is the orthogonal projection onto a Chebyshev subspace.
In particular, if MeN or N c M, then PM and P N commute.
Proof. (1) ==? (2). Suppose PM and P N commute. Then for any x E M, we
have that

Thus (2) holds.


(2) ==? (3). Suppose (2) holds, x E N, and Y E N-L. Then

(PM(X),y) = (PMPN(X),y) = (PN(X),PM(y)


= (X,PNPM(y) = (X,PMPNPM(y) (since PNPM(y) EM by (2))
= (PM (X), PNPM(y) = (PNPM(X), PM(Y)
= (PMPNPM(X),y) = (PNPM(X),y) by (2)
= ° (since y E N-L).

This proves that PM(X) E Nl.-L. But Nl.-L = N by Theorem 4.5(8). Hence
PM(X) EN for every x E Nand (3) holds.
(3) ==? (4). Suppose PM(N) c N. Let x E X and Xo := PMPN(X). To
show Xo = PMnN(X), it suffices by Theorem 4.9 to show that Xo E M n Nand
x - Xo E (M n N)l.. For every y E M n N,

(x - xo,Y) = (x,y) - (PMPN(X),y) = (x,y) - (PN(X),PM(y)


= (x, y) - (PN(X), y) = (x, y) - (x, PN(y) = (x, y) - (x, y) = 0.

Thus X-Xo E (MnN)-L. Clearly, Xo = PMPN(X) EM. Also, for every Z E Nl.,

(xo,z) = (PMPN(X),z) = (PNPMPN(X),z) (since PMPN(X) EN by (3))


= (PMPN(X),PN(Z) = (PMPN(X), 0) (by Theorem 5.8(4))
=0.

Thus Xo E Nl.l. = N using Theorem 4.5(8). Hence Xo EM n N, and this proves


Xo = PMnN(X).
The implication (4) ==? (5) is obvious.
(5) ==? (1). Let P = PMPN be the orthogonal projection onto some Chebyshev
subspace. By Lemma 9.1, (PMPN)* = PMPN and (PM PN )2 = PMPN. Using
Lemma 8.30, we obtain PMPN = (PMPN )* = PNPM = PNPM so (1) holds.
The last statement of the theorem follows from the equivalence of statements (1)
through (3). •
This theorem says that if the projections PM and P N commute, the product
PM PN is the projection PMnN onto M n N. What can be said if PM and PN do
THE CASE OF TWO SUBSPACES 195

not commute? We will see that if PM PN is iterated, then the iterates of PM PN


converge pointwise to P MnN .
More precisely, let Ml and M2 be closed subspaces of the Hilbert space X, let
x E X be arbitrary, and define the "alternating" sequence {x n } as follows:

Xo = x,
(9.2.1 ) X2n-l = P Ml (X2n-2),
X2n = P M2 (X2n-l) (= (PM2 PMl t(X)) (n = 1,2, ... ).

(see Figure 9.2.3 below). Then the next theorem shows that

(9.2.2)

Figure 9.2.3. Alternating projections

9.3 Von Neumann Theorem. Let Ml and M2 be closed subspaces in the


Hilbert space X. Then for each x EX,

(9.3.1)

Proof. For simplicity of notation, let Pi = P Mi (i = 1,2). Fix any x E X and


let

Xo = x,
(9.3.2) X2n-l = PI (X2n-2) (= P 1(P2P1)n-lx ),
X2n = P2(X2n-l) (= (P2P1)n X),

for n = 1,2, .... To complete the proof, it suffices to show that X2n -+ P Mln M2 (x).
In fact, we will prove that the whole sequence {xn} converges to PMlnM2(X).
196 THE METHOD OF ALTERNATING PROJECTIONS

Note that since IlPi I :::; I,


IIx2nli = IIP2X2n-lll :::; Il x 2n-lll = IIP1X2n-211 :::; Il x 2n-211·
This proves that the sequence of norms {llxnll} is decreasing, and hence
(9.3.3) '\:= lim Ilxn I
n-+oo
exists.
Using the facts that the Pi are self-adjoint (Theorem 5.13), X2n E M 2, and
X2n-1 E M 1, we deduce that
(X2n,X2m/ = (P2X2n-I,X2m/ = (X2n-l,P2X2m/ = (X2n-I,X2m/
= (P1X2n-I,X2m/ = (X2n-I,P1X2m/ = (X2n-I,X2m+I/.
That is,
(9.3.4)
Similarly, we can show that
(9.3.5) (X2n+l,X2m+l/ n,m E N.
= (X2n,X2m+2/,
Combining (9.3.4) and (9.3.5), we deduce that if n 2': m, then
(X2n,X2m/ = (X2n-I,X2m+l/ = (X2n-2,X2m+2/ = ... = (X2n-k,X2m+k/
for any integer 0 :::; k :::; 2n. In particular, for k = n - m, we obtain
(X2n, X2m/ = (xn+m,xn+m/ = Ilxn+mI12.
Thus, if m :::; n,
IIX2n - x2ml1 2 = IIx2nl12 - 2(X2n,X2m/ + IIx2ml12
= IIx2nl12 - 211xn+m112 + IIx2ml12
---+ ,\2 _ 2,\2 +,\2 = 0
as m --'; 00. This shows that {X2n} is a Cauchy sequence in M 2. Since X is complete
and M2 is closed, there exists y E M2 such that X2n --'; y. Also, since
(X2n, X2n-l/ = (P2X2n, X2n-l/ = (X2n, P2X2n-l/ = (X2n, X2n) = IIx2n 112
implies that
IIx2n - X2n_111 2 = IIx2n 112 - + Ilx2n-111 2
2(X2n, X2n-l/
= -llx2nl12 + Ilx2n-I!I2 ---+ _,\2 +,\2 = 0
as n --'; 00, it follows that
Il x 2n-1 - yll :::; Il x 2n-l - x2nll + II x 2n - yll ---+ O.
Since X2n-1 E MI for all nand MI is closed, y E MI. Thus, we have shown that
(9.3.6) limx n = y E MI n M 2·
To verify that y = PM ,nM2 (x), it suffices, by Theorem 4.9, to show that x - y l.
lvIl n M 2 . To this end, let Z E Ml n M 2. Then, using Theorem 4.9,
(X2n, z/ = (X2n - X2n-1 + X2n-l, z/ = (P1X2n-1 - X2n-l + X2n-l, z) = (X2n-l, z/.
Similarly,
(X2n-I, z/ = (X2n-2, z).
By induction, it follows that (X2n, z) = (xo, z/ = (x, z) for all n. Thus
(x - y, z/ = (x, z/ - (y, z/ = (x, z/ - lim(x2n, z/ = (x, z/ - (x, z) = O. •
ANGLE BETWEEN TWO SUBSPACES 197

Remarks. (1) An inspection of the proof shows that the theorem holds, more
generally, whenever Ml and M2 are any complete (hence Chebyshev) subspaces in
any inner product space X that is complete or not.
(2) Geometrically, it appears that the rate of convergence of the alternating
sequence {x n } defined in the theorem (see Figure 9.2.3) depends on the "angle"
between the subspaces Ml and M 2 . We shall see below that this is indeed the case.

Angle Between Two Subspaces

9.4 Definition. The angle between two subspaces M and N is the angle
a(M,N) between 0 and 7r/2 whose cosine, c(M,N) = cosa(M,N), is defined
by the expression

c (M, N) := sup{l(x, y) II x EM n (M n N)l., Ilxll ::::: 1,


(9.4.1)
yEN n (M n N)l., Ilyll ::::: I}.

Some basic facts about the angle between two subspaces are recorded next.

9.5 Lemma. Let M and N be closed subspaces in the Hilbert space X. Then:
(1) 0::::: c(M,N)::::: L
(2) c(M,N)=c(N,M).
(3) c(M,N) = c (M n (M n N)l.,N n (M n N)l.).
(4) c(M,N)=sup{l(x,y)llxEM, Ilxll:::::1,YENn(MnN)l., Ilyll:::::l}.
(5) c(M, N) = sup{l(x, y)11 x E M n (M n N)l., Ilxll ::::: 1, YEN, Ilyll ::::: I}.
(6) I(x, y) I ::::: c (M, N) /lxll Ilyll whenever x E M, YEN, and at least one of x
or y is in (M n N)l..
(7) c(M,N) = IIPNPM -PMnNII = IIPNPMP(MnN).l.11
= IIPMn(MnN).l.PNn(MnN).l.II·
(8) c (M, N) = 0 if and only if PM and P N commute. (In this case, P N PM =
PMnN .) In particular, c(M,N) = 0 if MeN or N c M.

Proof. Statements (1) and (2) are obvious.


(3) Let 8 = M n (M n N)l. and T = N n (M n N)l.. Then 8 n T = {O}, so
(8 n T)l. = X and

c(M,N) = sup{l(x,y)1 I x E 8, Ilxll::::: 1, yET, Ilyll::::: I}


= sup{l(x,y)11 x E 8n (8nT)l., Ilxll::::: 1, y E Tn (8 nT)l., Ilyll ::::: I}
=c(8,T).

(4) Using the readily verified facts that

M = MnN +Mn (MnN)l.

and
198 THE METHOD OF ALTERNATING PROJECTIONS

(see Exercise 18 at the end of the chapter), we deduce that

sup{l(x,y)11 x E M, Ilxll :':: 1, yEN n (M n N)-L, Ilyll :':: I}


= sup{I(PMx, PNn(MnN)l-y)llllxll :':: 1, Ilyll :':: I}
= sup{I(PMnNX + PMn(MnN)l-X, PNn(MnN)l-y)llllxll :':: 1, Ilyll :':: I}
= sup{I(PMn(MnN)l-X, PNn(MnN)l-Y) I I Ilxll :':: 1, Ilyll :':: I}
= sup{l(x,y)11 x E Mn (MnN)-L, Ilxll:,:: 1, yEN n (M nN)-L, Ilyll :':: I}
= c(M,N).

(5) Interchange the roles played by M and N in (4) and use (2).
(6) This follows easily from (4) and (5).
(7) Clearly,

c(M,N)
= sup{l(x, y)1 I x E M n (M n N)-L, Ilxll :':: 1, yEN n (M n N)-L, Ilyll:':: I}
= sUp{I(PMn(MnN)l-X,PNn(MnN)l-y)llllxll :':: 1, Ilyll :':: I}
= SUp{I(PNn(MnN)l-PMn(MnN)l-X,y)1 I Ilxll :':: 1, Ilyll :':: I}
= IIPNn(MnN)l-PMn(MnN)l-II,

where the last equality follows from (8.0.5). Using (2), this verifies the third equality
of (7).
Since MnN c M, PMnN = PMPMnN = PMnNPM by Theorem 5.14. Hence

PMP(MnN)l- = PM (I - PMnN) = PM - PMPMnN = PM - PMnNPM


= (I - PMnN)PM = P(MnN)l-PM .

That is, PM commutes with P(MnN)l-· Similarly, PN commutes with P(MnN)l-.


From Lemma 9.2, PMn(MnN)l- = PMP(MnN)l- and PNn(MnN)l- = PNP(MnN)L
Thus

c(M,N) = IIPNn(MnN)l-PMn(MnN)l-11 = IIPNP(MnN)l-PMP(MnN)l-11


= IIPN PM P(MnN)l-11,

which proves the second equality of (7). But

since M nNe M and M nNe N. This proves the first equality of (7).
(8) From (7) we see that c (M, N) = 0 if and only if PN PM = PMnN . By Lemma
9.2, this means that PN and PM commute. The last statement also follows from
Lemma 9.2. •
9.6 Examples. (1) Consider the subspaces (Le., "lines") in £2(2) given by

M = {x E £2(2) I x(l) = x(2)} and N={XE£2(2) I x(2) =O}

(see Figure 9.6.1). Then c(M,N) = 1/V2, and hence a(M,N) = 7r/4.
ANGLE BETWEEN TWO SUBSPACES 199

Figure 9.6.1

To see this, note that M n N = {OJ, and if x E M n (M n N)1- = M, [[xII ~ 1,


and y E Nn(MnN)1- = N, [[y[[ ~ 1, then [x(1)[ = [x(2)[ ~ 1/v'2, [y(l)[ ~ 1, and

(x,y) = x(l)y(l) ~ 1/v'2,


with equality holding when x(1) = x(2) = 1/v'2 and y(l) = 1.
(2) Consider the subspaces (i.e., "planes") in £2(3) given by
M = {x E £2(3) I x(3) = O} and N = {x E £2(3) I x(2) = O} (see Figure 9.6.2).
Then c (M, N) = 0 and hence a(M, N) = 7r /2. We leave the verification of this as
an exercise.
Remark. Note that when c (M, N) < 1, Theorem 9.5(6) represents a strength-
ening of the Schwarz inequality for certain pairs of vectors x and y. Moreover, the
rate of convergence of the method of alternating projections given in Theorem 9.3
is governed by the factor c (M, N)2n-l (see Theorem 9.8 below). For these reasons,
it is important to know when c (M, N) < 1. It will be verified later in Theorem
9.35 that c (M, N) < 1 if and only if M + N is closed. To prove this, we need first
to discuss some results concerning "weak convergence." However, at this point we
can establish another characterization of when M + N is closed.
9.7 Theorem. Let M and N be closed subspaces of the Hilbert space X such
that M + N is closed. Then
(9.7.1)

In particular, M + N is closed if and only if M 1- + N 1- is closed.


Proof. First observe that by Theorem 4.6(4),

(9.7.2)
200 THE METHOD OF ALTERNATING PROJECTIONS

3
N

-)--_ _ _ _ ~M

Figure 9.6.2

Thus to verify (9.7.1), it suffices to show that

(9.7.3)

To this end, let x E (M n N)l... Then Z := M + N is a closed subspace of X


so is a Hilbert space in its own right. If some z E Z has two representations
Z = Ul + Vl = Uz + vz, where Ui E M, Vi E N, then Ul - Uz = Vz - Vl E M n N, so
(Ul - uz,x) = 0 or (Ul'X) = (uz,x). It follows that the functional f, defined on Z
by

(9.7.4) f(u + v) := (u,x) for each U E M and v E N,

is well-defined, linear, and satisfies

(9.7.5) f(v) = 0 for every v E N.

We next show that f is bounded on Z, i.e., f E Z*. First we observe that A :


M x N -+ M + N, the "addition" mapping on M x N defined by A(u, v) = U + v,
is bounded, linear, and onto. The linearity and the surjectivity of A are clear. To
see that A is bounded, we note that for each (u, v) E M x N \ {(O, O)},

IIA(u, v)11
II(u, v)11
This proves that A is bounded (and, in fact, IIAII :s: V2). Since M + N is closed,
hence complete, in X, it follows by the open mapping theorem (Theorem 8.17) that
there exists b > 0 such that A(B(M x N)) ~ bB(M + N), where B(E) denotes the
RATE OF CONVERGENCE (TWO SUBSPACES) 201

unit ball in E. In particular, for each u E M, v E N with Ilu + vii s 1, there exists
(x, y) E H(M x N) such that x + y = A(x,y) = J(u + v). Then

1 1 1 1
!f(u + v)1 = Jlf(x + y)1 = JI(x,x)1 s Jllxllllxli S Jllxll.
Thus Ilfll S ~ Ilxll and f is bounded on Z.
By the Fnkhet-Riesz representation theorem (Theorem 6.10), there exists ¢ E Z
such that

(9.7.6) fez) = (z, ¢) for each z E Z.

In particular,

(9.7.7) f(u+v) = (u+v,¢) for each u E M and v E N.

By (9.7.5) it follows that ¢ E N1-. That is,

(9.7.8) f(u+v)=(u,¢) for each u E M and v E N.

Combining (9.7.4) and (9.7.8), we deduce (u,x) = (7L,¢) for every u E M, and so
x - ¢ E M 1.. Hence
x = (x - ¢) + ¢ E M1. + N1.,
and this completes the proof of (9.7.3). This proves (9.7.1).
In particular, since (M n N)1- is dosed, M1. + N1. must be dosed whenever
M + N is dosed. Conversely, if M 1. + N 1. is dosed, then the above proof shows
that M1.1. +N1.1. is dosed. However, by Theorem 4.5(8), we have M1.1. = M and
N1.1. = N. This proves that M + N is dosed when M1.1. + N1.1. is. •

Rate of Convergence for Alternating Projections (Two Subspaces)


The next theorem shows that the rate of convergence of the method of alternating
projections is fast (respectively slow) if the angle between the subspaces is large
(respectively small).
9.8 Rate of Convergence. Let MI and M2 be closed subspaces in the Hilbert
space X and c = c (MI , M2). Then, for each x E X,

for n = 1,2, .... Moreover, the constant C2n - 1 is the smallest possible independent
ofx.
Proof. Let M = MI n M 2 , Pi = P Mi (i = 1,2), and P = P2 P I . We first will
prove that

(9.8.2)

and hence

(9.8.3)
202 THE METHOD OF ALTERNATING PROJECTIONS

To see (9.8.2), let x E M.l. and y E M. Then, since P;, is self-adjoint and
idempotent,
(Pi(x),y) = (x,Pi(y)) = (x,y) = o.
Thus Pi(x) E M.l., which proves (9.8.2). Relation (9.8.3) follows, since

Next we verify that

(9.8.4) IIP(x)11 :::; ell PI (x) II :::; ellxll for all x E M.l..

To see this, let x E M.l.. Then

IIP(x)11 2 = (P(x) , P(x)) = (P(x) - PI (x) + Pr(x), P(x))


= (PI (x), P(x)) since P(x) - PI (x) E Mt , P(x) E M2
:::; eIIP1(x)llllF(x)11 by Lemma 9.5, (9.8.2), and (9.8.3)
:::; ellxIIIIP(x)11 since IlFill :::; 1.
Dividing both sides of the inequality by IIP(x)ll, we obtain (9.8.4).
Now suppose x E P(M.l.). Then x E M 2nM.l. by (9.8.3). Since PI (x) E M1nM.l.
by (9.8.2), we obtain, using Lemma 9.5 (6),

IIP1(x)11 2 = (P1(x), PI (x)) = (x,P1(x)):::; ellxIIIIPl(X)II·


Thus IIPI (x) II :::; ellxll. Substituting this into (9.8.4) yields

(9.8.5)

From (9.8.3), (9.8.4), and an induction on (9.8.5), we get that

II (P2P1)n(X) - PM(x)11 = Ilpn(x) - pnpM(X) II = Ilpn[x - PM (x)lII


= Ilpn-1[P(x - PM (x))lII :::; e2(n-l)IIP(x - PM (x)) II
:::; e2n - 1llx _ PM (x) II :::; e2n- 11Ixll·
This verifies (9.8.1).
The fact that the constant Jln-l is smallest possible, independent of x, will be
established in Theorem 9.31 below. •
Remarks. If e (Ml' M 2) = 0, then Lemma 9.5 implies that PM2 PMl = PMln M2'
and hence the iteration converges in one step (n = 1).
(2) We will see later (Theorem 9.34) that e (Ml' M2) < 1 if and only if Ml +M2 is
closed. This will be the case, for example, if either Ml or M2 is finite-dimensional.
(3) We shall also see later (Corollary 9.28) that Theorem 9.3 holds for more than
two subspaces. However, if the subspaces are replaced by arbitrary closed convex
sets, the analogous result fails.
To see this, consider the following example, in the plane £2(2) (see Figure 9.8.6
below). Let

01 = {x E £2(2) I x(2) ~ x(l) ~ O} and O2 = {x E £2(2) I x(2) = I}.


WEAK CONVERGENCE 203

Thus C 1 is a convex cone (the upper half of the first quadrant), and C 2 is a hori-
zontalline (one unit above the horizontal axis). Setting C = C 1 n C2 , we see that
C is the line segment joining the points (0,1) and (1,1).
Let Xo = (1, -1). Then Xl = Pc, (xo) = (0,0) and X2 = Pc, (xr) = (0,1). Since
X2 E C, it follows that all the terms of the sequence
(n = 2,3, ... )
are equal to X2. In particular,

However, it is easy to verify that

This shows that Theorem 9.3 fails if the two subspaces are replaced by more general
convex sets.

' · X=Xo=(J.-J)

Figure 9.8.6

In spite of this example, there is a variant of Theorem 9.3 that is valid for general
closed convex sets, not necessarily subspaces. Before turning to this, we need to
discuss some aspects of weak convergence in a Hilbert space.

VVeak Convergence

9.9 Definition. A sequence {xn} in an inner product space X is said to con-


verge weakly to X, written Xn ~ X or Xn --+ X weakly, if

(Xn, y) --+ (x, y) for every y E X.


204 THE METHOD OF ALTERNATING PROJECTIONS

Equivalently, Xn --+ x weakly if and only if x*(xn) --+ x*(x) for each x* E
X*. This follows obviously from Theorem 6.10 if X is complete, and slightly less
obviously from Theorem 6.S if X is not complete.
It is an easy consequence of the definition that weak limits are unique, and if
Xn ~ x and Yn ~ y, then aXn + j3Yn ~ ax + j3y. The relationship between
ordinary (or "strong") convergence and weak convergence can be summarized as
follows.
9.10 TheoreIll (Weak vs Strong Convergence).
(1) If Xn --+ x, then Xn ~ x (but not conversely in general).
(2) Xn --+ x if and only if Xn ~ x and Ilxn II --+ Ilxll·
(3) If X is finite-dimensional, then Xn ~ x if and only if Xn --+ x.
Proof. (1) Suppose Xn --+ x. Then for each Y E X,

so that (xn' y) --+ (x, y). Thus Xn ~ x. An example that shows that the converse
fails is given following this theorem.
(2) If Xn --+ x, then Xn ~ x by (1), and Ilxnll --+ Ilxll by Theorem 1.12.
Conversely, if Xn ~ x and Ilxnll --+ Ilxll, then

That is, Xn --+ x.


(3) If dim X = N, let {el' e2, .. . , eN} be an orthonormal basis for X. Then by
Corollary 4.10, we have that for each x EX,
N N
X = 2:)x, ei)ei and IIxI1 2= LI (x, ei) 12.
1

Hence if Xn ~ x, then
N N

IIxnl1 2 = LI(xn,eiW --+ LI(x,eiW = Ilx112.


1 1

Hence, by (2), Xn --+ x. The converse follows by (1). •


9.11 Exrunple (Weak convergence does not iIllply norIll convergence).
Consider the space £2 and the orthonormal sequence {xn} defined by xn(i) = Oni
for all n, i E N. Then for each y E £2,

(Xn, y) = y(n) --+ 0= (0, y),

since I:~ ly(n)j2 = IIyl12 < 00. Thus Xn ~ 0. But Ilxnll = 1 for all n, so that
{xn} does not converge to O.
The most important property of weak convergence, for our purposes, is the fol-
lowing result, which states that bounded sets satisfy a certain "weak compactness"
criterion.
WEAK CONVERGENCE 205

9.12 Weak Compactness Theorem. Every bounded sequence in a Hilbert


space has a weakly convergent subsequence.
Proof. Let {xn} be a sequence in the Hilbert space X, with Ilxnll S; c for all n.
Consider the sequence ofreal numbers {(Xl,Xn)}' This sequence is bounded (since
I(Xl,Xn)1 S; IlxI11 Ilxnll S; c2), so that there exists a subsequence that converges:
(XI,Xl(n)) --+ 6 ERas n --+ 00. Similarly, {(X2,Xl(n))} is bounded, so it also has a
convergent subsequence: (X2' X2(n)) --+ 6 as n --+ 00. Continuing in this fashion,
we see that for each kEN, there is a subsequence {Xk(n)} of {X(k-l)(n)} such that
(Xk, Xk(n)) --+ ~k as n --+ 00. Then the "diagonal" sequence {xn(n)} has the property
that {xn(n)} is a subsequence of {xn} such that for each kEN,

(9.12.1 )

Let
M = {x E X !li;n(x, Xn(n)) eXists}

and define f on M by
(9.12.2) f(x) := lim(x,xn(n)), x E M.
n

By (9.12.1), Xk EM (k = 1,2, ... ). Also, if x, Y E M and a,!3 E R, then

lim(ax + !3y, Xn(n)) = lim[a(x, Xn(n)) + !3(y, xn(n)) 1 = af(x) + !3f(y).


n n

Thus ax + !3y EM. This proves that M is a linear subspace and that f is linear
on M. Also, for any x E M,

If(x) I = lim
n
I(x, Xn(n)) I S; lim sup Ilxll Ilxn(n) II S; cIIxll·
n

Thus f is bounded on M and Ilfll S; c.


Next we show that M is closed. Fix any Y E M. Then there exists Ym E M such
that Ym --+ y. Since

If(Yn) - f(Ym)1 S; IlfllllYn - Ymll S; ellYn - Ymll

°
and {Yn} is Cauchy, it follows that {f(Yn)} is Cauchy in R Thus limn f(Yn) exists.
If also y~ EM and y~ --+ y, then Yn - y~ --+ and

If(Yn) - f(Y~)1 S; IlfllllYn - Y~II -+ 0.

This proves that F(y) := limn f(Yn) exists and is independent of the particular
sequence {Yn} in M that converges to y. Given E > 0, choose an integer NI such
that

(9.12.3) clly - YN111 < E/3 and If(YN,) - F(Y)I < E/3.

Then choose an integer N such that for all n 2': N,

(9.12.4)
206 THE METHOD OF ALTERNATING PROJECTIONS

Thus if n 2: N, then (9.12.3) and (9.12.4) imply that

l(y,xn(n)/- F(y)l:S l(y,Xn(n)/- (YN"Xn(n)/1 + I(YN"Xn(n)/- f(YN,)1


+ If(YN,) - F(Y)I
< Ily - YN,II Ilxn(n) II + 2E/3 :S clly - YN,II + 2E/3 < E.
This shows that F(y) = limn(y,Xn(n)1 exists and thus Y E M. Hence M is closed.
Since M is a closed subspace of the complete space X, M is a Hilbert space in
its own right, and M is a Chebyshev subspace of X. Theorem 6.10 implies that f
has a representer Xo E M; that is,

(9.12.5) f(x) = (x,Xo/, xEM.


Moreover, Theorem 5.8 implies that each x E X may be written as x = PM(X) +
PM~ (x). Since Xo and xn(n) are in M for all n, we obtain from (9.12.2) and (9.12.5)
that

This shows that xn(n) ~ Xo and completes the proof. •


Next we show that weakly convergent sequences are bounded and the norm is
"weakly lower semicontinuous."
9.13 Lemma. Let the sequence {xn} converge weakly to x. Then {xn} is
bounded and Ilxll :S liminfllxnll·
n
Proof. For any Y E X define fi on X* by

fi(x*) := X*(y) , x* E X*.


It is clear that fi : X* -+ lR is linear and

lfi(x*) I = Ix*(y)1 :S Ilx*llllyll, x* E X*.


Thus fi is bounded and lifill :S Ilyli. By Remark (2) after Theorem 6.23, choose
x* E X* with Ilx*11 = 1 and x*(y) = Ilyli. Then fi(x*) = x*(y) = Ilyll implies that
lifill 2: Ilyll and hence lifill = Ilyll· We have shown that fi E X** = B(X*,lR) and
lifill = Ilyll· Since X* is a Hilbert space (by Exercise 7 or 8 of Chapter 6), and since

xn(x*) = x*(xn) -+ x*(x) = x(x*)


for each x* E X*, it follows by Corollary 8.15 that

supllxnll = supllxnll < 00.


n n

Thus {xn} is bounded.


Again using Remark (2) after Theorem 6.23, choose x* E X* with Ilx* II = 1 and
x*(x) = Ilxll. Thus
Ilxll = x*(x) = limx*(xn) :S liminf Ilx*llllxnll = liminf Ilxnll· •
The following lemma shows that each continuous linear operator is also "weakly
continuous."
DYKSTRA'S ALGORITHM 207

9.14 Lemma. Let X and Y be inner product spaces with X complete, and let
A E B(X, Y). Then AXn -+ Ax weakly whenever Xn -+ x weakly.
Proof. Let Xn -+ x weakly. Then for each y E Y,

(Axn' y) = (xn, A*y) --+ (x, A*y) = (Ax, y).

That is, AXn -+ Ax weakly. •


9.15 Definition. A subset K of an inner product space X is called weakly
closed if Xn E K and Xn ~ x implies x E K.
The last result we need that concerns the weak topology is that a convex set is
closed if and only if it is weakly closed.
9.16 Theorem. (1) Every weakly closed set is closed.
(2) A convex set is closed if and only if it is weakly closed.
Proof. (1) Let K be weakly closed, Xn E K, and Xn -+ x. To show that K is
closed, it suffices to show that x E K. By Theorem 9.10 (1), Xn ~ x. Since K is
weakly closed, x E K.
(2) Owing to (1), if suffices to show that if K is closed and convex, then K is
weakly closed. Let Xn E K and Xn ~ x. If x rt K, then by Theorem 6.23, there
exists x* E X* with Ilx*1I = 1 and x*(x) > supx*(K). It follows that

x*(x) > sup x*(xn) 2': limx*(xn) = x*(x),


n n

which is absurd. Thus x E K, and hence K is weakly closed. •

Dykstra's Algorithm
Now we can describe, and prove the convergence of, an iterative algorithm called
Dykstra's algorithm that allows one to compute best approximations from an inter-
section nr Ki of a finite number of closed convex sets Ki (not necessarily subspaces),
by reducing the computation to a problem of finding best approximations from the
individual sets K i .
Let Ki be a closed convex subset of the Hilbert space X (i = 1,2, ... , r), and
let K := nrKi . We assume K =t 0. Given any x E X, we would like to compute
PK(X),
For each n E N, let [n] denote "n mod r"; that is,

[n] := {I, 2, ... , r} n {n - kr I k = 0, 1,2, ... }.


Thus [1] = 1, [2] = 2, ... , [r] = r, [r + 1] = 1, [r + 2] = 2, ... , [2r] = r, .... For a
fixed x E X, set

Xo : = x, e_(r-l) = ... = e_l = eo = 0,


Xn : = PK[n] (Xn-l + en- r ),
(9.16.1)
en : = Xn-l + en- r - Xn
= Xn-l + en- r - PK[n] (Xn-l + en- r ) (n = 1,2, ... )

(see Figure 9.16.2). We will show that

(9.16.3) lim Ilxn - PK(x)11 = 0.


n
208 THE METHOD OF ALTERNATING PROJECTIONS

.
:' ,I"
'

.'
','

X=Xo

Figure 9.16.2. Dykstra's algoritlun (r = 2)

It is convenient to decompose the proof of (9.16.3) into several smaller pieces,


which we designate as lemmas. For notational simplicity, we set Pi := P Ki (i =
1,2, ... , r) .
9.17 Lemma. For each n,
(9.17.1)

Proof. Using Theorem 4.1 with x = Xn - l + e n - r , Yo = p[nj(Xn-l + e n - r ), and


K = K[nj, we obtain, for all y E K[nJ'
(xn - y, en) = (p[nj(Xn-l + en - r ) - y , Xn-l + en - r - p[nj(Xn-l + en-r» 2: O. •
9.18 Lemma. For each n 2: 0,
(9.18.1) x - Xn = en-(r-l) + e n -(r-2) + ... + en- l + en-

Proof. By induction on n. For n = 0, X-Xo = x-x = 0 and e_(r- l) +e_(r -2) +


... + e _ l + e o = 0 by
(9.16.1). Now assume that the result is valid for some n 2: O.
Then
x - Xn+l = (x - xn) + (xn - x n +!)
= (en-(r -l) + e n -(r - 2) + ... + en-l + en) + (en+! - en+l - r )
= e n - (r-2) + e n -(r-3) + ... + en + en+l
= en+!-(r-l) + e n +l-(r - 2) + ... + en + en+l,
which shows that (9.18.1) is valid when n is replaced by n + 1. •
DYKSTRA'S ALGORITHM 209

9.19 Lemma. For each n E N, 0::; m ::; n, and Y E K,


n n
Ilxm - yII 2 = IIxn - YI12 + L Ilxk - Xk_111 2 + 2 L
(ek-r, Xk-r - Xk)
k=m+l k=m+l
(9.19.1)
n m
+2 E (ek,xk - y) - 2 E
(ek,xk - y).
k=n-(r-l} k=m-(r-l}

Proof. For any set of vectors {Ym, Ym+ 1, ... , Yn+ I} in X, we have the identity
IIYm - Yn+!11 2 = II(Ym - Ym+l) + (Ym+!- Ym+2) + ... + (Yn - Yn+l)11 2
n+l
= E IIYk-l - Ykl1 2 + 2
k=m+l m+l~i<j~n+l
n
(9.19.2) = llYn - Yn+111 2 + L IIYk-l - Yk 112
k=m+l

But

Ln L
( n+l
(Yi-l - Yi, Yj-l - Yj)
)
= Ln (
Yi-l - Yi, E (Yj-l -
n+l
Yj)
)
i=m+l j=i+l i=m+l j=i+l
n
(9.19.3) = E (Yi-l - Yi, Yi - Yn+l).
i=m+l
Substituting Yi = Xi for all i ::; n and Yn+! = Y into (9.19.3), we get
n n
L (Yi-l - Yi, Yi - Yn+!) = L (Xi-l - Xi, Xi - y)
i=m+l i=m+l
n
= E (ei - ei-r,Xi - y)
i=m+l
n n
= E (ei,Xi - y) - L (ei-roXi - y)
i=m+l i=m+l
n n
= E (ei,Xi-Y)- L [(ei-r,Xi-Xi-r) + (ei-r,Xi-r-Y)]
i=m+l i=m+l
n n n
= E (ei, Xi - y) - L (ei-r, Xi-r - y) + L (ei-r, Xi-r - Xi)
i=m+l i=m+l i=m+l
n n-r n
= E (ei,Xi - y)- L (ei, Xi - y) + L (ei-r, Xi-r - Xi)
i=m+l i=m+l-r i=m+l
n m n
E (ei, Xi - y) - L (ei, Xi - y) + L (ei-r, Xi-r - Xi).
i=n-r+l i=m-r+l i=m+l
2lO THE METHOD OF ALTERNATING PROJECTIONS

Substituting these same values for Yi into (9.19.2) and using the above relation,
we get

n n
Ilxm- YI12 = Ilxn - YI12 + L Ilxk-1 - Xk 112 + 2 L (ei' Xi - y)
k=m+l i=n-r+l
n m
+ 2 L (ei-r, Xi-r - Xi) - 2 L (ei' Xi - y).
i=m+l i=m-r+l
This verifies (9.19.1). •
9.20 Lemma. {xn} is a bounded sequence and

=
(9.20.1)

In particular,

(9.20.2) Ilxn-l - xnll ~ 0 as n ~ 00.

Proof. Setting m = 0 in (9.19.1), we note that the third and fourth terms on
the right of (9.19.1) are nonnegative by Lemma 9.17, while the last term is zero
using (9.16.1). Thus we obtain

n
Ilxo - YI12 ~ Ilxn - YI12 + Lllxk - xk_Iil 2
k=l
for all n. It follows that {llxn - yll} is bounded and (9.20.1) holds. Hence {xn} is
bounded as well. Finally, (9.20.1) implies (9.20.2). •
9.21 Lemma. For any n E N,

n
(9.21.1) Ilenll ::; Lllxk-l - xkll·
k=l
Proof. We induct on n. For n = 1,

Ilelll = Ilxo - Xl - el-rll = Ilxo - xIII,


since ei = 0 for all i ::; 0 by (9.16.1). Now assume that (9.21.1) holds for all m ::; n.
Then

n+l-r n+l
::; Ilxn - xn+111+ L Ilxk-1 - xkll ::; Lllxk-l - xkll
k=l k=l
implies that (9.21.1) holds when n is replaced by n + 1. •
DYKSTRA'S ALGORITHM 211

9.22 Lemma.
n
(9.22.1) limninf 2: I(Xk - Xn , ek)1 = O.
k=n-(r-l)

Proof. Using Schwarz's inequality and Lemma 9.21, we get


n n
2: I(Xk-Xn,ek)l:::: 2: Ilekllllxk - Xnll
k=n-(r-l) k=n-(r-l)

Letting ai = Ilxi-l - xiii, we see that it suffices to show that

(9.22.2)

But
00 00

A:= 2:a; = 2:llxi-l - xil1 2 < 00


1 1

by Lemma 9.20. Using the Schwarz inequality, we get

for each n. Hence to verify (9.22.2), it is enough to verify


n
(9.22.3) limninf Vn 2: ai = o.
i=n-(r-2)

To establish (9.22.3), let


n
p := lir~,inf Vn 2: ai·
i=n-(r-2)

If (9.22.3) were false, then we would have p > O. We may assume p < 00. (The
proof for p = CXl is similar.) Then

n 1
In
V" ""
L..t ai > 2" p "eventually"
i=n-(r-2)
212 THE METHOD OF ALTERNATING PROJECTIONS

(i.e., for n sufficiently large). Thus Z=~=n-(r-2) ai > p/(2.,fii) eventually. Using
Schwarz's inequality, we deduce that

(. t
,=n-(r-2)
ai) 2 ::::: (r - 1) t
i=n-(r-2)
a; eventually.

Thus for some integer N,


p2 n
-4n < (r - 1) '"
L.J'
a 2 for all n 2 N.
i=n-(r-2)

Hence
n

L: a;
i=n-(r-2)
00 00

::::: (r - 1) L: [a;_(r_2) + a;-(r-3) + ... + a;] ::::: (r - l?L:a; < 00,


n=N 1

which is absurd. This contradiction proves that (9.22.3) must hold. •


9.23 Lemma. There exists a subsequence {x nj } of {xn} such that

(9.23.1) lim sup (y - Xnj , X - x nj ) ::::: 0 for each y E K, and


j

nj
(9.23.2) lim L: I(Xk -xnj,ek)1 = o.
J k=nj -(r-l)

Proof. Using Lemma 9.18, we have for all y E K, n 2 0, that

(y - Xn , X - xn) = (y - X n , en-(r-l) + e n -(r-2) + ... + en)


n
L: (y - xn,ek)
(9.23.3) k=n-(r-l)

k=n-(r-l) k=n-(r-l)

By Lemma 9.17, the first sum is no more than O. Hence

(9.23.4)
k=n-(r-l)

By Lemma 9.22, we deduce that there is a subsequence {nj} of N such that


nj
(9.23.5) lim L: I(Xk - x nj ' ek)1 = O.
J k=nj-(r-l)
DYKSTRA'S ALGORITHM 213

Note that the right side of (9.23.4) does not depend on y. In view of (9.23.4), it
follows that (9.23.5) implies that (9.23.1) and (9.23.2) hold. •
Now we are ready to state and prove the main theorem. Recall our setup: X is
a Hilbert space, K 1 , K 2 , ... , Kr are closed convex subsets, and K = nlKi i 0. For
each x EX, define
Xo = x, e_(r-l) = ... = e_l = eo = 0,
(9.23.6) Xn = PK[n] (Xn-l + en - r ),
(n = 1,2, ... ),
where
[nJ = {1,2, ... ,r} n {n - krlk = 0, 1,2, ... }.
9.24 Boyle-Dykstra TheoreIn. Let Kb K2' ... ' Kr be closed convex subsets
of the Hilbert space X such that K := nlKi i 0. For each x E X, define the
sequence {xn} as in (9.23.6). Then
(9.24.1)

Proof. By Lemma 9.23, there exists a subsequence {x nj } such that

(9.24.2) lim sup (y - x nj ' X - x nj ) :S; 0 for each y E K.


j

Since {xn} is bounded by Lemma 9.20, it follows by Theorem 9.12 (by passing to
a further subsequence if necessary), that there is Yo E X 3uch that
w
(9.24.3) Xnj ----t Yo,
and

(9.24.4) lim Ilxnj II exists.


J

By Theorem 9.13,

(9.24.5) IIYoll :S; liminf Ilxnj II = lim Ilx nj II·


J J

Since there are only a finite number of sets K i , an infinite number of the xnj's
must lie in a single set Kio. Since Kio is closed and convex, it is weakly closed by
Theorem 9.16, and hence Yo E KiD. By (9.20.2), Xn - Xn-l -+ O. By a repeated
application of this fact, we see that all the sequences {xnj+d, {X nj +2}, {x nj +3}, ...
converge weakly to Yo, and hence Yo E Ki for every i. That is,

(9.24.6) Yo E K.
For any y E K, (9.24.5) and (9.24.2) imply that

(y - Yo, x - Yo) = (y, x) - (y, Yo) - (Yo, x) + Ilyol12


:S; lim[ (y, x) - (y, xnj ) - (x nj , x) + IIX nj 112]
J
= lim(y - X nj ' X - X nj ) :S; O.
J
214 THE METHOD OF ALTERNATING PROJECTIONS

By Theorem 4.1,
(9.24.7)
Moreover, putting y = Yo in the above inequalities, we get equality in the chain of
inequalities and hence
(9.24.8)

and
(9.24.9) lim(yo - x nj ' X - x nj ) = O.
]

By (9.24.3) and (9.24.8), it follows from Theorem 9.10 (2) that


(9.24.10) [[Xnj - YO[[ -+ O.
Hence
(9.24.11)
To complete the proof, we must show,that the whole sequence {xn} converges
to Yo. From equation (9.23.3) with y = Yo and n = nj, we get

(9.24.12) (Yo - x nj ' X - x nj ) =


k=nj-(r-l)
The left side of (9.24.12) tends to zero as j --+ 00 by (9.24.9), while the second sum
on the right tends to zero by (9.23.4). Hence
nj
(9.24.13) lim L (Yo - Xk, ek) = O.
] k=nj-(r-l)
Using Lemmas 9.19 and 9.17 with m = nj and y = Yo, we see that for all n:::: nj,
nj
IIx nj - YO[[2 :::: [[xn - YO[[2 - 2 L (ek,Xk - Yo),
k=nj-(r-l)
or
nj
(9.24.14) [[xn - YO[[2 :::; [[xnj - YO[[2 +2 L (ek,xk - Yo).
k=nj-(r-l)
But both terms on the right of inequality (9.24.14) tend to zero as j --+ 00 by
(9.24.10) and (9.24.13). It follows that limn [[xn - YO[[ = 0, and this completes the
proof. •
Remark. A close inspection of the proof of the Boyle-Dykstra theorem reveals
the following more general result, which does not require completeness of X, but
only that of the K i .
9.25 Theorem. Let K 1 , K 2 , . .. ,Kr be complete convex sets in an inner product
space X such that K := n1Ki =I 0. Then
(1) K is a convex Chebyshev set, and
(2) For each x EO X, the sequence {xn} defined by (9.23.5) satisfies
lim [[xn - PK(x)[[ = O.
n
THE CASE OF AFFINE SETS 215

The Case of Affine Sets

The Dykstra algorithm described in (9.16.1)-(9.16.3) substantially simplifies in


the special case where all the sets Ki are subspaces or, more generally, where all
the Ki are affine sets (i.e., translates of subspaces). Specifically, the computation
of the residuals en may be omitted entirely from the algorithm!
To see this, we first record the following result that characterizes best approxi-
mations from affine sets (see also Exercise 2 at the end of Chapter 4).

9.26 Characterization of Best Approximations from Affine Sets. Let


V be an affine set in the inner product space X. Thus V = M + v, where M is
a subspace and v is any given element of V. Let x E X and Yo E V. Then the
following statements are equivalent:

(1) Yo = Pv(x);
(2) X-YoEMi.;
(3) (x - Yo, Y - v) = 0 for all y E V.

Moreover,

(9.26.1) Pv(x + e) = Pv(x) for all x E X, e E Mi..

Proof. Using Theorem 2.7 (1) (ii), we see that Yo = Pv(x) <==? Yo = PM+v(x) =
PM(x - v) + v <==? Yo - v = PM(x - v). Using Theorem 4.9, we deduce that
Yo-V = PM(x-v) <==? X-Yo = (x-v)-(YO-v) E Mi.. This proves the equivalence
of (1) and (2). The equivalence of (2) and (3) is obvious, since M = V-v.
To prove (9.26.1), let x E X be arbitrary and e E Mi.. Using Theorem 2.7
(1) (ii) again, along with the facts (Theorems 5.7 and 5.12) that PM is linear and
PM(e) = 0, we see that

Pv(x + e) = PM+v(x + e) = PM(x + e - v) + v


= PM(x v) +v = PM+v(x) = Pv(x). •

The geometric interpretation of Theorem 9.26 is that the error vector x - Pv(x)
must be orthogonal to the translate of V that passes through the origin (see Figure
9.26.2). Note that in the special case where V = !vI is a subspace, we recover
Theorem 4.9.
216 THE METHOD OF ALTERNATING PROJECTIONS

, , - - - - - - - - - + - - - - - - 7 Y = M+ V

y •.....

r- Py(X)

Y-'J~m7M
/ Figure 9.26.2
Let us return now to Dykstra's algorithm in the particular case where one of the
closed convex sets Ki is an affine set. For definiteness, suppose KI is a closed affine
set. Then
KI = MI +VI,
where MI is a closed subspace and VI E K I .
If n is any integer with [n] = 1, Theorem 9.26 implies that

en = Xn-I + e n - r - PK[n] (Xn-I+ en - r )


= Xn-I + e n - r - PK, (Xn-I + en - r ) E Mt.

Since [n - r] = 1 also, it follows that en - r E Mf", and (9.26.1) implies that

That is, to determine X n , we need only project Xn-I onto K[n]; it is not necessary
first to add e n - r to Xn-I and then project the sum Xn-I + e n - r onto K[n].
In particular, if all the sets Ki are closed affine sets, then it is not necessary to
compute any of the en's; each term Xn in Dykstra's algorithm may be determined
by projecting Xn-I onto K[nJ=

(9.26.3) (n = 1,2, ... ).


Summarizing all this, we can state the following generalization of Theorem 9.3
to the case of more than two subspaces, or even when the subspaces are replaced
by affine sets.
9.27 Theorem. Let Vi, l/2, ... , Vr be r closed affine sets in the Hilbert space
X such that n'lVi =1= 0. Then for each x E X,

(9.27.1)
RATE OF CONVERGENCE FOR ALTERNATING PROJECTIONS 217

Proof. Fix x EO X. By Dykstra's algorithm, the sequence {xn} determined


by (9.16.1)-(9.16.3) converges to Pn,vi (x). But by (9.26.3), we have that Xn =
PVjn] (Xn-i) for each n :::: 1. In particular, replacing n by nr and inducting on this
relation, we get

Xnr = PV[nr] (xnr-d = PVr [PV[nr-l] (Xnr -2)]


= PVrPVr - 1 [PV[nr-2] (Xnr -3)] = .. ,
= PVrPVr- 1 ••• PV, (xnr- r) = PVrPVr - 1 ••• PV, (X(n--i)r)
= (PVrPVr_1 ••• PV, ?(X(n-2)r) = ...
= (PV,PVr_l ... PV,)n(x).

Since {xn} converges to Pn,v;(x), so does the subsequence {x nr }. Thus (9.27.1)


follows. •
The subspace case of this theorem is important enough to be stated as a separate
corollary.

9.28 Corollary (Halperin). Let M i , M 2 , ... , Mr be closed subspaces in the


Hilbert space X and M = n1Mi . Then

(9.28.1)

for each x E X.
Note that this corollary generalizes Theorem 9.3 to any finite number of sub-
spaces, not just two.

Rate of Convergence for Alternating Projections


Now we turn to the question of the rate of convergence of the method of alter-
nating projections. If M i , M 2 , ... ,Mr are r closed subspaces in the Hilbert space
X and M = nrMi , then Corollary 9.28 states that

for every x EX. Since

(9.28.2)

for all x, where

(9.28.3)

it follows that Br(n) is the sharpest (Le., smallest) bound available in (9.28.2) that
is independent of x.
We will compute B2(n) exactly and give upper bounds on BrCn) when r > 2.
To do this, we first note that there are alternative ways of describing Br(n).
218 THE METHOD OF ALTERNATING PROJECTIONS

9.29 Lemma. Let M l , M 2 , ... , Mr be closed subspaces of the Hilbert space X


and M = nr
Mi. Then:
(1) PM~ is idempotent and self-adjoint.
(2) PM~ commutes with each PMi and with PMrPMr _1 ·· ·PM1 .
(3) For i = 1,2, ... , r, PMiX = x for each x E M and P Mi (Ml.) C Ml.. Hence

and

Proof. Let Pi = PMi .


(1) This follows from Theorem 5.13.
(2) Since M C Mi for each i, it follows from Theorem 9.2 that PM commutes
with Pi. Moreover, using Theorem 5.8, we obtain

and thus PM~ commutes with each Pi. It follows that PM~ commutes with
PrPr - l ·· ·Pl ·
(3) If x E M, then x E Mi for all i, so that PiX = x. The rest of the proof
follows from (2) and Theorem 9.2. •
9.30 Lemma. Let M l , M 2 , . .. , Mr be closed subspaces in the Hilbert space X
and M = nr
Mi. Then for each n E N,

(PMrPMr _1 ... PM1)n - PM = (PMrPMr _1 ... PM1)n PM~


(9.30.1) = (PMrPMr-l·· ·PMIPM~)n = (QrQr-l·· ·Ql)n,

where Qj = PMjnM~ (j = 1,2, ... , r).


In particular,

II(PMrPMr-l .. ·PM1t - Pn,Mill = II(PMrPMr-l ... PMJnpM~11


(9.30.2) = II (PMrPMr_l ... PMIPM~nl
= II(QrQr-l··· Qr)nll = II(QlQ2·· ·Qr)nll·

Proof. Let Pi = P Mi and P = PrPr - l ··· Pl· Since the adjoint of (Qr··· Ql)n
is (Ql·· ·Qr)n, the last equality in (9.30.2) follows from Theorem 8.25 (3). From
(8.25.2), it suffices to prove (9.30.1). Using Lemma 9.29 and Theorem 5.14, we
obtain PM = pn PM and
pn _ PM = pn - pnpM = pn(I - PM) = pnpM~

= pn P~~ (since PM~ is idempotent)


= (pPM~)n (since P commutes with PM~)

= [(PrPM~)(Pr-lPM~)··· (PlPM~)ln,

since Pi commutes with PM~ and PM~ is idempotent. Finally, Lemma 9.2 implies
that PiPM~ = PMinM~ = Qi (i = 1,2, ... , r) and this completes the proof. •
Now we consider the special case where r = 2.
RATE OF CONVERGENCE FOR ALTERNATING PROJECTIONS 219

9.31 Theorem. For each n EN,

(9.31.1)

Proof. From Lemma 9.30, we have

(9.31.2)

where Qi = PMi n(M1 nM2)"-- Since the adjoint of (Q1Q2)n is (Q2Ql)n, it follows
from Theorem 8.25 that

(9.31.3)

Further, by a simple inductive argument (using the fact that Qi is idempotent), we


obtain

(9.31.4)

Since Ql Q2Ql is self-adjoint, it follows from (9.31.4) and Lemma 8.32 that

(9.31.5)

Using Theorem 8.25 and Lemma 9.5(7), we deduce

IIQ1Q2Q111 = II(Q1Q2)(Q2Ql)1I = II(Q1Q2)(Q1Q2)*1i


= IIQ1Q2112 = IIPM,n(MlnM2)-,-PM2n(M,nM2)-,-1I2
= C(Ml,M2)2.

Thus (9.31.5) implies that II (Q1Q2)nIl2 = c(M1,M2)2(2n-l), and substituting this


into (9.31.2) yields the result. •

9.32 Remarks. (1) Proposition 9.31 shows that the bound in Theorem 9.8 is
sharp.
(2) A crude bound on Br(n) can be immediately obtained from (9.30.2):

nr
for all n, where M := Mi.
(3) In particular, for r = 2, we obtain from (9.30.2) the inequality

Thus the extra care taken in the proof of Proposition 9.31 resulted in an improve-
ment of this cruder bound by the multiplicative factor c (M1, M 2)n-l.
For r > 2, we do not compute Br(n) exactly, but we are able to get an upper
bound for it.
220 THE METHOD OF ALTERNATING PROJECTIONS

9.33 Theorem. Let M l , M 2, ... , Mr be closed subspaces in the Hilbert space


X, let M = nIMi , and let

(9.33.1) (i=1,2, ... ,r-1).

Then for each x E X,

(9.33.2)

where

IT (1- en
r-l ] 1/2
(9.33.3) c:= [ 1 -
<=1

Proof. Let Pi = PM, and P = PrPr - l ... PI, and write X = M EEl M.l.. We first
observe that from Lemma 9.29(3),

(9.33.4) Px = x for all x E M,

and

(9.33.5)

We will next show that for every x E M.l.,

(9.33.6)

We proceed by induction on r. For r = 1, M = M l , P = PI = PM, and hence the


left side of (9.33.6) is O.
Now assume that the result holds for r - 1 subspaces, where r :::0: 2. Let

M' = M2 n M3 n ... n Mr and p' = PrPr - l ... P2.

Then by Lemma 9.29(3), we have

(9.33.7) P'(x') = x' for every x' EM',

and

(9.33.8)

For any x E M.l., write x = Xl + x', where Xl E Ml and x' E Mt- C M.l.. Since
x' E M.l., we see that
Xl = X - x' E M.l. n MI.
Next write Xl = x~ + x~, where x~ E M' and x~ E M' .l.. Since x~ E M.l., we have

x~ = Xl - x~ E M.l. n M'.
RATE OF CONVERGENCE FOR ALTERNATING PROJECTIONS 221

Using (9.33.7), we get

P'(XI) = P'(X~ + xn = X~ + P'(x~).


Since p'(xn E Mt.l.. by (9.33.8), we obtain

(9.33.9)

The induction hypothesis yields

Substituting this and the equation Ilx~112 = IIxll12 - Ilx~ 112 into (9.33.9), we get

IIP'(XI)11 2 ~ Ilx~112 + [1-[1(1- C T)] (11xIil 2 _llx~112)


= [[1(l-C n] Ilx~112+ [l-[1(1-CT)] Ilx1112.
Now,
Ilx~ 112 = (X~, X~) = (Xl - X~, X~) = (Xl, X~)
~ C(MI,M')llxlllllx~11 = cIilxIilllx~11
implies Ilx~ II ~ clllxlll· Hence, from above,

liP' (Xl) 112 ~ [[1(1 - cD] cf IIxll12 + [1 -[1 (1- cn] IIxIil 2

= [l-(l-Ci )[1(1-c n] IIxIil 2


= [1- g(l-C n] Ilx1112=C211x1112
= c2(llxl12 _ Ilx~ 112) ~ c211x11 2 .

Since X = Xl + x~ with Xl E MI and x~ E Mr, we obtain Px~ = P r ... P2PIX~ = 0


and hence
P(X) = PXI = Pr '" P2PIXI = Pr '" P2XI = P'(XI).
This proves that

which completes the induction and proves (9.33.6). In particular, this proves

(9.33.10) IIPII ~ c.

Further, we have
222 THE METHOD OF ALTERNATING PROJECTIONS

Hence, for any x E X, we have that PM.LX E M.1. and pn-1(PM .Lx) E M.1. by
(9.33.5). Thus, by (9.33.6),

Ilpnx - PMxl1 = IlPn PM.Lxll = liP p n- 1(PM.LX)11


S; cllp n- 1(PM.LX)11 S; c2 Ilp n- 2 (PM .Lx) II
s; ... ::; en IIPM.L xii ::; cnllxll·
This completes the proof of (9.33.2). •
A more general result than Theorem 9.33 can be deduced from it. Let V1 , ... , Vr
denote r closed affine sets in X such that n1v; =1= 0. Then (see, e.g., Exercise 13 at
the end of Chapter 2) there exist unique closed subspaces M, M 1 , •.. , Mr such that

(i=I,2, ... ,r),

and
nV; =
r
M +vo
1

for any Vo E n1v;.


9.34 Corollary. For each x in the Hilbert space X,

(9.34.1)

(n = 1,2, ... ), where

r-1 ] 1/2
(9.34.2) c= [
1- II(1-cD
.=1

and

(9.34.3) n
Ci=C(M., j=i+1 M.)
J
(i=I,2, ... ,r-l).

The proof follows from the fact that PViX = PM,(X - vo) + Vo (see Theorem 2.7
(ii) and Theorem 9.33). We leave the details as an exercise.
Before we turn to some applications of these theorems, we need a deeper prop-
erty of the angle between subspaces. It characterizes when the angle between two
subspaces is positive. We shall prove it now.
9.35 Theorem. Let M and N be closed subspaces of the Hilbert space X.
Then the following statements are equivalent.
(1) c(M,N) < 1;
(2) M n (M n N).1. + N n (M n N).1. is closed;
(3) M + N is closed.
(4) M.1. + N.1. is closed.
Proof. For notational simplicity, let Y := Mn(MnN).1. and Z := Nn(MnN).1..
We first verify the following two subspace identities.
Claim. (a) Y + Z = (M + N) n (M n N).1..
RATE OF CONVERGENCE FOR ALTERNATING PROJECTIONS 223

(b) M + N = Y + Z + M n N.
To verify (a), first note that Y + Z c (M + N) n (M n N)J.. is obvious. Next
fix any x E (M + N) n (M n N)l.. Then x = Y + z, where Y E M, zEN. Since
x E (M n N)l., it follows that PMnNX = 0: Thus

x = x - PMnNX = (y - PMnNY) + (z - PMnNZ)


EM n (M n N)l. +N n (M n N)l. = Y+Z

using Theorem 4.9. This proves (a).


To prove (b), first note that Y + Z + M nNe M + N is obvious. Next fix any
x E M + N. Then x = Y + Z for some Y E M and zEN. By Theorem 5.8(2),

But since M nNe M, P MnN commutes with PM by Lemma 9.2. It follows


that P(MnN)l. = I - P MnN also commutes with PM. By Lemma 9.2 again,
P(MnN)l. (y) EM n (M n N)l. = Y. Similarly, P(MnN)l. (z) EN n (M n N)l. = Z.
This proves that x E M n N + Y + Z and verifies (b).
(2) ~ (3). If M + N is closed, then since (M n N)l. is always closed, Y + Z
is closed by Claim (a).
Conversely, suppose Y + Z is closed. Since Y + Z c (M n N)l., the sum
(Y + Z) + M n N is closed because it is a sum of orthogonal closed subspaces (see
Exercise 14 at the end of the chapter). Using Claim (b), it follows that M + N is
closed.
(1) =? (2). Let c = c (M, N) < 1. We must show that Y + Z is closed. By
Lemma 9.5(6),

(9.35.1) l(y,Z)I:<:; cilylllizil


for all Y E Y and Z E Z. Now let Xn E Y + Z and Xn -+ x. It suffices to show that
x E Y + Z. Write Xn = Yn + Zn, where Yn E Y, Zn E Z. Then {xn} is bounded, so
there exists a constant p > 0 such that

p ~ IIxnl12 = llYn + znl1 2 = IIYnl1 2 + 2(Yn, zn) + I znl1 2


~ IIYnl1 2 - 21(Yn,zn)1 + II znl1 2
~ IIYnl1 2- 2cllYnllllznli + IIznl12 (using (9.35.1))
= (1IYnll- Ilznll? + 2(1- c)IIYnllllznll·
Since 0:<:; c < 1, it follows that both sequences {IIYnll-llznll} and {IIYnllllznll} are
bounded. From this it follows that both sequences {IIYnll} and {llznll} are bounded.
By passing to a subsequence, if necessary, Theorem 9.12 implies that {Yn} converges
weakly to some y, and {zn} converges weakly to some z. Since Y and Z are closed
subspaces, they are weakly closed by Theorem 9.16(2). Thus Y E Y, Z E Z and
Xn = Yn + Zn -+ Y + Z weakly. But Xn -+ x implies Xn -+ x weakly, and hence
x = Y + Z E Y + Z. This proves (2).
(2) =? (1). Suppose that Y + Z is closed. Then H ;= Y + Z is a Hilbert space
and Y n Z = {O}. By Propo..'lition 8.24, the linear mapping Q ; H -+ Y defined by

Q(y + z) = Y , Y E Y, Z E Z,
224 THE METHOD OF ALTERNATING PROJECTIONS

is continuous.
If (1) failed, then c := c (M, N) = 1, and there would exist Yn E Y, Zn E Z with
llYn I = 1 = Ilzn II and 1 = c = lim (Yn, zn). Then
n--+oo

implies that
Yn = Q(Yn - zn) --+ Q(O) = 0,
which is absurd. Thus c < 1 and this proves (1).
We have shown the equivalence of the first three statements. But Theorem 9.7
already established the equivalence of (3) and (4). •
Remark. It is also true that c(M·l..,Nl.) = c(M,N). However, since an ele-
mentary proof of this fact seems fairly lengthy, and since we use this fact in only
one example below (Example 9.40), we have omitted it.
9.36 Lemma. Let M be a closed subspace and N a finite-dimensional subspace
in the inner product space X. Then M + N is a closed subspace.
Proof. The proof is by induction on k = dimN. For k = 1, N = span{b1 } for
some bi EX \ {O}. If bi EM, then M + N = M is closed. Thus we may assume
bi tt M. Let Xn E M + Nand Xn --+ x. We must show that x E M + N. Then
Xn = Yn + anb i for some Yn EM and an E R. We have

so that

which implies that {an} is bounded. Thus there is a subsequence {anJ and a E R
such that an; --+ a. Hence

since l'vI is closed, and x = Y + ab i E M + N.


Now assume that the result is true for all subspaces of dimension at most k. If
N has dimension k + 1, then we can write N = Nk + N I , where dim Nk = k and
dim NI = 1. By the induction hypothesis, Mk := M + Nk is closed. By the proof
when k = 1, Mk + NI is closed. Since M + N = M + Nk + NI = Mk + N I , it follows
that M + N is closed. •
9.37 Corollary. Let M and N be closed subspaces in the Hilbert space X such
that M or N has either finite dimension or finite codimension. Then c (M, N) < 1.
Proof. If dim N < 00, then M + N is closed by Lemma 9.36. If codimN < 00,
then dimNl. < 00 by Theorem 7.12, and MJ. +Nl. is closed by Lemma 9.36. Thus
either M + Nor MJ. + Nl. is closed. By Theorem 9.35, c (M, N) < 1. •

Before turning to some possible applications of the von Neumann and Dykstra
algorithms, we will state two lemmas that will prove useful in these applications.
RATE OF CONVERGENCE FOR ALTERNATING PROJECTIONS 225

9.38 Lemma. Let M be a closed subspace of the Hilbert space X and


x E X \ {OJ. Then

(9.38.1) c(span{x},M) = {oIIPM(X)ll/llx ll if x M, tt


if x E M.
In particular, c (span {x}, M) < l.
Proof. If x tt M, span{x} n M = {OJ, so that
c(M,span{x}) = sup{l(y,ax)11 y E M, Ilyll::; 1, a E JR, lal::; Ilxll- 1 }

= w sup{l(y,x)11 y E M,
1
Ilyll ::; I}

= w sup{I(PMz,x)11 z
1
E X, Ilzll ::; I}

= w PM x) I z
1
sup{l(z, I E X, Ilzll::; I}
1
=WIIPMxll.

If x E M, then span{ x} C M and Lemma 9.5 (8) implies that c (span{ x}, M) =
O. This proves (9.38.1).
tt
If x M, then since x = PM(x) + PMJ.. (x), we have

IIxl1 2 = IIPM(x)11 2 + IIPMJ..(x)112 > IIPM(x)11 2,


so that IIPM(X)ll/llxll < 1. This proves the last statement. •
It is convenient to restate here a special consequence of Theorem 6.31 concerning
best approximations from half-spaces.
9.39 Theorem (Best approximation from half-spaces). Let X be a Hilbert
space, z EX \ {OJ, c E JR, and
(9.39.1) H = {y E X I (y,z) ::; c}.
Then:
(1) H is a Chebyshev half-space.
(2) For any x E X,

(9.39.2) ( ) _ _ [(x,z)-c]+z
PH X - X II z l12

and

(9.39.3) d( H) = [(x, z) - c]+


x, Ilzll'
where
if a ~ 0,
if a < O.
What formula (9.39.2) implies, in particular, is that it is easy to compute best
approximations from the closed half-spaces (9.39.1), and hence Dykstra's algorithm
is effective for the intersection of a finite number of closed half-spaces (i.e., for a
polyhedron) in a Hilbert space.
226 THE METHOD OF ALTERNATING PROJECTIONS

Examples
We now list several examples, which can all be effectively handled by Dykstra's
algorithm or von Neumann's alternating projections algorithm.
9.40 Example (Linear equations).
Let X be a Hilbert space, {Yl, Y2, .. . , Yr} in X \ {O}, and {Cl, C2, .. . , cr } C R
We want to find an x E X that satisfies the equations

(9.40.1) (x, Yi) = C; (i=I,2, ... ,r).

Setting
(i=I,2, ... ,r)
and
r
(9.40.2) K=nHi,

we see that x satisfies (9.40.1) if and only if x E K. Assuming K =1= 0, we see that
since the Hi are closed hyperplanes and, by Theorem 6.17,

1
PHi(Z)=z-IIYiI12[(z,Yi)-Ci]Yi

for any z E X. It follows from Corollary 9.34 that (PHrPHr _1 ••• PH, )n(x) converges
to PK(X), for any x E X, at the rate

(9.40.3)

where Va E K is arbitrary,

r-1 ] 1/2
(9.40.4) C= [ 1- }J(I- cD ,

(9.40.5) (i=I,2, ... ,r-l),

and

(9.40.6) Mi = Hi - va = {y E X I (y, Yi) = O}.

But using Theorem 4.6(5) and the remark preceding Lemma 9.36, we obtain

Ci = C(M/-, C~lMj).l) = c(M/-,Mi1.1+ ... + Mf)


=c(M/-,Mft1 +... +M;-),
EXAMPLES 227

since each Mf is finite-dimensional so that the sum Ml


t-l + ... + M;!- is closed
(Lemma 9.36). Since M j = yt
= [span{Yj }]J., it follows by Theorem 4.5(S) that

(9.40.7)

Thus by Lemma 9.3S,

Ci = C(span{Yd, span{Yi+1' Yi+2, ... , Yr})


0 if Yi E span{Yi+l,.··, Yr},
{
= IlPs p an{Yi+l ,... ,Yr} (Yi) III IIYi II otherwise,

and Ci < 1 for all i. This shows that c < 1.


In particular, if X = £2(N) and Yi = (ail, ai2, ... , aiN) (i = 1,2, ... , r), equations
(9.40.1) can be rewritten as

N
(9.40.S) I>ijX(j) = Ci (i=1,2, ... ,r).
j=l

Thus we can use von Neumann's alternating projections algorithm to compute a


particular solution to any consistent system of linear equations (9.40.S). In par-
ticular, starting with Xo = 0, the von Neumann algorithm produces a sequence
of points {xn} in £2(N) that converges geometrically to the unique minimal norm
solution of (9.40.S).

9.41 Exrunple (Linear inequalities).


Let X be a Hilbert space, {Yl, Y2, ... , Yr} C X \ {O}, and {Cl, C2, . .. , cr } C R.
We want to find an x E X that satisfies the inequalities

(9.41.1) (x, Yi) :::: Ci (i=1,2, ... ,r).

Setting K = n1Hi, where

(9.41.2) Hi = {y E X I (Y,Yi):::: cd,


we see that x solves (9.41.1) if and only if x E K. We assume K i= 0.
To obtain a point in K, we start with any Xo E X and generate the sequence
{xn} according to Dykstra's algorithm that converges to PK(xo). This algorithm
is effective, since it is easy to obtain best approximations from the half-spaces Hi
by Lemma 9.39.
In particular, taking X = £2 (N) and Yi = (ail, ai2, ... , aiN), the inequalities
(9.41.1) become

(9.41.3) (i=1,2, ... ,r).

Thus Dykstra's algorithm can be used effectively to solve a system of linear in-
equalities.
228 THE METHOD OF ALTERNATING PROJECTIONS

9.42 Example (Isotone regression). An important problem in statistical


inference is to find the isotone regression for a given function defined on a finite
ordered set of points. In the language of best approximation, a typical such problem
may be rephrased as follows. Given any x E £2(N), find the best approximation to
x from the set of all increasing functions in £2(N):

(9.42.1 ) C = {y E £2(N) I y(l) ::::: y(2) ::::: ... ::::: yeN)}.

Note that we can write C as


N-1
(9.42.2) C= n Hi,
i=l
where

Hi = {y E £2(N) I y(i) ::::: y(i + I)} = {y E £2(N) I (y, Yil ::::: O}


is a closed half-space, and

if j = i,
if j = i + 1,
otherwise.

Thus we are seeking Pc (x). Since C is the intersection of a finite number of half-
spaces, this reduces to an application of Dykstra's algorithm just as in Example
9.41 beginning with Xo = x.
9.43 Example (Convex regression). Another problem in statistical inference
is to find the convex regression for a given function defined on a finite subset of the
real line. In the language of best approximation, the standard such problem can be
rephrased as follows. Given any x E £2(N), find the best approximation to x from
the set of all functions in £2(N) that are convex, that is, from the set
N-2 N-2
(9.43.1 ) C = i~:\ {y E £2(N) I y(i + 1) - y(i) ::::: y(i + 2) - y(i + I)} = i;;l Hi,
where
Hi = {y E £2(N) I y(i + 1) - y(i) ::::: y(i + 2) - y(i + I)}.
But

Hi = {y E £2(N) I- y(i) + 2y(i + 1) - y(i + 2) ::::: O} = {y E MN) I (y, Yil ::::: O}

is a closed half-space, where

-I ifjE{i,i+2},
Yi(j) =
{
~ if j = i + 1,
otherwise.

Since we are seeking Pc(x) , and C is the intersection of a finite number of


closed half-spaces, we again can apply Dykstra's algorithm (just as in Example
9.41) starting with Xo = x.
EXAMPLES 229

9.44 Example (Shape-Preserving Interpolation). In the next chapter we


will see that a large class of problems, which fall under the general heading of
"shape-preserving interpolation," can be put into the following form.
Let X = L 2 [0, 1], {YllY2, ... ,Yr} eX \ {O}, and {Cl,C2,'" ,er} c R. Assume
that the set

(9.44.1 ) K = {y E X IY ? 0, (y, Yi) = Ci (i = (1,2, ... , r)}

is nonempty. We want to find the element of K having minimal norm; that is, we
seek PK(O).
Setting
(i=1,2, ... ,r)

and
Kr+l = {y E X IY ? O},
we see that Kr+l is a closed convex cone, K 1 , ••• , Kr are closed hyperplanes, and

(9.44.2)

Since best approximations from K r + 1 are given by the formula (Theorem 4.8)

and best approximations from the Ki (i = 1,2, ... , r) are given by (Theorem 6.17)

we see that this problem can be effectively handled by Dykstra's algorithm starting
with Xo = 0. In fact, since each of the Ki for i = 1,2, ... , r is an affine set, we
can ignore all the residuals when projecting onto K 1 , K 2 , ... ,Kr (see (9.26.3)). We
need only keep track of the residuals obtained when projecting onto K r +1 • This is
a substantial simplification of Dykstra's algorithm for this case.
More precisely, set

[n]= {1,2, ... ,r+ I} n {n - k(r+ 1) I k = 0,1,2,3 ... },


Xo = 0, e- r = e_(r-l) = ... = e_l = eo = 0,
Xn = PK[n] (Xn-l) if [n] i= r + 1,
and

if [n] = r + l.
Then the sequence {xn} converges to PK(O). More generally, start with any
x E X and put Xo = x. Then the sequence {xn} converges to PK(X),
230 THE METHOD OF ALTERNATING PROJECTIONS

Exercises

1. Let {x n } be the sequence defined in the von Neumann theorem (Theorem


9.3). Verify the following statements:
(a) IIX2nli ~ IIX2n-ili for all nand IIx2nli = IIx2n-ril if and only if X2n-i E
M2·
(b) IIX2n+11l ~ IIx2nli for all nand IIX2nHII = IIx2nli if and only if X2n E Mi·
(c) If x2n E Mi for some n, then Xk = X2n for all k 2 2n and X2n =
PM1 nM2(X).
(d) If X2n-i E M2 for some n, then Xk = X2n-i for all k 2 2n - 1 and
X2n-i = PMlnM2(X).
(e) (X2n+i,X2mH) = (X2n,X2m+2) for all n,m E No
2. Verify Example 9.6 (2). That is, show that the angle between the subspaces
M = {x E £2(3) I x(3) = O} and N = {x E £2(3) I x(l) = O} is 7r/2
(radians).
3. What is the angle between the following pairs of subspaces?
(a) M and M.L (M an arbitrary subspace),
I
(b) M = {x E £2 x(2n) = 0 (n = 1,2, ... n
and
I
N = {x E £2 x(2n -1) = 0 (n = 1,2, ... n·
4. Some authors have defined the "minimal angle" between subspaces M and
N as the angle between 0 and 7r/2 whose "cosine" is

eo(M,N):= sup{l(x,y)11 x E M, Ilxll ~ 1, yEN, lIyll::; I}.


Verify the following statements.
(a) 0 ~ eo(M, N) ~ l.
(b) eo(M, N) = eo(N, M).
(c) c(M,N) ~ co(M,N) and c(M,N) = co(M,N) if MnN = {O}.
(d) eo(M, N) < 1 if and only if M n N = {O} and M + N is closed.
[Hint: co(M, N) = c (M, N) in the cases ofrelevance.]
(e) If M + N is closed and M n N f= {O}, then c (M, N) < caCM, N) = l.
5. Verify the following statements.
(a) Weak limits are unique; that is, if Xn ~ x and Xn ~ y, then x = y.
(b) If Xn ~ x and Yn ~ y, then Xn + Yn ~ X + y.
(c) If Xn ~ x and an -----+ a (in 1R), then anxn ~ ax.
(d) Xn ~ x weakly if and only if x*(xn) -----+ x*(x) for all x* E X*.
[Hint: Theorem 6.8.]
6. In a Hilbert space, show that a sequence {xn} is bounded if and only if each
subsequence of {xn} has a subsequence that converges weakly.
[Hint: Theorem 9.12 and Lemma 9.13.]
7. Let X be a Hilbert space and {xn} a sequence in X. Show that the following
statements are equivalent.
(a) {xn} converges weakly to x.
(b) Every subsequence of {xn} has a subsequence that converges weakly to
x. [Hint: Lemma 9.13 and Exercise 6.]
8. Call a set K weakly compact if each sequence in K has a subsequence
that converges weakly to a point in K. Prove the following statements.
(a) Every weakly compact set is weakly closed.
EXERCISES 231

(b) In a Hilbert space, a set is weakly compact if and only if it is weakly


closed and bounded.
[Hint: Exercise 6.]
9. Let M be a Chebyshev subspace in X, v E X, and V = M + v. Prove that
V is an affine Chebyshev set, and for all x, Y E X and a E JR,

Pv(x + y) = Pv(x) + PM(Y)


and
Pv(ax) = aPM(x) + Pv(O).
10. In the text, Corollary 9.28 was deduced from Theorem 9.27. Assuming
Corollary 9.28, deduce Theorem 9.27.
[Hint: Translate niVi by an element in this set.]
11. In the proof of Theorem 9.31, we used the fact that if Q1 and Q2 are
idempotent, then

(n = 1,2, ... ).
Prove this by induction on n.
12. Let A,B E 8(X,X). Show that if A and B commute, then

A(R(B)) c R(B) and A(N(B)) c N(B).

In words, the range and null space of anyone of two commuting operators
are invariant under the other operator.
13. Prove Corollary 9.34.
[Hint: Use Theorem 2.7 (ii), Exercise 13 at the end of Chapter 2, and
induction to deduce that

and
Pnrv,(X) = PnrMi(X)(X - vo) + Vo·
Then apply Theorem 9.33.]
14. (Sum of orthogonal subspaces is closed) Suppose that M and N are
closed orthogonal subspaces of the Hilbert space X, Le., M C NJ.. Show
that M + N is closed. More generally, M + N is closed if M and N are
orthogonal Chebyshev subspaces in the inner product space X.
15. Suppose X and Yare Hilbert spaces and A E 8(X, Y). Show that A is
bounded below if and only if A is one-to-one and has closed range. [Hint:
Lemma 8.18 for the "if" part, and Theorem 9.12 and Lemma 9.14 for the
"only if" part.]
16. An operator A E 8(X, X) is called nonnegative, denoted by A 2': 0, if
(Ax, x) 2': 0 for all x E X. If A, B are in 8(X, X), we write A 2': B or
B ::; A if A - B 2': o.
If M and N are Chebyshev subspaces in X, show that the following three
statements are equivalent:
(a) PM::; PN ;
232 THE METHOD OF ALTERNATING PROJECTIONS

(b) MeN;
(c) IIPMxl1 :::; IIPNXII for all x E X.
17. Let M and N be closed subspaces of the Hilbert space X and A E B(X,X).
Show that
IIPN PM AI1 2 :::; c211PMAI12 + (1- c2)IIPMnNAI12,
where c = c (M, N).
18. Let M and N be closed subspaces of the Hilbert space X. Show that
(a) M = M n N + M n (M n N)J.,
(b) PM = PMnN + PMn(MnN)~·
[Hint: For (a), let x E M, y = PMnNX, and z = x - y. Note that z E
Mn (MnN)J. and x E M nN +Mn (MnN)J.. For (b), use (a) and show
that x - (PMnNX + PMn(MnN)~X) E MJ. for any x.]
19. For closed subspaces M and N in a Hilbert space X, prove that the following
five statements are equivalent.
(a) PM and PN commute;
(b) PM and PN~ commute;
(c) PM~ and PN commute;
(d) PM~ and PN~ commute;
(e) M=MnN+MnNJ..
20. Let M 1 , M 2 , and M3 be closed subspaces in the Hilbert space X with
M2 U M3 C Ml and M2 J.- M 3 · Suppose PM := PM! - P M2 - P M3 is an
orthogonal projection. Show that M = {O} if and only if Ml = M2 + M 3 .
21. Let M and N denote the following subspaces of X = £2(4):
M = span{el, e2 + ed and N = span{el + e2, e3 + e4},
where ei are the canonical unit basis vectors for X: ei (j) = (iij. It was
shown in Chapter 4 (Exercise 26 at the end of Chapter 4) that M and N
are 2-dimensional Chebyshev subspaces in X with M n N = {O}, and for
every x = z=i
x(i)ei,

and

(a) Compute c (M, N).


(b) Compute PNPM(X) for any x E X.
(c) Compute (PNPM)n(X) for any x E X and every n E N.
(e) Give a direct proof that
(PNPM )n(x) -+ 0 as n -+ 00

for every x EX. That is, do not appeal to Theorem 9.3.


22. (Opial's condition) Let X be an inner product space and let {x n } be a
sequence in X that converges weakly to x. Show that for each y E X \ {x},
liminf Ilx n
n
- yll > liminf
n
Ilx n - xii·

[Hint: Expand Ilx n - Yl12 = II(x n - x) + (x - y)112.]


HISTORICAL NOTES 233

Historical Notes
From Lemma 9.2, we see that P M2 PM, = P M ,nM2 if PM, and P M2 commute,
that is, the product of two commuting projections is a projection. Von Neumann
(1933) was interested in what could be said in the case where PM, and PM2 do not
commute. He subsequently proved Theorem 9.3, which states that the sequence of
higher iterates of the product of the two projections {(PM2 PM,)n I n = 1,2, ... }
converges pointwise to P M ,nM2. This theorem was rediscovered by several authors
including Aronszajn (1950), Nakano (1953), Wiener (1955), Powell (1970), Gor-
don, Bender, and Herman (1970), and Hounsfield (1973)~the Nobel Prize winning
inventor of the EMI scanner.
There are at least ten different areas of mathematics in which the method of al-
ternating projections has found applications. They include solving linear equations,
the Dirichlet problem, probability and statistics, computing Bergman kernels, ap-
proximating multivariate functions by sums of univariate ones, least change secant
updates, multigrid methods, conformal mapping, image restoration, and computed
tomography (see the survey by Deutsch (1992) for a more detailed description of
these areas of applications and a list of references).
There are several posssible ways of extending the von Neumann theorem (Theo-
rem 9.3). In a strictly convex reflexive Banach space X, every closed convex subset
C is a Chebyshev set (see Riesz (1934), who proved it in case X is a Hilbert space,
and Day (1941; p. 316) for the general case) so Pc is well-defined. Stiles (1965a)
proved the following result.
Theorem (Stiles). If X is a strictly convex reflexive Banach space having
dimension at least 3, and if for each pair of closed subspaces M and N in X and
eachxEX,

then X must be a Hilbert space.


Thus trying to extend Theorem 9.3 to more general Banach spaces seems fruit-
less. However, by replacing the subspaces by their orthogonal complements in
Theorem 9.3, and using the relationship between their orthogonal projections (see
Theorem 5.8(2)), we obtain the following equivalent version of Theorem 9.3.
Theorem. Let X be a Hilbert space and let M and N be closed subspaces.
Then, for each x EX,
(9.44.3)

Interestingly enough, in contrast to the equivalent Theorem 9.3, this theorem


does extend to more general Banach spaces. Indeed, Klee (1963) exhibited a 2-
dimensional non-Hilbert space X in which (9.44.3) holds for all closed subspaces
M and N. (Klee's result answered negatively a question of Hirschfeld (1958), who
asked, If X is strictly convex and reflexive and (9.44.3) holds for all closed subspaces
M and N, must X be a Hilbert space?) Extending this even further, Stiles (1965c)
showed that (9.44.3) holds if X is any finite-dimensional smooth and strictly convex
space. For a variant of the Stiles (1965c) theorem when strict convexity is dropped,
see Pantelidis (1968). Franchetti (1973), Atlestam and Sullivan (1976), and Deutsch
(1979) have each generalized the Stiles (1965c) theorem. The most inclusive of these
is in the latter paper, and it may be stated as follows.
234 THE METHOD OF ALTERNATING PROJECTIONS

Theorem (Deutsch). Let X be a uniformly smooth and uniformly convex


Banach space (i.e., both X and X* are uniformly convex). Let M and N be closed
subspaces of X such that M + N is closed. Then

(9.44.4) lim [(1 - PN )(1 - PM)t (x) = (1 - PM +N )(x) for each x E X.


n->oo

Boznay (1986) stated that this theorem holds without the restriction that M +N
be closed, but his proof is not convincing. Franchetti and Light (1984) proved
that (9.44.4) holds for all pairs of closed subspaces M and N if, in addition, the
modulus of convexity Ox of X (respectively ox< of X*) satisfies inf,>o C20X(E) >
o (respectively inf,>o E-20x< (E) > 0). They also conjecture that there exists a
uniformly convex and uniformly smooth Banach space X such that (9.44.4) fails
for some pair of closed subspaces M and N (whose sum is not closed). Finally, they
proved that in the Deutsch theorem, the condition that X be uniformly convex could
be weakened to X is an "E-space." (An E-space is a strictly convex Banach space
such that every weakly closed set is approximatively compact.)
If A and B are closed convex subsets of a Hilbert space X, Bauschke and Borwein
(1993) made a thorough study of the convergence of the iterates {(PBPA)n(X)},
even when A and B have empty intersection.
The angle between two subpaces given in Definition 9.4 is due to Friedrichs
(1937). Dixmier (1949) gave a different definition (obtained by deleting the factors
(MnN)~ in the Friedrichs definition). These definitions agree when MnN = {O},
but not in general.
Lemma 9.5(7)-(8) is from Deutsch (1984). Theorem 9.7, even in more general
Banach spaces, can be found in Kato (1984; Theorem 4.8). The rate of convergence
in Theorem 9.8 is essentially due to Aronszajn (1950), while the fact that the
constant C2n - 1 is best possible is due to Kayalar and Weinert (1988).
The weak compactness theorem (Theorem 9.12) is due to Hilbert (1906) in the
e
space 2 (who called it a "principle of choice"), and to Banach (1932) in any reflexive
space.
Dykstra's algorithm was first described, and convergence proved, by Dykstra
(1983) in the case where X is Euclidean n-space and where all the closed convex
sets Ki were actually convex cones. Later, Boyle and Dykstra (1985) proved their
general convergence theorem (Theorem 9.24) by essentially the same proof as de-
scribed in this chapter. Dykstra's algorithm is applicable for problems involving
isotone regression, convex regression, linear inequalities, quadratic programming,
and linear programming (see the exposition by Deutsch (1995) for a more detailed
description of these problems).
Corollary 9.28 was first proved by Halperin (1962) by a more direct approach.
A short proof of Halperin's theorem was given by Stiles (1965b). Halperin's proof
was adapted by Smarzewski (1996) (see also Bauschke, Deutsch, Rundal, and Park
(1999)) to yield the following generalization.
Theorem (Smarzewski). Let T1 , T 2 , ... , Tk be self-adjoint, nonnegative, and
nonexpansive bounded linear operators on the Hilbert space X. Let T = Tl T2 ... Tk
and M = FixT. Then

lim Tn(x) = PM (x) for each x E X.


n->oo
HISTORICAL NOTES 235

Aronszajn (1950) proved the inequality

(9.46.3)

and Kayalar and Weinert (1988) sharpened this by showing that equality actually
holds in (9.46.3) (Le., they proved Theorem 9.31). The rate of convergence theo-
rem (Theorem 9.33) is due to Smith, Solmon, and Wagner (1977). Sharper rates
of convergence theorems were later established by Kayalar and Weinert (1988) and
Deutsch and Hundal (1997). Also, in contrast to Theorem 9.31 for two subspaces,
Deutsch and Hundal (1997) showed that none of the bounds given by Smith, Sol-
mon, and Wagner (1977), by Kayalar and Weinert (1988), and by Deutsch and
Hundal (1997) is sharp for more than two subspaces!
The equivalence of the first two statements in Theorem 9.35 is due to Deutsch
(1984), while the equivalence of the second two is due to Simonic, who privately
communicated his results to Bauschke and Borwein (1993; Lemma 4.10). Re-
lated to this and to the error bound (9.32.1), we have the following result: Let
M 1 ,M2 , •.• ,Mr be closed subspaces of a Hilbert space and M = nIMi . Then
IIPMrPMr-, ... PMl PM.L II < 1 if and only if Mf + Mt + ... + M! is closed. This
follows from a combination of results of Bauschke and Borwein (1996; Lemma 5.18
and Theorem 5.19) and Bauschke, Borwein, and Lewis (1996; Proposition 3.7.3 and
Theorem 3.7.4). (See also Bauschke (1996; Theorem 5.5.1).)
Hundal and Deutsch (1997) extended Dykstra's algorithm in two directions.
First they allowed the number of sets Ki to be infinite; second, they allowed a
random, rather than cyclic, ordering of the sets K i .
Finally, we should mention that the main practical drawback of the method of
alternating projections (MAP), at least for some applications, is that it is slowly
convergent. Both Gubin, Polyak, and Raik (1967) and Gearhart and Koshy (1989)
have considered a geometrically appealing method of accelerating the MAP that
consists in adding a line search at each step (but no proofs were given of convergence
of the acceleration scheme in either paper). Bauschke, Deutsch, Hundal, and Park
(1999) considered a general class of iterative methods, which includes the MAP
as a special case, and they studied the same kind of acceleration scheme. They
proved that the acceleration method for the MAP is actually faster in the case of
two subspaces, but not faster in general for more than two subspaces. They also
showed that the acceleration method for a symmetric version of the MAP is always
faster (for any number of subspaces). (See also the brief survey by Deutsch (2001)
of line-search methods for accelerating the MAP.) Whether a similar acceleration
scheme works for the Dykstra algorithm is an interesting open question.
CHAPTER 10

CONSTRAINED INTERPOLATION FROM A CONVEX SET

Shape-Preserving Interpolation
In many problems that arise in applications, one is given certain function values
or "data" along with some reliable evidence that the unknown function that gener-
ated the data has a certain shape. For example, the function may be nonnegative
or nondecreasing or convex. The problem is to recover the unknown function from
this information.
A natural way to approximate the unknown function is to choose a specific func-
tion from the class of functions that have the same shape and that also interpolate
the data. However, in general there may be more than one such function that inter-
polates the data and has the same shape. Thus an additional restriction is usually
imposed to guarantee uniqueness. This additional condition can often be justified
from physical considerations.
To motivate the general theory, let us consider a specific example.
10.1 Example. Fix any integers k ~ 0 and m ~ 1, and let L~k)[O, 1J denote
the space of all real-valued functions f on [O,lJ with the property that the kth
derivative f(k) of f exists and is in L 2 [0, 1J. Consider the prescribed m + k points
o S h < t2 < ... < tm+k S l.

Suppose some fa E L~k)[O, 1J satisfies f~k) ~ 0 on [O,lJ. Thus if k = 0,1, or 2, fa


would be nonnegative, nondecreasing, or convex, respectively.
Now let
G = Gk(Jo) := {g E L~k)[O.lJ I g(ti) = fo(ti) (i = 1,2, ... , m + k) and g(k) ~ O}.

That is, G is the set of all functions in L~k)[O, 1J that interpolate fa at the points ti
and that have the "same shape" as fa. Suppose all that we know about the function
fa are its values at the m + k points ti and that its kth derivative is nonnegative.
To approximate fa on the basis of this limited information, it is natural to seek an
element of G. But which element of G?
Certainly, Gis nonempty (since fa E G), and if there is more than one element
of G, then there are infinitely many, since G is a convex set. For definiteness, we
shall seek the element go E G whose kth derivative has minimal L 2 -norm:

(10.1.1) Ilg6k ) II = inf {llg(k) III g E G} .

It can be shown, by using integration by parts, that this problem is equivalent


to one of the following type. Determine the element ho of minimal norm from the
set
(10.1.2) H := {h E L 2 [0, III h ~ 0, (h, Xi) = bi (i = 1,2, ... , m)}.

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
238 CONSTRAINED INTERPOLATION FROM A CONVEX SET

That is, choose ho E H such that

(lO.1.3) Ilholl = inf {llhlll h E H}.

Here the scalars bi and the functions Xi E L2 [0, 1] (i = 1, 2, ... , m) are completely
determined by the points ti and the function values fO(ti) (i = 1,2, ... , m + k).
In fact, if ho is a solution to (lO.1.3), then the element go E L~k) [0,1] with the
gb
property that k ) = ho solves the problem (lO.1.1).
Let us rephrase the problem (10.1.3). Let {Xl,X2, ... ,X m } be a set in L2 =
L2[0,1], b = (bl, b2, ... , bm ) E £2(m), C = {x E L2 I x ~ O}, and

(lO.1.4) K={XECI(x,xi)=b i , i=1,2, ... ,m}.

Clearly, K is a closed convex subset of the Hilbert space L 2 , so that if K is


nonempty, best approximations in K to any x E L2 always exist and are unique.
Briefly, K is Chebyshev (Theorem 3.5). The problem (lO.1.3) can then be stated
as "determine the minimum norm element PK(O) in K."
If we define a linear mapping A : L2 -+ £2 (m) by

then we see that K may be rewritten in the compact form

(lO.1.5) K = {x E C I Ax = b} = C n A-l(b).
In this chapter we shall consider the following generalization of this problem.
Let X be a Hilbert space, C a closed convex subset of X, b E £2(m), A a bounded
linear operator from X to £2(m), and

K = K(b) := {x E C I Ax = b} = C n A-l(b).
Our problem can be stated as follows. Given any x E X, determine its best ap-
proximation PK(b)(X) in K(b). We shall be mainly interested in establishing a
useful characterization of best approximations from K(b). What we shall see is
that PK(b)(X) is equal to the best approximation to a perturbation x + A*y of x
from C (or from a convex extremal subset C b of C). The merit of this characteriza-
tion is threefold. First, it is generally much easier to compute best approximations
from C (or C b ) than from K(b). Second, when X is infinite-dimensional, the prob-
lem of computing PK(b) (x) is intrinsically an infinite-dimensional problem, but the
computation of the y E £2(m) for which Pc(x + A*y) E A-l(b) depends only on a
finite number mof parameters. Third, in many cases there are standard methods
for solving the latter problem.

Stong Conical Hull Intersection Property (strong CHIP)


Before proceeding with establishing the characterization of best approximations
from K(b) = C n A-l(b), it is convenient first to make a definition of a property
(the "strong CHIP") that will prove fundamental to our analysis. Loosely speaking,
when the strong CHIP is present (respectively is not present), then PK(b)(X) =
Pc(x + A*y) (respectively PK(b)(X) = PCb (x + A*y)) for some y E £2(m). Then we
give some examples, and establish some general results, concerning this property.
STONG CONICAL HULL INTERSECTION PROPERTY (STRONG CHIP) 239

10.2 Definition. A collection of closed convex sets {C 1 ,C2, ... ,Cm } in X,


which has a nonempty intersection, is said to have the strong conical hull inter-
section property, or the strong CHIP, if

(10.2.1) for each x E n


m

Ci ·

There are useful alternative ways of describing the strong CHIP. We state one
next, and include others in the exercises.

10.3 Lemma. Let {C1 , C2, . .. , Cm} be a collection of closed convex sets in a
Hilbert space X. Then the following statements are equivalent:
(1) {C 1 ,C2, ... ,Cm } has the strong CHIP;
(2) For each x E nl'ci ,

(10.3.1)

and
m
(10.3.2) is closed

(3) For each x E nl'Ci ,

(10.3.3) :) n
m

1
con (Ci - x)

and (10.3.2) holds.

Proof. (1) ===} (2). Suppose (1) holds. Then for each x E nl'Ci , (10.2.1) holds.
Taking dual cones of both sides of (10.2.1), we obtain

Using 4.5(9) and 4.6(3), we deduce

which proves (10.3.1). Also, (10.3.2) is a consequence of (10.2.1) and 4.5(1).


(2) ===} (3) is obvious.
240 CONSTRAINED INTERPOLATION FROM A CONVEX SET

(3) ==? (1). If (3) holds, then since the inclusion con (nrCi -x) c nrcon(Ci-
x) always holds, it follows that (2) holds. Taking dual cones of both sides of (10.3.1),
and using 4.5(3) and 4.6(4), we obtain

m m
= 2:: [con (Ci - xW = 2::(Ci - x)o
1 1
m

since the latter set is closed. Hence {C1 , C2 , .. . , Cm} has the strong CHIP. •
When one is approximating from an intersection of sets having the strong CHIP,
a strengthened characterization of best approximations can be given.
10.4 Theorem. Let {CI, C 2 , ... , Cm} be a collection of closed convex sets with
the strong CHIP in X, x E X, and Xo E C:= nrCi . Then Xo = Pc(x) if and only
if
m
(10.4.1) X - Xo E 2::(Ci - xot·
1

Proof. By Theorem 4.3, Xo = Pc(x) if and only if x - Xo E (C - xo)o. By the


strong CHIP, the result follows. •
The main usefulness of this result is in the case where the sets C i have the
additional property that (Ci - xo)O does not depend on the particular Xo E nrCi .
The next few examples we give are of this type.
10.5 Example of the Strong CHIP (Subspaces). Let {M1 ,M2 , ... ,Mm }
be a collection of closed subspaces in the Hilbert space X. If each of the M i , except
possibly one, has finite co dimension, then

(10.5.1)

for each x E nr Mi and so {Ml, M 2 , ... , Mm} has the strong CHIP.
In particular, if X is finite-dimensional, then every finite collection rof subspaces
in X has the strong CHIP.
Proof. Let x E nr Mi. Since Mi - x = Mi and nr Mi - x = nr M i , we obtain

m m
= 2::(Mi - x).l = 2::(Mi - x)O,
1
STONG CONICAL HULL INTERSECTION PROPERTY (STRONG CHIP) 241

where the second equality follows from Theorem 4.6(4), the third from Theorem
4.5(6), the fourth from Theorem 9.36, the fifth from Mi = Mi - x, and the seventh
from Theorem 4.5(6). •

Before giving our next example of the strong CHIP, it is convenient to restate
here Lemma 6.38, which is a result about when the conical hull operation commutes
with the operation of intersection.

10.6 Lemnm. Let {Cl , C 2 , .•. , Cm} be a collection of convex sets with aE
nrCi . Then

(10.6.1)

Remark. We note that if the conical hull "con" is replaced by the closed conical
hull "con" in Lemma 10.6, the result is no longer valid! (See Exercise 25 at the
end of Chapter 4.)

10.7 Example of the Strong CHIP (Half-Spaces). Let X be a Hilbert


space, hi E X\ {a}, ai E JR, and

(i=1,2, ... ,m).

Then, for each x E nr Hi,

(10.7.1)

and so {Hi, H 2 , .•. ,Hm} has the strong CHIP. Here lex) denotes the set of "active"
indices for x; that is, lex) := {i I (x, hi) = ad.
Proof. This result is just a restatement of Theorem 6.40. •

Other examples of the strong CHIP can be constructed from these examples by
using the next lemma.

10.8 Lemma. Let {CI ,C2 , ... ,Cm} have the strong CHIP. Let J I , ... ,In be
nonempty disjoint sets of indices whose union is {I, 2, ... , m}. Then

has the strong CHIP.


Proof. Let Aj := niEJjCi (j = 1,2, ... , n). To show that {AI, A 2 , . .. , An} has
242 CONSTRAINED INTERPOLATION FROM A CONVEX SET

r (0 r
the strong CHIP. Let x E nf Aj . Then x E nrCi and

(0 Aj - x = Ci - x = ~(Ci - x t
= 'L)Ci - x)O + ... + 'L)Ci - xt

(by Theorem 4.6(2))

n
= (AI - xt + ... + (An - xt = ·~.)Aj - xt
I

(by Theorem 4.6(2)).

It follows that the two subset symbols "e" may be replaced by equality symbols

-xr
"=", and hence

(0 Aj = ~(Aj -xt·
This proves that {AI, A 2, ... , An} has the strong CHIP. •
10.9 Example of the Strong CHIP (Hyperplanes). Let X be a Hilbert
space, hi E X\{O}, ai E JR, and
H i := {x E X [(x,h i ) = ad (i = 1,2, ... ,m).
Then, for each x E nr Hi,
(10.9.1) (0 Hi - x) ° = span {hI, ... ,hm } = f(Hi - xt,

and so {HI, H 2 , ... ,Hm} has the strong CHIP.


Proof. Each Hi is the intersection of two half-spaces, Hi = ii2i- I n ii2i , where
ii2i- I = {x E[ (x,h i ) :s; ad and ii2i = {x E[ (x, -hi) :s; -ai}.
Since the collection {iiI, ii2, ... , ii2m - I , ii2m } has the strong CHIP by Theorem
10.7, it follows by Theorem 10.8 that {HI, ... ,Hm} = {ii1 nii2 , ... ,ii2m-lnii2m}
has the strong CHIP, and hence

Moreover, Theorem 10.7 implies that for each x E nr Hi, we have x E n?m iii, so
the active index set is lex) = {l, 2, ... , 2m}. Thus

(0 Hi - X) ° = con {±h i [ i = l, 2, ... , m} = span {hi [ i = l, 2, ... , m} . •

Recall that a polyhedral set is the intersection of a finite number of closed half-
spaces. In particular, every half-space and every hyperplane is a polyhedral set.
STONG CONICAL HULL INTERSECTION PROPERTY (STRONG CHIP) 243

10.10 Example ofthe Strong CHIP (Polyhedral Sets). Any finite collec-
tion of polyhedral sets in a Hilbert space has the strong CHIP.
Proof. Any collection of half-spaces has the strong CHIP by Example 10.7.
Using Lemma 10.8, the result follows. •
10.11 Lemma. Suppose that X is a Hilbert space and {Co, Cl , ... ,Cm } is a
collection of closed convex subsets such that {Cl , ... , Cm} has the strong CHIP.
Then the following statements are equivalent:
(1) {Co, Cl , ... , Cm} has the strong CHIP;
(2) {Co, nrC;} has the strong CHIP.

Proof. (1) ===} (2). This is a consequence of Lemma 10.8.

(2) ===} (1). Suppose that {Co, nrC;} has the strong CHIP. Then for each

r
m
x E nCi , we have that
a

(~Ci - x r = [Co n (0 Ci) - x

= (Co - x)O + (0Ci _x) ° by (2)

by hypothesis

= 2:(Ci -xt·
a
Thus {Co, c l ,.·., Cm} has the strong CHIP. •
Let us now return to the main problem of this chapter. Unless explicitly stated
otherwise, the standing assumptions for the remainder of the chapter are that X is
a Hilbert space, C is a closed convex subset of X, A is a bounded linear operator
from X into £2(m), b = (b l , b2, ... , bm ) E £2(m), and
(10.11.1)
Our main goal is to characterize PK(b)(X), and to describe methods to compute
PK(b)(X), for any x E X.
By Theorem 8.28, A has a representation of the form
(10.11.2)
for some Xi E X, and A* : £2(m) -7 X has the representation
m
(10.11.3)

If Xi = 0 for some i, then we must have bi = 0, since A-l(b) =1= 0 by (10.11.1).


Thus, without loss of generality, we may assume that Xi =1= 0 for every i. Letting
Hi := {x E X I (x, Xi) = bi} (i = 1,2, ... , m), we can rewrite K(b) in the form

(10.11.4) K(b)=cnA-l(b)=cn(0Hi).
244 CONSTRAINED INTERPOLATION FROM A CONVEX SET

Since the Hi are hyperplanes, it follows by Example 10.9 and Lemma 10.11 that
{C,A-1(b)} has the strong CHIP if and only if {C,H1, ... ,Hm } has the strong
CHIP.
By specializing Lemma 10.3 to the situation at hand, we can deduce the following.
Recall the notation R(T) and N(T) for the range and null space of T (see Definition
8.7).
10.12 Lemma. The following statements are equivalent:
(1) {C, A-1(b)} has the strong CHIP;
(2) For every Xo E C n A-1(b),

(10.12.1)

(3) For every Xo E C n A-1(b),


[C n A-1(b) - xot c (C - xot + R(A*);
(4) For every Xo E C n A-1(b),
(10.12.3) con (C - xo) nN(A) c con [(C - xo) nN(A)]

and

(10.12.4) (C - xo)O + R(A*) is closed.

Proof. First observe that for every Xo E C n A-1(b), we have


C n A-1(b) - Xo = (C - xo) n [A-1(b) - xo] = (C - xo) nN(A).

Next note that by Lemma 8.33,

N(At = N(A)J.. = R(A*) = R(A'),


where the last equality holds because R(A*) is finite-dimensional, hence closed by
Theorem 3.7. Using these facts, an application of Lemma 10.3 yields the result .
• Now we can prove our main characterization theorem. It shows that the strong
CHIP for the sets {C, A -1 (b)} is the precise condition that allows us always to
replace the approximation of any x E X from the set K(b) by approximating a
perturbation of x from the set C.
10.13 Theorem. The following statements are equivalent:
(1) {C, A-1(b)} has the strong CHIP;
(2) For every x E X, there exists y E C2 (m) such that

(10.13.1) A[Pc(x + A*y)] = b;

(3) For every x E X, there exists y E C2 (m) such that

(10.13.2) PK(b) (x) = Pc (x + A*y).


STONG CONICAL HULL INTERSECTION PROPERTY (STRONG CHIP) 245

In fact, for any given y E £2(m), (10.13.1) holds if and only if (10.13.2) holds.
Proof. First note that if (10.13.2) holds, then Pc(x + A*y) E K and hence
(10.13.1) holds. Conversely, suppose (10.13.1) holds. Then Xo := Pc(x+A*y) E K
and by Theorem 4.3,
(10.13.3)
Hence X-Xo E (C-xo)o+R(A*) c (K -xo)o. By Theorem 4.3 again, Xo = PK(x).
That is, (10.13.2) holds. This proves the equivalence of (2) and (3) as well as the
last statement of the theorem.
(1) ===} (3). If (1) holds and x E X, let Xo := PK(X). Then Theorem 4.3 implies
x - Xo E (K - xo)o. Using the strong CHIP,
(K - xo)O = (C - xot + R(A*).
Thus x - Xo + A*y E (C - xo)O for some y E £2(m). By Theorem 4.3 again,
Xo = Pc(x + A*y). That is, PK(X) = Pc(x + A*y) and (3) holds.
(3) ===} (1). Suppose (3) holds and let Xo E K. Choose any z E (K - xo)O and
set x := z + Xo. Note that x - Xo = z E (K - xo)O so Xo = PK(x) by Theorem
4.3. Since (3) holds, there exists y E £2(m) such that Xo = Pc(x + A*y). Hence
Theorem 4.3 implies that
z = x - Xo = x - Pc(x + A*y) = x + A*y - Pc(x + A*y) - A*y
E [C - Pc(x + A*y)Jo - A*y c (C - xot + R(A*).
Since z was arbitrary in (K - xo)O, we see that (K - xot c (C - xo)O + R(A*).
Since Xo E K was arbitrary, it follows by Lemma 10.12 that {C, A-1(b)} has the
strong CHIP. •
ReInarks. (1) This theorem allows us to determine the best approximation
in K (b) to any x E X by instead determining the best approximation in C to a
perturbation of x. The usefulness of this is that it is usually much easier to determine
best approximations from C than from the intersection K(b). The price we pay
for this simplicity is that now we must determine precisely just which perturbation
A*y of x works! However, this is determined by the (generally nonlinear) equation
(10.13.1) for the unknown vector y. Since y E £2(m) depends only on m parameters
(the coordinates of y), (10.13.1) is an equation involving only a finite number of
parameters and is often amenable to standard algorithms (e.g., descent methods)
for their solution.
(2) Of course, to apply this theorem, one must first determine whether the pair
of sets {C, A-1(b)} have the strong CHIP. Fortunately, some of the more interesting
pairs that arise in practice do have this property (see Corollary 10.14 beloW).
But even if { C, A -1 (b)} does not have the strong CHIP, we will show below that
there exists a certain convex (extremal) subset C b of C such that K(b) = CbnA-1(b)
and {Cb,A- 1(b)} does have the strong CHIP! This means that we can still apply
Theorem 10.13 but with C replaced by Cb .
(3) As an alternative to applying Theorem 10.13 for the numerical computation
of PK(b) (x), we will show below that Dykstra's algorithm (see Chapter 9) is also
quite suitable for computing PK(b)(X). Moreover, to use Dykstra's algorithm, it
is not necessary to know whether {C,A-1(b)} has the strong CHIP. All that is
needed to apply Dykstra's algorithm is to be able to compute Pc(x) for any x. For
example, in the particular important case when C = {x E L2 I x ::::: O}, we have the
formula Pc(x) = x+ (see Theorem 4.8).
246 CONSTRAINED INTERPOLATION FROM A CONVEX SET

10.14 Corollary. If C is a polyhedral set, then {C,A-l(b)} has the strong


CHIP. Hence for each x EX, there exists y E £2 (m) that satisfies
(10.14.1) A[Pc(x + A*y)] = b.
Moreover, for any y E £2(m) that satisfies (10.14.1), we have
(10.14.2) PK(b) (x) = Pc(x + A*y).

Proof. Note that A-I (b) = nl' Hi, where each Hi is the hyperplane
Hi := {z E X I (z, Xi) = bi } .
It follows by Example 10.10 that {C,A-l(b)} has the strong CHIP. The result now
follows from Theorem 10.13. •
10.15 An Application of Corollary 10.14. Let

C1 = {x E£2 I x(l) ::; O}, C2 = {x E£2 I ~ ~~7~ -2} ,


=

and K = C 1 n C2 . What is the element of K having minimal norm? That is, what
is PK(O)?
Letting hI = (1,0,0, ... ) and h2 = (1/2~, 1/2~, 1/2~, ... , 1/2~, ... ), we see
that hi E £2 for i = 1,2,
C 1 = {x E £21 (x, hI) ::; O} =: C, and C 2 = {x E £21 (x,h 2) = -2} = A- 1 (-2),
where A: £2 -7 JR is defined by Ax = (x,h 2 ). Thus K = CnA-l(-2), and we
want to compute PK(O).
Since C is a half-space, hence polyhedral, it follows by Corollary 10.14 that
(10.15.1)
for any y E JR with
(10.15.2) A[Pc(A*y)] = -2.
But by Theorem 6.31,

Pc(A*y) = Pc(yh2) = yh2 - Ilh~112 {(yh2,h 1 )}+ hI


(10.15.3) = yh2 - {y(h2, hI)} + hI

=yh2 {~Y}+ hI =yh2 - ~y+hl.


Substituting this expression into (10.15.2), we see that we want y E JR such that
(yh2 - ~y+hl,h2) = -2, or

1 +
(10.15.4) Y - -y = -2.
2
If y 2': 0, then y- h+ = h 2': 0 and (10.15.4) cannot hold. Thus we must have
y < O. Then we deduce from (10.15.4) that y = -2, and hence that
(10.15.5)
using (10.15.3). (Note also that IIPK(O) II = 2.)
AFFINE SETS 247

Affine Sets
In general, if C is not polyhedral, it may not be obvious that {C, A-1(b)} has the
strong CHIP, and hence not obvious whether Theorem 10.13 can be applied. What
we seek next is a fairly simple sufficient condition that guarantees that {CA-1(b)}
has the strong CHIP. Such a condition is that b E riA(C), where ri (5) denotes the
"relative interior" of 5 (see Definition 10.21 below). Before we prove this, we must
develop the necessary machinery.
10.16 Definition. A set V in X is called affine ifax + (1- a)y E V whenever
x,y E V and a E R
In other words, V is affine if it contains the whole line through each pair of its
points. Affine sets are also called linear varieties or fiats. Obviously, every affine
set is convex.
10.17 Affine Sets. Let V be a nonempty subset of the inner product space X.
Then:
(1) V is affine if and only ifL.; aivi E V whenever Vi E V and ai E lR satisfy
2::;ai = 1.
(2) V is affine if and only if V = M +v for some (uniquely determined) subspace
M and any v E V. In fact, M = V-V.
(3) If {V;} is any collection of affine sets, then ni V; is affine.

Proof. (1) The "if" part is clear, since we use only the case where n = 2. For
the converse, assume that V is affine. Let Vi E V (i = 1,2, ... , n) and ai E lR
satisfyL.; ai = 1. We must show that L.;
aiVi E V. Proceed by induction on
n. For n = 1, the result is a tautology. Assume the result holds for some n 2: 1.
Suppose that V1, V2,'" ,vnH are in V and 2::;+1 ai = 1. Clearly, not all ai = 1,
since 2::;+1 ai = 1. By reindexing, we may assume that a n +1 # 1. Then

where V := L.~1 00;/(1 - anH)vi E V by the induction hypothesis (because


L.~1 00;/(1- anH) = 1). Then

n+1
2:= aivi = (1 - anH)v + a n+1 Vn+1 E V,
1

since V is affine. This completes the induction.


(2) Let V be affine and v E V. Set M := V-v. We first show that M is a
subspace. Let x,y E M and O! E R Then x = V1 - v, Y = V2 - v for some Vi E V
(i = 1,2). Further, V1 + V2 - v E V by part (1). Thus

x + y = (V1 + V2 - v) - v E V - V = M

and
ax = aV1 - av = [av1 + (1 - a)v]- v E V - v = M.
248 CONSTRAINED INTERPOLATION FROM A CONVEX SET

Thus M is a subspace and V = M + v.


Next we verify the uniqueness of M by showing that M = V-V. Let V = M +v
for some v E V. Then M = V - v C V-V. Conversely, if x E V - V, then
x = V1 - V2 for some Vi E V, so x = (V1 - V2 + v) - v E V - v = M using part (1).
Thus V - V c M, and so M = V - V as claimed.
Conversely, let M be a subspace, v E V, and V = M + v. To show that V is
affine, let x, y E V and a E R Then x = m1 + v and y = m2 + v for some mi E M
(i = 1,2) implies that

ax+ (l-a)y = a(m1 +v) + (1-a)(m2 +v) = [am1 + (1-a)m2l +v EM +v = V.

Thus V is affine.
(3) Let x, y E niVi and a E R Then x, y E Vi for each i and thus ax+ (l-a)y E
Vi for each i, since Vi is affine. Thus ax + (1- a)y E niVi, and so niVi is affine .
• Part (2) states that a set is affine if and only if it is the translate of a subspace.
10.18 Definition. The affine hull of a subset S in X, denoted by aff(S), is
the intersection of all affine sets that contain S.
10.19 Affine Hull. Let S be a subset of the inner product space X. Then:
(1) aff (S) is the smallest affine set that contains S.
(2) aff(S) = {I:~ aiSi I Si E S, I:~ ai = 1, n E N}
(3) S is affine if and only if S = aff(S).
(4) If 0 E S, then aff (S) = span (S).

We leave the proof of this as an easy exercise (see Exercise 12 at the end of the
chapter).
10.20 Definition. The dimension of an affine set V, denoted by dim V, is -1
if V is empty, and is the dimension of the unique subspace M such that V = M + v
if V is nonempty. In general, the dimension of any subset S of X, denoted by
dim S, is the dimension of the affine hull of S.
For example, a single point has dimension 0, the line segment joining two distinct
points has dimension 1, and the unit ball B(O, 1) in X has the same dimension as
X.
Using Lemma 1.15(7), we see that the interior of a set S is the set of all xES
such that B(x, E) C S for some E> o. Thus

int S = {x E SIB (x, E) C S for some E > O} .

10.21 Definition. The relative interior of a set S, denoted by ri S, is the


interior of S relative to its affine hull. More precisely,

ri S := {x E S I B(x, E) n aff S C S for some E > O} .

Note that int S C ri S, but the inclusion is strict in general. Indeed, if X = £2(2)
is the Euclidean plane and S = {x E X I x(2) = O} is the "horizontal axis," then
intS = I/) but riS = S. Moreover, if Sl c S2, then it is obvious that intS1 C intS2.
AFFINE SETS 249

However, it is not true in general that ri 8 1 C ri 8 2 . For example, if D is a cube in


Euclidean 3-space, 8 1 is an edge of D, and 8 2 is a face of D that contains 8 1 , then
the relative interiors of 8 1 and 8 2 are nonempty and disjoint.
There is another important distinction between interior and relative interior.
While it is easy to give examples of nonempty convex sets in Euclidean n-space
(n 2': 1) with empty interior (e.g., a point in I-space), we will see below that the
relative interior of any finite-dimensional nonempty convex set is always nonempty!
The next lemma often helps us to reduce problems concerning relative interiors
to ones involving (ordinary) interiors.
10.22 Lemma. Let 8 be a nonempty subset of X. Then

(10.22.1) aff8 +y = aff(8 + y)

for each y E X.
Moreover, for any x E 8, the following statements are equivalent:
(1) x E ri8;
(2) 0 E ri (8 - x);
(3) 0 E int (8 - x) relative to the space Xo := span (8 - x).

Proof. Let y E X. If x E aff8 + y, then x = L~ O!iSi + y for some Si E 8


and O!i E R with L~ O!i = 1 (by 10.19(2)). Thus x = L~ O!i(Si + y) E aff(8 + y).
This proves that aff 8 + y c aff (8 + y). Since 8 and y were arbitrary, it follows
that aff(8 + y) - y c aff(8 + y - y) = aff8 implies aff(8 + y) C aff8 + y, and so
(10.22.1) holds.
(1) =? (2). If x E ri 8, then B(x, E) n aff 8 c 8 for some E > 0. Since B(x, E) =
X + B(O, E), it follows from (10.22.1) that

(10.22.2) B(O, E) naff(8 - x) c 8 - x,

which is precisely statement (2).


(2) =? (3). If (2) holds, then so does (10.22.2), so that

B(O, E) n Xo c 8 - x,

which is precisely statement (3).


(3) =? (1). If (3) holds, then by retracing the above steps we obtain (1). •
Our next result is the analogue of Theorem 2.8 for relative interiors.
10.23 Theorem. Let K be a convex subset of an inner product space. Then
for each x E ri K, y E K, and 0 < A :s: 1, it follows that

(10.23.1) AX + (1 - A)y E ri K.

In particular, ri K is convex.
The proof is similar to that of Theorem 2.8 using Lemma 10.22.
The next theorem states in particular that the relative interior of a nonempty
finite-dimensional convex set is nonempty.
250 CONSTRAINED INTERPOLATION FROM A CONVEX SET

10.24 Theorem. Let K be a finite-dimensional convex subset in an inner prod-


uct space. Then K and ri K are convex sets that have the same affine hull, and
hence the same dimension, as K. In particular, ri K =1= 0 if K =1= 0, and ri K = K.
Proof. K (respectively riK) is convex by Theorem 2.8 (respectively Theorem
10.23). Moreover, if ri K =1= 0, then ri K = K by Theorem 10.23.
Next suppose that C and Dare nonempty finite-dimensional convex sets with
0= D. We first show that

(10.24.1) affC C affO = affD C affD = affD.

To see this, note that the first inclusion is obvious, since 0 =:> C, while the last
equality holds because aff D is finite-dimensional, hence closed by Theorem 3.7. It
remains to show that aff D C aff D. For if x E aff D and E > 0, choose d1, ... ,dn in
D and a; E R with .E~ a; = 1 such that x = .E~ aid;. Next choose d; E D such
that lid; - d; II < Ej.E j laj I for each i. Then x := .E~ aid; E aff D and

Hence x E aff D and aff D C aff D. This verifies (10.24.1), and thus aff C C aff D.
By interchanging the roles of C and D in (10.24.1), we obtain the reverse inclusion
and, in particular,

(10.24.2) affC = affO = affD = affD.

This verifies that C, 0, D, and D all have the same affine hull and hence the same
dimension. It follows that ifriK =1= 0, then riK, K, and K all have the same affine
hulls, and hence the same dimensions.
To complete the proof, it remains to show that ri K =1= 0 if K =1= 0.
Fix any Xo E K and let n := dimK. Then n 2: 0, and by Theorem 10.17,
aff K = M + Xo for some n-dimensional subspace M. If n = 0, then M = {O} and
aff K = {xo} = K implies that ri K = K. Thus we may assume that n 2: l.
Since
M = affK - Xo = aff(K - xo) = span(K - xo)

is n-dimensional, it follows that there exist n vectors Xl, ... ,Xn in K such that the
set {Xl - XO, X2 -Xo, . .. , Xn -xo} is a basis for M. Put z = n~l (xo +XI + ... +xn).
We will show that z EriK.
First note that

1 1 1
z = --(Xl - xo)
n+1
+ --(X2
n+1
- xo) + ... + -n+1
- ( xn - xo) + Xo.

Since aff K = M + Xo and {Xl - Xo, ... , Xn - Xo} is a basis for M, we see that for
each y E aff K,

n
(10.24.3) y = La;(x; -xo) +Xo
I
RELATIVE INTERIORS AND A SEPARATION THEOREM 251

for some uniquely determined scalars ai. Also,

Ily - zll = II(y - xo) - (z - xo)11 = II~ (a i - n: 1) (Xi - XO)II·


By Theorem 3.7(4), we deduce that two points in a finite-dimensional space are
close if and only if their corresponding coefficients (relative to a given basis) are
close. Hence there exists E > 0 such that if y E B(z, E) n aff K, then

(10.24.4)
l -1«
ai --
n+l
1
n+l
1)2 for i=I,2, ... ,n.

In particular, (10.24.4) implies that ai > 0 for all i and

0< ~a. <~ (la.--


~ ,-~
1
, n+l 1+_1
n+l )
n
< (n+l)2 +--
n+l
n
1 i=l
n + n(n + 1) (n + 1)2
=
(n+l)2
< (n+l)2 =1 .

But then (10.24.3) implies that

n
y = La;(xi - xo) +xo = Laixi
n + (n)
La; Xo
1-
1 1 1

is a convex combination of xo, xl,.·., Xn , and hence y E K.


We have shown that B(z, E) n affK c K, that is, z EriK. •

Relative Interiors and a Separation Theorem


Recall Definition 6.20 of two sets being separated by a hyperplane. The following
separation theorem (which will be needed in a few places later) states that in a
finite-dimensional Hilbert space, two convex sets having disjoint relative interiors
can always be separated by a hyperplane.
10.25 Finite-Dimensional Separation Theorem. IfG, Dare nonempty con-
vex subsets of the finite-dimensional Hilbert space X such that ri G n ri D = 0, then
there exists Xo E X\{O} such that

(10.25.1) sup(Xo, c) :::; inf (xo, d).


cEC dED

In particular, if a is any constant satisfying SUPcEC(xO, c) :::; a :::; infdED (xo, d),
then the hyperplane {x E X I (xo,x) = a} separates G and D.
Proof. We have ri G = C and ri D = D by Theorem 10.24. By continuity,

sup (xo, c) = sup(xo, c) and inf (xo, d) = ini(xo, d),


cEriC cEC dEnD dED

so that by replacing G and D by their respective relative interiors, we may assume


that G nD = 0.
252 CONSTRAINED INTERPOLATION FROM A CONVEX SET

Let E = C-D. Then E is convex by Proposition 2.3(5), and 0 ¢:. E. We consider


two cases.
Case 1: 0 ¢:. E.
In this case, set Xo := -PE"(O) =1= O. Deduce from Theorem 4.1 that

(Xo, Y + xo; ::; 0 for all y E E,

and so (xo, y) ::; -(Xo, Xo) < 0 for all y E E. Thus

(10.25.2) (Xo, c - d) < 0 for all c E C, dE D,

which proves (10.25.1).


Case 2: 0 E E.
In this case, let Xo := spanE = affE = affE. Working in Xo rather than X,
we see that
intE = riE = riE = intE.
Using Lemma 2.9, we deduce

o E E\E C E\intE = E\intE = bdE = bd(Xo\E) c Xo\E.

Thus there exist Xn E XO \E such that Xn -+ O. Setting Zn := PE"(X n ), we obtain


that Xn - Zn =1= 0 and, by Theorem 4.1,

(10.25.3) (Xn - Zn, Y - zn) ::; 0 for all y E E, n E N.

Letting en := (xn - zn)/llxn - znll for each n, we see that Ilenll = 1 and

(10.25.4) for all Y E E, n E N.

Since 0 E E, it follows that

(10.25.5)

That is, Zn -+ O.
Since {en} is a sequence of norm-one elements, Theorem 3.7 implies that there
exists a subsequence {e nk } and a point Xo E Xo such that en, -+ Xo. Since
Ile nk II = 1 for all k, IIxoll = 1. Moreover, in (10.25.4), we obtain

(10.25.6) (Xo, y) ::; 0 for all y E E.

Equivalently,
(xo,c-d)::;O forall cEC,dED,
and so (10.25.1) holds. •
Our next result gives three alternative characterizations of the relative interior of
a convex set in a finite-dimensional space. Each characterization will prove useful
in what follows.
RELATIVE INTERIORS AND A SEPARATION THEOREM 253

10.26 Relative Interior. Let D be a convex subset of a finite-dimensional


inner product space, and let xED. Then the following statements are equivalent:
(1) xEriD;
(2) For each d E D, there exists d' E D and 0 < A < 1 such that x = Ad + (1-
A)d';
(3) For each dE D, there exists f1 > 0 such that x + f1(x - d) ED;
(4) (D - x)O = (D - x).l..

Proof. Since x E ri D if and only if 0 E ri (D - x), by replacing D with D - x,


we may assume x = O.
(1) ===} (2). Suppose 0 E riD. Then there exists E > 0 such that B(O, E) n
spanD c D. Thus for each d E D\{O}, it follows that the element -E(21Idll)-ld is
in D. Then d' := -E(21IdID- 1d ED and 0 = Ad + (1 - A)d', where

E(21Idll)-1
A := 1 + E(21Idll)-1 E (0,1).

(2) ===} (3). If (2) holds, let d E D. Then there exist d' E D and 0 < A < 1 such
that 0 = Ad + (1 - A)d'. Taking fL = l~A > 0, we see that -fLd = d' E D and (3)
holds.
(3) ===} (4). Suppose (3) holds but (4) fails. Then we can choose Z E DO\D.l..
Thus (z, d) :::; 0 for all d E D and (z, do) < 0 for some do E D. Since (3) holds,
there exists f1 > 0 such that d~ := -fLdo ED. Hence

o 2': (z, d~) = -f1(z, do) > 0,


which is absurd.
(4) ===} (1). Suppose DO = D.l.. If 0 1'- riD, then by setting Y = spanD = affD,
we see that 0 1'- int D relative to Y. By the separation theorem (Theorem 10.25) in
the space Y, there exists Xo E Y\{O} such that sup {(xo, d) IdE D}:::; O. Hence

Xo E DO = D.l. = [spanD].l. = y.l. = {O},

which contradicts Xo =I O. •
Remark. Statements (2) and (3) can be paraphrased geometrically to yield
that x E ri D if and only if each line segment in D having x for an endpoint can be
extended beyond x and still remain in D.
In an analogous way, we can characterize the interior of a finite-dimensional
convex set.
10.27 Interior and Relative Interior. Let D be a convex subset of the finite-
dimensional inner product space Y. Then

(10.27.1) riD = {x E D I (D - x)" = (D - x).l.}

and

(10.27.2) intD = {x E D I (D - x)O = {O}}.


254 CONSTRAINED INTERPOLATION FROM A CONVEX SET

Proof. Relation (10.27.1) is a consequence of Theorem 10.26. If x E int D, then


B(x, E) cD for some E> O. Hence B(O, E) CD - x implies that

{OJ = B(O, Et :::> (D - xt :::> {OJ,


so that (D - x)O = {OJ.
Conversely, if xED and (D - x)O = {OJ, then

{OJ = (D - xt :::> (D - x).l :::> {OJ


implies that (D - x)O = (D - x).l = {OJ. By Theorem 10.26, x E riD, or 0 E
ri (D - x). Then
{OJ = (D - x).l = [span (D - x)].l
implies that

Y = [span (D - x)].l.l = span (D - x) = span (D - x)

using Theorem 4.5(9) and the fact that finite-dimensional subspaces are closed.
Since 0 E D - x, span (D - x) = aff (D - x), and hence 0 E int (D - x), or x E int D.
This verifies (10.27.2). •
There are certain relationships between relative interiors of two convex sets that
we record next.
10.28 Lemma. Let K, Kr, and K2 be nonempty convex subsets of a finite-
dimensional Hilbert space. Then:
(1) riK=riK.
(2) KI = K2 if and only ifriK I = riK2 .
(3) riKI nriK2 cri(co{KI UK2 }).

Proof. (1) Since K C K and both K and K have the same affine hull by
Theorem 10.24, we have ri K c ri K. Conversely, let x E ri K. We must show that
x EriK. Choose any Y EriK. If y = x, we are done. Thus assume y oF x. Then
by Theorem 10.26,
z:= x + f.L(x - y) E K

for f.L > 0 sufficiently small. Solve this equation for x to obtain that

1 f.L
x=~~z+~~y,
1+f.L 1+f.L

which is in ri K by Theorem 10.23.


(2) If KI = K 2, then (1) implies that riKI = riKI = riK 2 = riK2. Conversely,
if ri KI = ri K 2 , then Theorem 10.24 implies that K I = ri KI = ri K2 = K 2.
(3) Let x E riKI n riK2 and let y E Ko := co (KI U K2)' By Theorem 10.26, it
suffices to show there exists f.L > 0 such that x + f.L(x - y) E Ko. First note that
y = ).,YI + (1- ).,)Y2 for some Yi E Ki and 0:::: ).,:::: 1. Since x E riKi (i = 1,2),
Theorem 10.26 implies that there exist f.Li > 0 such that

(10.28.1) (i = 1,2).
RELATIVE INTERIORS AND A SEPARATION THEOREM 255

By convexity of each K i , (10.28.1) must also hold when f-li is replaced by f-l :=
min{f-ll> f-l2}:

(10.28.2) (i = 1,2).

It follows that f-l > 0 and, from (lO.28.2), that


x + f-l(x - y) = x + f-l[x - {AYI + (1 - A)Y2}]
= A[x + f-l(x - Yl)] + (1 - A)[x + f-l(x - Y2)] E Ko· •

10.29 Theorem. Let X be a Hilbert space, C a closed convex cone, and M a


finite-dimensional subspace such that C n M is a subspace. Then C + M is closed.
Proof. Assume first that C n M = {O}. Let Xn E C + M and Xn -+ x. We must
show that x E C + M. Write Xn = Cn + Yn, where en E C and Yn EM. If some
subsequence of {Yn} is bounded, then by passing to a subsequence if necessary, we
may assume that Yn -+ Y E M. Then Cn = Xn - Yn -+ X - y. Since C is closed,
c:= x - Y E C and x = c + Y E C + M.
Hno subsequence of {Yn} is bounded, then IIYnll-+ 00. It follows that {Yn/IIYnll}
is bounded in M, so by passing to a subsequence if necessary, we may assume that
Yn/IIYnll -+ Y E M and Ilyll = 1. Then en/IIYnll E C for all nand
cn Xn Yn
llYn II = IIYnl1 - IIYnl1 -+ 0 - Y E C,

since C is closed. Thus -Y E CnM = {O}, which contradicts IIYII = 1. This proves
the theorem when C n M = {O}.
In general, V = C n M is a closed subspace of M, so we can write M = V +
V..L nM. Then

(10.29.1) C+M=C+ V+ V..LnM=C+ V..LnM

and

C n (V..L n M) = V n V..L = {O}.

By the first part of the proof (with M replaced by V..L nM), we see that C+ V..LnM
is closed. Thus C + M is closed by (10.29.1). •
The next result shows that the unit ball in a finite-dimensional space is contained
in the convex hull of a finite set. This result is obviously false in an infinite-
dimensional space.
10.30 Theorem. Let {e 1 , e2, ... , en} be an orthonormal basis for the inner
product space Y. Then

(10.30.1)

In other words, if Y E Y and Ilyll < 1/ yin, then there exist 2n scalars Ai 2' 0 with
Lin Ai = 1 such that Y = L~=l (Ai - An+i)ei.
256 CONSTRAINED INTERPOLATION FROM A CONVEX SET

Using Schwarz's inequality in £2(n), we obtain

Moreover, since ak = at - a k , we set

if 1 :s: i :s: n,

if n < i :s: 2n,

and deduce that Ai 2': 0, Lin Ai = 1, and

n 2n
LAiei+ L Ai(-ei-n) Eco{±el,±e2, ... ,±en }.
i=l i=n+l

Further, 0 E co {±el, ±e2, ... , ±e n } implies that


n n
Y= Latei + Lai(-ei)
1 1

= t 1
lajl [t
1
Aiei + f
n+l
Ai( -ei-n)] + (1 -t 1
lajl) ·0
E co{±el, ±e2,·.·, ±en}· •

It turns out that the constant 1/ fo in the inclusion (10.30.1) is "best possible" in
the sense that (10.30.1) fails if l/fo is replaced by any larger number (see Exercise
17 at the end of the chapter).
We need one more result of a general nature concerning the dual cones of images
of linear mappings.
10.31 Dual Cone of Images. Let A be a bounded linear mapping from a
Hilbert space X to an inner product space Y. If S is any nonempty subset of X,
then

(10.31.1)

and

(10.31.2)
RELATIVE INTERIORS AND A SEPARATION THEOREM 257

Proof. We have that y E [A(SW ~ (y, As) :s: 0 for all s E S ~ (A*y, s) :s: 0
for all s E S ~ A*y E So ~ Y E (A*)-l(SO). This proves (10.31.1). The proof
of (10.31.2) is strictly analogous. •
Now we can prove one of our main results. It gives an important sufficient
condition that guarantees the strong CHIP.

For the remainder of this chapter, unless explicitly stated otherwise,


we assume the following hypothesis: X is a Hilbert space, C is a closed
convex subset, bE £2(m), A: X -+ £2(m) is defined by

where {Xl,X2, ... ,Xm } C X\{O} is given, and b E £2(m). Moreover, we


assume that

(i = 1, ... ,m)}

is nr)t empty. That is, bE A(C).

10.32 Theorem. Ifb E riA(C), then {C,A-l(b)} has the strong CHIP.
Proof. By Lemma 10.12, it suffices to show that for each fixed x E C n A-l(b),

(10.32.1) (C - xt + R(A*)
is closed and

(10.32.2) con (C - x) nN(A) c con [(C - x) nN(A)].

Since b E riA(C), Theorem 10.26 implies that

(10.32.3) [A(C) - W= [A(C) - b].l.

Since Ax = b, we deduce that

[A(C - x)]O = [A(C - x)].l.

Using Theorem 10.31, it follows that

or, equivalently,

(10.32.4) R(A*) n (C - x)O = R(A*) n (C - x).l.


Since R(A*) and (C-x).l are both subspaces, (10.32.4) shows that R(A*)n(C-x)O
is a subspace. Using Theorem 10.29 (with C and M replaced by (C-x)O and R(A*),
respectively), we deduce that (10.32.1) holds.
It remains to verify (10.32.2). Let Z E con (C - x) n N(A). To complete the
proof, we will show that there exists a sequence {zn} in con [( C - x) n N (A)] such
258 CONSTRAINED INTERPOLATION FROM A CONVEX SET

that Zn -+ z, and hence Z E con [(C - x) nN(A)]. If Z = 0, we can take Zn = 0 for


each n. Thus we may assume that Z f O.
Since Z E con(C - x), there exists a sequence {z~} in con(C - x) such that
z~ -+ z. By Lemma 4.4(5), there exist Cn E C and Pn > 0 such that z~ = Pn(cn -x).
If z~ E N(A) for infinitely many n, say Z~k E N(A) for k = 1,2, ... , then

_l_Z~k = cnk - x E (C - x) nN(A),


Pnk
so Z~k E con [(C - x) nN(A)] for all k. Setting Zk = Z~k' we see that

Zk E con [(C - x) nN(A)] and Zk -+ Z.

Thus, by passing to a subsequence if necessary, we may assume that z~ 1'. N(A)


for each n. If Ilcn - xii> 1 for some n, then since C - x is convex and 0 E C - x,
we see that

so (cn - x) /llcn - xii = c~ - x for some c~ E C and IJc~ - xii = 1. Hence

Z~ = Pn(en - x) = p~(c~ - x),

where P~ = Pnllcn - xii> 0 and Ilc~ - xii = 1. In short, we may assume that each
z~ has the representation
Z~ = Pn(cn - x),
where Pn > 0, Cn E C, and lien - xii :s: l.
Since z~ -+ z, by passing to a subsequence, we may assume that

(lO.32.4) ~llzll < Ilz~11 = Pnllcn - xii :s: Pn


for each n. Since Az = 0, we see that

(lO.32.5) Yn := Az~ -+ Az = O.

Then Yn f 0 for each n; and

Yn = PnA(cn - x) = Pn[Acn - b],

so
1
(lO.32.6) -Yn +b= ACn E A( C) for each n.
Pn
Hence, using (lO.32.4), we see that

(lO.32.7) In := II ~: II = P~ IIYnl1 < II~II llYn II -+ O.


RELATIVE INTERIORS AND A SEPARATION THEOREM 259

Since b E ri A( C), there exists 5 °


> such that
B(b, 25) naffA(C) c A(C).
Subtract Ax = b from both sides of this inclusion to obtain
(10.32.8) B(O, 25) n span [A(C - x)] c A(C - x).
Next we choose an orthogonal basis {el, ... , eN} of span [A(C-x)] such that Ileill =
5 for each i. Then {±el, ±e2, ... , ±eN} C A(C - x) by (10.32.8), and by convexity
of A(C - x), there even follows co {±el, ±e2, .. ·, ±eN} c A(C - x). Now choose
Ui E C such that
(i = 1,2, ... ,N),
and set
p,:= max{llui -xii + 25 I i = 1,2, ... ,2N}.
Using Theorem 10.30, we deduce that
el e2 eN}
A[(C - x) n B(O,p,)] ::J co{±el, ±e2,···, ±eN} = 5 co { ±"5' ±"5' ... ' ±T

(0, ~ ) = B (0, ~ ) .
::J 5B

°: ; ), : ; Thus, for each °< ), ::; 1, we have


Sinc'C C - x is convex and C - x, it follows that )'(C - x)
0 E C C - x for every
1.
A[(C - x) n B(O, ),p,)] ::J A[)'(C - x) n )'B(O, p,)]
= )'A[(C - x) n B(O,p,)]
(10.32.9)
::J )'B (0, ~ ) = B (0,), ~ ) .
Setting I := ),51 m, it follows from (10.32.9) that ),p, = (rm I 5)p, = IV' where
V := (mI 5)p, > 0, and hence
(10.32.10) A[(C - x) n B(O, IV)] ::J B(O, I)
for all °< ,::; 51m. Add Ax = b to both sides of (10.32.10) to get
(10.32.11) A[C n B(x, IV)] ::J B(b, ,)
°
for all < I ::; 51m. By (10.32.7), In = IIYnlPnl1 -+
many n, we may assume that
°so by omitting finitely
(10.32.12)
In = II~:II < ~ for all n.

Note that In > 0, In -+ 0, and InPn = IIYnl1 -+ 0. By (10.32.11), we can choose


XnE C n B(x, Inv) such that

AXn = b - Yn for all n.


Pn
°
Thus Ilxn - xii < InV -+ and A(xn - x) = -Ynl Pn implies that A[Pn(xn - x)] =
-Yn. Hence the element Zn := z~ + Pn(x n - x) is in con (C - x) and AZn = 0. That
is,
Zn E con (C - x) nN(A) = con [(C - x) nN(A)],
where we used Lemma 10.6 for the last equality, and
Ilzn - zil ::; Ilzn - z~11 + Ilz~ - zil ::; Pnllxn - xii + Ilz~ - zil
< PninV + Ilz~ - zll = IIYnll v + Ilz~ - zll -+ 0. •
260 CONSTRAINED INTERPOLATION FROM A CONVEX SET

10.33 Corollary. Suppose b E riA(C). Then, for each x E X, there exists


y E £2(m) such that

(10.33.1) (j=1,2, ... ,m).

Moreover, for any y E £2(m) that satisfies (10.33.1), we have

(10.33.2)

Proof. By Theorem 10.32, {C,A-l(b)} has the strong CHIP. By Theorem


10.13, PK(X) = Pc(x + A*y) for any y E £2(m) that satisfies A[Pc(x + A*y)] = b.
By virtue of (10.11.2) and (10.11.3), the result follows. •
10.34 An Application. Let T be any set that contains at least m points and
let X = £2 (T). Fix any m points tl, t2, . .. , tm in T, define Xi E X by

ift i= ti,
if t = ti,

for i= 1,2, ... ,m, put C = {x EX! x(t);:>: 0 for all t E T}, and define A: X-t
£2(m) by

Clearly,
A(C) = {y E £2(m) ! y(i);:>: 0 for i = 1,2, ... ,m}.
Claim 1: ri A(C) = int A(C) = {y E £2(m) ! y(i) > 0 (i = 1,2, ... , m)}.
To verify this claim, let

D={yE£2(m)!y(i»0 for i=1,2, ... ,m}.

Note that D c A(C). Now let Yo ED and set E:= ~min{yo(i)! i = 1,2, ... ,m}.
Thus E > 0 and if lIy - Yoll < E, then L~l !y(i) - yo(i)J2 < E2 implies that
!y(i) - yo(i)! < E for each i, so that

y(i) > yo(i) - E ;:>: 2E - E = E >0

for each i. That is, y E D. Thus

(10.34.1) Dc intA(C) c riA(C).


Conversely, let y E riA(C). Then y(i) ;:>: 0 for each i. If y(i o) = 0 for some
io E {I, 2, ... , m}, define d E £2(m) by setting d(i o) = 1 and d(i) = 0 if i i= io.
Then dE A(C), so that by Theorem 10.26, there exist d' E A(C) and 0 < ..\ < 1
such that y = ..\d + (1 - ..\)d'. Then

0= y(io) = ..\d(i o) + (1 - ..\)d'(io) = ..\ + (1- ..\)d'(io) ;:>: ..\ > 0,


EXTREMAL SUBSETS OF C 261

which is absurd. Thus we must have y( i) > 0 for all i. This proves that ri A( C) c D
and, combined with (10.34.1), verifies Claim l.
Now fix any bE C2 (m) with b(i) > 0 for i = 1,2, ... ,m, and let

K = K(b) := C n A-l(b).

By Claim 1, bE riA(C), and hence K # 0.


Claim 2: For each x E X, PK(x) is the function defined by

x+(t) if t E T\{h, t2, ... , t m },


PK(X)(t) = { b(i)
ift = k

To verify this claim, we note that by Corollary 10.33, there exists y E C2 (m) such
that

(10.34.2) (j = 1,2, ... ,m)

and

(10.34.3)

for any y E C2 (m) that satisfies (10.34.2). By Application 4.8, we have that Pc(z) =
z+ for any z E X. Using this and the fact that (z,Xj) = z(tj ) for any z E X, we
see that (10.34.2) may be reduced to

or, more simply,

(10.34.4) for j = 1,2, ... , m.

But y(j) := b(j) - x(tj) (j = 1,2, ... , m) clearly satisfies (10.34.4). Using this, we
deduce from (10.34.3) that

Evaluating this at any t E T, we see that Claim 2 is verified.

Extremal Subsets of C
The above argument would not be valid if b E A(C)\riA(C). How would we
proceed in this case? It turns out that there is a result analogous to Corollary
10.33 even if b <Ie riA(C)! However, in this case, the set C in equations (10.33.1)
and (10.33.2) must be replaced by a certain convex extremal subset of C. To
develop the theory in this more general situation, we first must define the notion
of an extremal set.
262 CONSTRAINED INTERPOLATION FROM A CONVEX SET

10.35 Definition. A subset E of a convex set S is called an extremal subset


of S if x , y in S, 0 < A < 1, and AX + (1 - A)Y E E implies that X, y E E. In other
words, if an interior point of a line segment in S lies in E, the whole line segment
must lie in E.
An extreme point of S is an extremal subset of S consisting of a single point.
Thus e E S is an extreme point of S if X, YES, 0 < A < 1, and e = AX + (1 - A)y
implies that X = Y = e.
For example, the convex extremal subsets of a cube in 3-space are the whole cube,
the six faces, the twelve edges, and the eight vertices. The set of extreme points are
the eight vertices. Also, for the unit disk in the plane, the convex extremal subsets
are the whole disk and each point on the boundary (see Figure 10.35.1) .

.............

Cube Disk

Figure 10.35.1

It is easy to verify that any intersection of extremal subsets of S is again an


extremal subset of S (possibly empty).
Now we shall consider what can be said in the case where {C, A-1(b)} does
not have the strong CHIP. The following proposition is the main result in this
regard. It states that there exists a closed convex extremal subset C b of C such
that C b n A-1(b) = C n A- 1(b), b E riA(Cb), and {Cb , A- 1(b)} has the strong
CHIP! In other words, since the set K(b) = C n A-1(b) is unchanged if we replace
C with Cb, we can apply Theorem 10.33 (with C replaced by Cb ) to obtain a
characterization of best approximations from the set K(b) that does not require
bE riA(C).

10.36 Proposition. There exists a unique convex subset C b of C such that


b E ri A( C b) and Cb is larger than any other subset with this property. That is, if
C' is a convex subset ofC and b E riA(C'), then C' c Cb. Moreover,
(1) C b is closed,
(2) Cb n A-1(b) = C n A-l(b)(= K(b)),
(3) A(Cb) is an extremal subset of A(C),
(4) Cb is an extremal subset ofC,
(5) {Cb, A-I (b)} has the strong CHIP,
(6) C b = C if and only if bE ri A(C).

Proof. Let

C:= {C' I c' c C, C' is convex, bE riA(C')}.


EXTREMAL SUBSETS OF C 263

The first statement of the proposition is that C has a unique largest element C b .
Let us assume this for the moment, and show how the remaining six statements
follow.
(1) If Cb were not dosed, then Co := C b would be a convex subset of C that is
strictly larger than Cb . Moreover,

where the first inclusion follows from the continuity of A. It follows that

By Lemma 10.28, riA(Cb ) = ri A(Co). Since b E riA(Cb), we have that b E riA(Co)


and thus Co E C. But this contradicts the maximality of Cb.
(2) If CbnA-I(b) =I K, choose x E K\CbnA-I(b). Then x E C\Cb and Ax = b.
Consider the set Co := co (Cb U {x}). Then Co is a convex subset of C that is
strictly larger than Cb. Also, since Ax = b E riA(Cb) C A(Cb), and since A(Cb) is
convex, we deduce that

bE riA(Cb) = ri{coA(Cb)} = ri {co [A(CbU{X})]} = riA[co(CbU{X})] = riA(Co).

Thus Co E C, and this contradicts the maximality of Cb .


(3) If A(Cb) were not extremal in A(C), there would exist YI,Y2 in A(C), YI t/:.
A(Cb), and 0 < A < 1 such that

Choose Xl E C\Cb and X2 E C such that AXi = Yi (i = 1,2). Then

Since bE riA(Cb), Theorem 10.26 implies that there exists p > 0 such that

(10.36.1) Y3 := b + p(b - y) E A(Cb )·

Choose X3 E Cb such that AX3 = Y3. Solve (10.36.1) for b to obtain

Thus

This proves that C' := co {Xl, X2, X3} E C, and hence C' C Cb. But this contradicts
Xl t/:. C b •
(4) Let XI,X2 E C, 0 < A < 1, and x = AXI + (1- A)X2 E Cb. Then
264 CONSTRAINED INTERPOLATION FROM A CONVEX SET

By (3), A(Cb) is extremal in A(C), and hence AXi E A(Cb) for i = 1,2. If Xl tI. C b,
then Co := co (Cb U {Xl}) is a convex subset of C strictly larger than C b and
aff A( Co) = aff A( Cb ). It follows that

bE riA(Cb) C riA(Co).

Hence Co E C, which contradicts the maximality of Cb. This proves that Xl E C b.


Similarly, X2 E Cb· Hence C b is extremal in C.
(5) Since bE ri A(Cb) and K = Cb n A-1(b) by (2), it follows by Theorem 10.32
(with Cb in place of C) that {Cb, A -1 (b)} has the strong CHIP.
(6) If C = C b, then b E riA(Cb) = riA(C). Conversely, if bE riA(C), then since
C b is the largest convex subset C' of C with b E riA(C'), we have that C b = C.
Thus, to complete the proof, we must show that C has a unique largest element.
For this, we use Zorn's lemma (see Appendix 1). Order C by containment: C 1 >- C 2
if and only if C 1 =:> C2 . Let T be a totally ordered subset of C and set Co :=
U{C' I c' E T}.
Claim 1: Co E C.
First we show that Co is convex. Let X, y E Co and 0 < A < 1. Then X E C 1
and y E C 2 for some C i E T (i = 1,2). Since T is totally ordered, either C 1 C C 2
or C 2 C C 1 • We may assume C 1 C C 2. Then X,y E C 2 , and since C 2 is convex,
AX + (1 - A)y E C 2 C Co, and Co is convex.
Next we show that bE riA(Co). Let Yo E A(Co). Then Yo E A(C') for some
C' E T, and since b E riA(C'), Theorem 10.26 implies that there exists f1 > 0 such
that b + f1(b - yo) E A(C') c A(Co). By Theorem 10.26 again, bE ri A(Co). Thus
Co E C and Claim 1 is proved.
Moreover, it is clear that Co =:> C' for every C' E T. Thus Co is an upper bound
for T. By Zorn's lemma, C has a maximal element C b .
Claim 2: C b =:> C' for all C' E C.
For otherwise C1 rt. Cb for some C 1 E C. Then the set C2 := co{C1 U C b} is a
convex subset ofC that is strictly larger than C b · Also, since bE riA(C1 )nriA(Cb ),
it follows from Lemma 10.28(3) that

Thus C 2 E C, C2 =:> C b, and C2 =1= Cb. This contradicts the maximality of Cb, and
proves Claim 2.
Finally, if C b and C~ are both maximal elements for C, then Claim 2 shows that
Cb =:> Cb and Cb =:> Cb. That is C£ = C b, and the maximal element is unique. •
There is an alternative description for the set C b that is often more useful for
applications. We first observe the following lemma.
10.37 Lemma. If S is a nonempty subset of a convex set C, then there exists
a (unique) smallest convex extremal subset of C that contains S, namely,

(10.37.1) n{E I E is convex, extremal, and SeE c C}.

Proof. Since the intersection of extremal sets is extremal, we see that the set
in (10.37.1) is extremal, convex, and is the smallest such set that contains S. •
EXTREMAL SUBSETS OF C 265

10.38 Proposition. Let Eb denote the smallest convex extremal subset of C


that contains K(b) = C n A-l(b). Then C b = Eb.
In particular, Cb is the largest convex subset of C with the property that b E
riA(Cb ), and C b is also the smallest convex extremal subset of C that contains
C n A-l(b).
Proof. Assume first that the following three properties hold:

(10.38.1) bE riA(Eb),
(10.38.2) A(Eb) is extremal in A(C),
(10.38.3) A- 1 (A(Eb)) n C = Eb.

Under these conditions, we will prove that Cb = E b. Indeed, from (10.38.1) and
the definition of Cb, we see that C b ::::l Eb. For the reverse inclusion, let y E A( C b).
Since b E ri A( C b ) by Theorem 10.36, there exist (by Theorem 10.26) y' E A( Cb )
and 0 < A < 1 such that AY + (1 - A)Y' = b E A(Eb). Since A(Eb) is extremal in
A(C), y and y' are in A(Eb). That is, A(Cb) C A(Eb) and hence

by (10.38.3). Thus C b = Eb, and the result is verified.


Hence to complete the proof, it suffices to verify (10.38.1)-(10.38.3). To this end,
let Fe denote the smallest convex extremal subset of A( C) that contains b, and set

(10.38.4)

That is, Fe is the intersection of all convex extremal subsets of A( C) that contain
b.
Clearly, CF.::::l CnA-l(b) = K. Now, C Fe is convex, since Fe is convex and A
is linear. Next let x, yin C, 0 < A < 1, and z:= AX + (1- A)y E C F •. Then z E C
and
AAx + (1 - A)Ay = Az E Fe.
Since Fe is extremal in A(C), we have that Ax and Ay are in Fe. It follows that
x and yare in C n A-l(Fe) = C F •. This proves that C F • is extremal in C. Since
CFe ::::l K, we deduce C Fe ::::l E b .
The mapping Ae := Ale from C to A(C) is surjective, so that

That is,

(10.38.5)

Next we verify that bE ri Fe. This is equivalent to 0 E ri (Fe-b). Since 0 E Fe-b,


we see that air (Fe - b) = span (Fe - b). By working in the space Yo := span (Fe - b)
rather than £2 (m), it suffices to show that 0 E int (Fe - b). If 0 rJ. int (Fe - b), then
the separation theorem (Theorem 10.25) implies that there exists y E Yo\{O} such
that

(10.38.6) (y, f - b) ::: 0 for all f E Fe·


266 CONSTRAINED INTERPOLATION FROM A CONVEX SET

If y E (Fe - b)..L, then


Y E [span (Fe - b)]..L = Yo..L = {O},
which contradicts y # O. Thus there exists fa E Fe such that (y, fa - b) < O. Set
(10.38.7) H = {z E Yo I (y, z) = O} .
Then the hyperplane H supports Fe - bat 0, and E := H n (Fe - b) is a convex
extremal subset of Fe - b such that E # Fe - b (since fa - b <t- E). [To see that E
is extremal, let fi E Fe (i = 1,2), 0 < A < 1, and suppose e := A(fI - b) + (1 -
A)(12 - b) E E. Then e E H, so
0= (e,y) = A(fI - b,y) + (1- A)(12 - b,y):S: 0
by (10.38.6). Thus equality must hold, so (Ii - b, y) = 0 for i = 1,2, and therefore
1; - bEE (i = 1,2).] It follows that E + b is a convex extremal subset of Fe with
bEE + band E + b # Fe. Since Fe is extremal in A(G), E + b is also extremal in
A(G). But this contradicts the minimality of Fe. Hence 0 E int (Fe - b) and thus
(10.38.8) bE riFe = riA(GFJ.
We next show that GFe C Eb and hence, since GFe :::l Eb, that GFe = Eb. Let
Xo E GFe' Then Xo E G and Axo E Fe. Since b E riFe by (10.38.8), Theorem 10.26
implies that there exist y' E Fe and 0 < A < 1 such that b = AAxo + (1- A)Y' E Fe.
By (10.38.5), we can choose x~ E GFe such that y' = Ax~. Then
b = AAxO + (1 - A)Ax~ = A[AXo + (1 - A)X~].
Thus the element x := AXo + (1 A)X~ is in G and Ax = b. That is, x EKe Eb.
Since Eb is extremal in G, it follows that Xo, x~ E Eb. Thus GFe C Eb, and hence
(10.38.9)
Using (10.38.9), (10.38.5), and the fact that Fe is extremal, we obtain (10.38.2).
From (10.38.8) and (10.38.9) we obtain (10.38.1).
Finally, using (10.38.9) and (10.38.5), we obtain
A-1(A(Eb)) n G = A-1(A(GFJ) n G = A-1(Fe ) n G = GFe = Eb.
This verifies (10.38.3) and completes the proof. •
For the actual computation of Gb , the following result is sometimes helpful.
10.39 Proposition. Let E be a convex extremal subset of G that contains
G n A-l(b). Then the following statements are equivalent:
(1) E = Gb ;
(2) bEriA(E);
(3) [A(E) - W = [A(E) - b]..L.

Proof. The equivalence of (2) and (3) is from Theorem 10.27. The implication
(1) ==> (2) is a consequence of Proposition 10.36. Finally, if (2) holds, then Gb :::l E
follows from Proposition 10.36. But Proposition 10.38 implies that E :=> Gb. Thus
E=Gb . •
As an immediate consequence of Proposition 10.36 and Corollary 10.33, we ob-
tain the main result of this chapter.
EXTREMAL SUBSETS OF C 267

10.40 Perturbation Theorem. For each x E X, there exists y E R2(m) such


that

(10.40.1)

For any y E R2(m) that satisfies (10.40.1), we have

(10.40.2) PK(b) (x) = PCb (x + ~ Y(i)X i) .

Moreover, C b = C if and only if b E ri A( C).


Thus the problem of computing PK(b) (x) reduces to that of solving the (generally
nonlinear) m equations (10.40.1) for the m unknowns Y = (y(I), y(2), ... , y(m)).
This in turn requires being able to compute PCb(Z) for any Z E Z. If bE riA(C),
then C b = C, and this requires being able to compute Pc(z) for any z E X. If
b i riA(C), then Cb of C, and we first have to determine the set C b. In this regard,
it is important to know how to recognize when C b = C, that is, when b E riA(C).
The following result is helpful in this regard.
10.41 Lemma. The following six statements are equivalent:
(1) C=Cb ;
(2) bE riA(C);
(3) [A(C) - W = [A(C) - W;
(4) R(A*) n [C - K(bW c [C - K(b)].L;
(5) R(A*) n (C - x)o C (C - x).L for each x E K(b);
(6) R(A*) n (C - x)o C (C - x).L for some x E K(b).

Proof. The equivalence of the first three statements follows from Proposition
10.39.
Now fix any x E K(b). Then
(10.41.1) A(C - K(b)) = A(C) - A(K(b)) = A(C) - b = A(C - x).
Using (10.41.1), we see that (3) holds if and only if
(10.41.2) [A(e - K(b)W = [A(C - K(b))].L
if and only if
(10.41.3) [A(C - xW = [A(C - x)].L.
Using Proposition 10.31, we deduce that (3) holds if and only if
(10.41.4) (A*)-l(C - x)o = (A*)-l(C - x).L
if and only if
(10.41.5)
Since it is always true that S.L c So for any set S, the equal signs in (10.41.4) and
(10.41.5) may be replaced by the inclusion symbol "c." From this, it is now easy
to verifY the equivalence of (3)-(6), and we omit the details. •
Before giving some applications of these results in the most important case (i.e.,
when C it; the cone of nonnegative functions), we shall establish some improvements
of the previous results in the particular case where C is a convex cone.
268 CONSTRAINED INTERPOLATION FROM A CONVEX SET

10.42 Definition. A convex subset C of X is said to be generating for X if


its span is dense in X, i.e., span C = X.
10.43 Example. Let X = .e2 (T) and let C be the positive cone in X:

C = {x E X I x(t) ::::: 0 for all t E T}.


Then it is easy to see that for each x E X, x = x+ - x-, and since both x+ and
x- are in C, X = C - C. In particular, X = span (C), so C is generating.
Similarly, the positive cone is generating in C2 [a, bj and in L 2 [a, bj.
10.44 Lemma. A convex subset C of a Hilbert space X is generating if and
only ifCl.. = {O}.
Proof. If C is generating, then span C = X. For any x E Cl.., choose a sequence
{x n } in spanC such that Xn -t x. Thus Xn = L~\n) anicni for some ani E lR and
Cni E C. Then (x, Cni) = 0 for all nand i, so that
N(n)
IIxl1 2 = (x, x) = lim(xn,
n
x) = lim'"'
n L.....t
ani (Cni, x) = O.
i=l

Hence x = 0, and thus Cl.. = {O}.


Conversely, if Cl.. = {O}, then

X = {O}l.. = C.l...l.. = {spanC} 1..1.. = spanC


using Theorem 4.5(3) and (9). Thus C is generating. •
When C is a convex cone, there is a strengthening of Lemma 10.41.
10.45 Lemma. Let C be a closed convex cone in the Hilbert space X. Then
the following five statements are equivalent:
(1) C = C b ;
(2) bE riA(C);
(3) [A(CW n bl.. c [A(C)jl..;
(4) R(A*) nco n [K(b)jl.. c cl..;
(5) R(A*) nco n {x}l.. C Cl.. for some x E K(b).
In addition, if C is generating, then these five statements are equivalent to each
of the following three:
(6) [A(CW n bl.. c N(A*);
(7) R(A*) nCo n [K(b)jl.. = {O};
(8) R(A*) nCo n {x}l.. = {O} for some x E K(b).
Finally, if C is a generating closed convex cone and A is surjective, these eight
statements are equivalent to each of the following two:
(9) [A(CW n bl.. = {O};
(10) bEintA(C).

Proof. The equivalence of the first four statements follows from Lemma 10.41,
the fact that A(C) is a convex cone, and Theorem 4.5(5). The equivalence of (4)
and (5) follows from the fact that

(10.45.1) R(A*) nCo n [K(bW = R(A*) nCo n {x}l..


EXTREMAL SUBSETS OF C 269

is an identity for any x E K(b). To verify this, it clearly suffices to show the
inclusion

(10.45.2) R(A*) nCo n [K(b)].L ~ R(A*) nCO n {x}.L.

But if Z E R(A*) nCo n {x}.L, then z = A*y for some y E £2(m), z E Co, and
(z, x) = O. Then for any Xl E K(b), we have that

(z, Xl) = (A*y, Xl) = (y, AXI) = (y, b) = (y, Ax) = (A*y, x) = (z, x) = o.
Thus z E [K(b)].L and (10.45.2) is verified.
If C is also generating, then C.L = {O} by Lemma 10.44 and

by Theorem 10.31. This proves the equivalence of (3) and (6), and of (4) and (7).
The equivalence of (7) and (8) follows from (10.45.1).
If, moreover, A is surjective, then Lemma 8.33 implies that

Thus (6) and (9) are equivalent. But using 4.5(5), we can rewrite (9) as

{O} = [A(C) - W·
Since {O} C [A(C) - b].L c [A (C) - W is always true, it follows that

{O} = [A(C) - W = [A(C) - b].L.

Using Theorem 10.27 applied to D = A(C), we deduce that (9) is equivalent to


(10). •
Next we observe that if C is a convex cone, then any convex extremal subset of
C must be a cone. In particular, Cb is a convex cone.
10.46 ExtreInal Subsets of Cones are Cones. If E is a convex extremal
subset of a convex cone C, then E is a convex cone.
Proof. Let x E E. If p > 1, then since 0 and px are in C and since

~ (px) + (1 - ~ ) 0= x E E,

it follows by extremality that px and 0 are in E. In particular, px E E for all p ~ l.


If 0 ::: A ::: 1, then by convexity of E, AX = AX + (1 - A)O E E. This proves that
px E E for all p ~ O. Finally, if x, y E E and p, fl, > 0, then by convexity of E,

p fl,
z := - - x + - - y E E,
P+fl, P+fl,

and px + fl,y = (p + fl,)z E E by the statement above. Thus E is a convex cone .


270 CONSTRAINED INTERPOLATION FROM A CONVEX SET

Constrained Interpolation by Positive Functions


Now we will exploit these results for what is perhaps the most important case:
where C is the convex cone of nonnegative functions. In this case, C is a generating
cone.
Let T be a nonempty set, X = £2(T),

C = {x E X I x(t) 2: °
for all t E T},

By Theorem 8.28, the adjoint mapping A* : £2(m) -+ X is given by


m

For any given bE £2(m), let

K(b) : = C n A-l(b)
(10.46.1)
= {x E X I x 2: 0, (x, Xi) = b(i) for i = 1,2, ... ,m)}.

We assume that K(b) i= 0, i.e., bE A(C).


We will give a useful formula for PK(b) (x) for any x E £2(T).
For any x E £2(T), the support of x is defined by

(10.46.2) suppx:= {t E T I x(t) i= O}.


Also, for any subset S of T, we define the subset C(S) of C by

(10.46.3) C(S) := {x E C I suppx c S}.


The importance of the set C(S) is that it is equal to the minimal convex extremal
subset C b of C for a particular choice of S! Before we verify this, it is convenient
to establish some basic properties of C(S).
10.47 Properties of C(S). Fix any subset S ofT. Then:
(1) C(S) is a closed convex cone that is extremal in C.
(2) C(T) = C and C(0) = {O}.
(3) C(S)O = {x E £2(T) I x(t) :s; 0 for all t E S}.
(4) C(S)1. = {x E £2(T) I x(t) = 0 for all t E S}.

Proof. (1) Let (x n ) be a sequence in C(S) and Xn -+ x. Then xn(t) -+ x(t)

°
for all t, and SUPPXn C S for all n. Since xn(t) 2: 0 for all t, it follows that
x(t) 2: for all t and x E C. If suppx rt S, choose to E T\S such that x(to) > o.
Thus xn(to) > 0 for n sufficiently large implies that to E SUPPXn for large n, a
contradiction to SUPPXn C S. This proves suppx C S, and thus C(S) is closed.
If x, y E C(S) and a, fJ 2: 0, then z := ax + fJy satisfies z 2: 0, and supp z C S.
Thus z E C (S), so C (S) is a convex cone.
CONSTRAINED INTERPOLATION BY POSITIVE FUNCTIONS 271

Finally, let x, y E C, 0 < >. < 1, and z := >.x+(I- >.)y E C(S). Then supp z C S.
Since x, y :0:: 0 and >., 1 >. > 0, it follows that supp xeS, supp yeS and hence
x, y E C(S). This proves that C(S) is extremal in C.
(2) This is obvious.
(3) Let D = {z E £2(T) I z(t) :s:
0 for all t E S}. If zED, then for every
x E C(S), x = 0 off S, x :0:: 0 on S, and hence

(z, x) = I: z(t)x(t) = I: z(t)x(t) :s: o.


tET tES

Thus z E C(S)O and hence D C C(S)o.


Conversely, let z E £2(T)\D. Then z(to) > 0 for some to E S. The element
Xo E £2(T) defined by xo(to) = 1 and xo(t) = 0 for t =f. to is in C(S), but (z,xo) =
z(to) > O. Thus z f/:. C(S)o. This proves C(S)O cD and verifies (3).
(4) The proof is similar to (3), or it can be deduced from (3) and the fact that
C(S)J. = C(S)O n [-C(S)O]. •
There is a simple formula for the best approximation to any x E £2 (T) from the
set C(S). Recall that the characteristic function of a set S is defined by

ift E S,
XS(t) = { ~ otherwise.

10.48 Best Approximation from C(S). Let S be a given subset ofT. Then
C(S) is a Chebyshev convex cone in £2(T), and

(10.48.1) PC(S)(x) = xs . x+ for each x E £2(T).

In particular,

(10.48.2) Pc(x) = x+ for each x E £2(T).

Proof. By Theorem 10.47, C(S) is a closed convex cone L1 the Hilbert space
£2(T), so C(S) is Chebyshev by Theorem 3.5. Given any x E £2(T), let Xo :=
XS· x+. Clearly, Xo :0:: 0 and suppXo C S, so Xo E C(S). Further, if t E S, then
x(t) -xo(t) = x(t) -x+(t) :s: 0 and X-Xo E C(S)O by Theorem 10.47(3). Moreover,

(x - xo,xo) = I:[x(t) - xo(t)]xo(t) = I:[x(t) - xo(t)]xo(t)


tET tES

= I: [x(t) - x+(t)]x+(t) = 0,
tES

since [x(t) - x+(t)]x+(t) = 0 for every t E T. We have shown that x - Xo E


C(S)O n x~. By Theorem 4.7, Xo = PC(S) (x).
The last statement follows from the first, since C = C(T) by Theorem 10.47(2) .
• The next proposition shows that the important extremal subset Cb of C is just
C(S) for a prescribed subset S of T.
272 CONSTRAINED INTERPOLATION FROM A CONVEX SET

10.49 Proposition. Let Sb := UxEK(b)SUPPX. Then

(10.49.1)

Proof. Since Sb :::) suppx for each x E K(b), it follows that K(b) C C(Sb).
Since C(Sb) is a convex extremal subset of C by Theorem 10.47(1), it follows
by Proposition 10.38 that C(Sb) :::) Cb. For the reverse inclusion, it suffices by
Proposition 10.36 to show that b E ri A( C(Sb)). But by Theorem 10.27, this is
equivalent to showing that

(10.49.2)

Since C(Sb) is a convex cone and A is linear, A(C(Sb)) is a convex cone and
bE A(K(b)) C A(C(Sb))' By Theorem 4.5(5), we get

[A(C(Sb)) - W= [A(C(Sb)W n bl..


Similarly,
[A(C(Sb)) - b]l. = [A(C(Sb))]l. n bl..
Thus (10.49.2) may be rewritten as

(10.49.3)

To verify (10.49.3), let y E [A(C(Sb)W n bl.. Then, using Theorem 10.31, we


have that

(10.49.4) (y,b) = 0

and

(10.49.5)

By Proposition 10.47(3), (A*y)(t) :S 0 for all t E Sb. Also, for any x E K(b),
Ax = b, so that from (10.49.4) we obtain

(10.49.6) 0 = (y,b) = (y,Ax) = (A*y,x) = J::(A*y)(t)x(t) = 2)A*y) (t)x(t),


tET

where we used suppx C Sb for the last equality. Since (A*y)(t) :S 0 and x(t) 2 0
for all t E Sb, (10.49.6) implies that

(10.49.7) (A*y)(t)x(t) = 0 for all t E Sb.

If (A*y)(to) < 0 for some to E Sb, choose Xo E K(b) such that xo(to) > O. Then
(A*y)(to)xo(to) < 0, which contradicts (10.49.7). Thus (A*y)(t) = 0 for all t E Sb,
and using 10.47(4), it follows that A*y E C(Sb)l., or y E (A*)-l(C(Sb)l.). By
(10.31.2), it follows that y E [A( C(Sb)W, which verifies (10.49.3). •
Now we are in position to state and prove the main result of this section.
CONSTRAINED INTERPOLATION BY POSITIVE FUNCTIONS 273

10.50 Interpolation Constrained by the Positive Cone. Let K(b) be de-


fined as in (10.46.1) and Sb := UxEK(b)SUPPX. Then for each x E PdT), there exists
y E £2 (m) that satisfies

(10.50.1) \ [x + ~ y(i)Xir XSb, Xj) = b(j) for j = 1,2, ... ,m.

Moreover,

(10.50.2)

for any y E £2(m) that satisfies (10.50.1).


Finally, the characteristic function XS b may be deleted from both (10.50.1) and
(10.50.2) if {X1XSb' X2XS b, . .. ,XmXSb} is linearly independent.
Proof. Combining Theorem 10.40, Proposition 10.49, and Theorem 10.48, we
obtain the first paragraph of the theorem. To prove the last sentence of the theorem,
suppose {X1XSb' X2XS b, ... ,XmXSb} is linearly independent. We will show that b E
riA(C). It then follows that Cb = C = C(T) and Theorems 10.40 and 10.48 show
that the characteristic function XS b may be dropped from (10.50.2).
Since C is generating (Example 10.43), it follows by Lemma 10.45 that it suffices
to show that

(10.50.3) [A(CW n bl. c N(A*).

To this end, let y E [A( CW n bl.. Using Theorem 10.31, we see that A*y E Co.
Then A*y ~ 0 by Proposition 10.47(3), so that for any x E K(b),

0= (y,b) = (y,Ax) = (A*y,x) = I)A*y)(t)x(t) = 2:)A*y) (t)xsb(t)x(t),


tET tET

since suppx C Sb for every x E K(b). Since A*y ~ 0 and x 2: 0, it follows that
(A*y)(t)XSb(t) = 0 for all t. But A*y = L:~l y(i)Xi implies that L:~l y(i)XiXSb =
O. Since {X1XS b, ... ,xmXs.} is linearly independent, we must have y(i) = 0 for all
i. That is, y = O. Hence y E N(A*) and (10.50.3) holds. •
Remark. The proof actually has shown that if {X1XS b, ... ,XmXSb} is linearly
independent, then b E intA(C). To see this, we need only note that b E intA(C) if
and only if [A(C)-W = {O} (by Theorem 10.27(2)) if and only if [A(CWnbl. = {O}
(by Theorem 4.5(5». And it was this latter condition that was established in
verifying (10.50.3).
10.51 Example. Let X = £2, let ei denote the canonical unit basis vectors

and let
K:={XE£2Ix2:0, (x,xl)=I, (X,X2)=0}.
274 CONSTRAINED INTERPOLATION FROM A CONVEX SET

What is the minimal norm element of K? That is, what is PK(O)?


Letting b = (1,0), we see that K = K(b), and Theorem 10.50 applies. We first
must compute Sb = UxEK(b)SUPPX. If x E K(b), then
(10.51.1) x(t) ::::: 0 for all tEN,
(10.51.2) x(l) + x(2) = 1,

and
1
L
00

(10.51.3) -x(n) = O.
n=3 n

Now (10.51.1) and (10.51.3) imply that


(10.51.4) x(n) = 0 for all n ::::: 3.
Properties (10.51.1) and (10.51.2) imply that
(10.51.5) x(2) = 1 - x(l), 0:<:; x(l) :<:; 1.
In particular, x = el and x = e2 are both in K(b). It follows from (10.51.4)-
(10.51.5) that
(10.51.6) Sb = U suppx = {1,2}.
XEK(b)

It follows from Theorem 10.50 that


(10.51.7)
for any choice of y = (y(l), y(2)) E £2(2) such that
(10.51.8) «(Y(l)XI + y(2)X2)+X{I,2}, Xl) = 1,

(10.51.9)
Now, (10.51.9) is automatically satisfied for any choice of y, since X2X{I,2} = O.
Thus we need only satisfy (10.51.8). For the same reason, (10.51.8) reduces to
(10.51.10) ((y(l)XI)+, Xl) = 1.
Clearly, y(l) = ~ satisfies (10.51.10) since (XI,XI) = 2. Thus
1 1
(10.51.11) PK(O) = "2 Xl = "2 (el + e2).
We should observe that although PK(O) = ~XI is the unique solution to this
problem, there are infinitely many solutions y to equations (10.51.8)-(10.51.9),
namely, y = (~, y(2)) for any y(2) E lit
There is an analogous result that holds in the space L2 [c, d], and it can be proved
in a similar way.
Let {XI,X2, ... ,xm } C L 2[c,d]' bE £2(m), and
K(b):= {x E L 2[c,dll x(t)::::: 0 for almost all t E [c,d]'
(10.51.12)
and (x, Xi) = b(i) for i = 1,2, ... ,m}.
For any x E L 2 [c, dj, its support is defined by
(10.51.13) suppx:= {t E [c,dll x(t) =1= O}.
Of course, supp x is defined only up to a set of measure zero.
CONSTRAINED INTERPOLATION BY POSITIVE FUNCTIONS 275

*10.52 Theorem. Let K(b) be as defined in (10.51.2) and let


Sb := UxEK(b)SUPPX. Then for each x E L2[c,d], there exists y E £2(m) that
satisfies

(10.52.1) ( [X + ~ y( i)Xi] + Xsb' Xj) = b(j) for j = 1,2, ... , m.

Moreover,

(10.52.2)

for any y E £2(m) that satisfies (10.52.1).


Finally, the characteristic function XS b may be deleted from both (10.52.1) and
(10.52.2) if {X1XSb,X2XSb"'" xmXsJ is linearly independent.

*10.53 Example. In the Hilbert space L2[0, 3], let

Xi(t) := (1- It - iJ)+ (i = 1,2)

and

What is the minimal norm element PK(O) of K?


Using the fact that 2X[O,1] E K, it is easy to verify that

U{suppx I x E K} = [0,1].

By Theorem 10.52,

(10.53.1)

for any choice of y = (y(l), y(2)) E £2(2) that satisfies

(10.53.2) ([y(I)Xl + y(2)X2]+X[O,1],Xl) = 1,


(10.53.3) ([y(l)Xl + y(2)X2]+X[O,1], X2) = O.

Since X2X[O,1] = 0, it follows that (10.53.3) is satisfied for any choice of y. Thus it
suffices to determine a y that satisfies (10.53.2). Since

[y(I)Xl + y(2)X2]+X[O,1] = [y(1)X1X[O,1] + y(2)X2X[O,1]]+ = [y(I)X1X[O,1]]+


= y(I)+X1X[O,1],

it follows that (10.53.2) implies that


276 CONSTRAINED INTERPOLATION FROM A CONVEX SET

SO y(l)+ = 3 = y(l). Substituting this into (10.53.1), we obtain that

(10.53.4)

By the uniqueness of PK(O), it is now clear that it is not possible to drop the
characteristic function X[O,l] from the expression (10.53.1) for PK(O).
There is an alternative approach to computing PK(b) (x) as described in Theorem
10.50. It uses Dykstra's algorithm (see Chapter 9). We describe this approach now.
Recall our setup. We are working in the Hilbert space X = f!2(T) (or, more
generally, in X = L2 (T, ft), where ft is any measure on T). Let {Xl, X2, ... , x m } C
X\{O}, C = {X E X I x(t) ~ 0 for all t E T}, and define A: X -+ f!2(m) by

Ax := ((x, Xl), (x, X2)'" ., (x, x m )), x E X.

For any b E A(C), set

K(b) := CnA-l(b) = {x E C I (x, Xi) = b(i) for i = 1,2, ... ,m} = n


m+l
Ki,

where
Ki := {x E X I (x, Xi) = b(i)}
is a hyperplane for i = 1,2, ... , m, and K m + l = C.
As we saw in Example 9.44, Dykstra's algorithm substantially simplifies in this
application. Indeed, for any x E X, we have that

(10.53.5)

by (10.48.2), and

1 .
(10.53.6) PKi(x)=x-lixiI12 [(x,xi)-b(z)]Xi

for i = 1,2, ... ,m by (6.17.1). Fix any Z E X and set

[n] := {I, 2, ... , m + I} n {n - k(m + 1) I k = 0, 1,2, ... },

Zo = z, eo = 0,

(10.53.7) if [n] # m+ 1,

and

(10.53.8)

and

(10.53.9)

if [n] = m + 1 (n = 1,2, ... ). Then the Boyle-Dykstra theorem (Theorem 9.24)


implies that the sequence {zn} converges to PK(b) (z): Ilzn - PK(b)(z)11 -+ O.
EXERCISES 277

For the practical implementation of these equations, it is convenient to rewrite


them in the following form. Set

(10.53.10) Zo = z, eo = 0,
and for n ~
(10.53.11)
°
compute

1
Z(m+l)n+l = Z(m+1)n - IIxll12 [(Z(m+l)n, Xl) - b(l)] Xl,

1
z(m+1)n+2 = Z(m+l)n+l - IIx2112 [(Z(m+1)n+1' X2) - b(2)] X2,

1
Z(m+1)n+m = z(m+1)n+m-1 - Il x ml!2 [(Z(m+l)n+m-l, xm) - b(m)] Xm ,

(10.53.12)

(10.53.13) e(m+l)(n+l) = Z(m+l)n+m - Z(m+l)(n+l) + e(m+l)n-


Exercises

(1) 1. Let S = {p(a,,6, 1) I p ~ 0, a 2 +,62 <::: 1} and T = {(a, -1, 1) I a E JR.}.


Verify the following statements.
(a) S is a closed convex cone and T is a closed affine set.
(b) (SnT)O 1= So +To.
(c) con (S n T) = con (S) n con (T). Why does this example not contradict
Lemma 10.37
2. Let X and Y be inner product spaces, A : X --+ Y be linear, and suppose
that KeY is a convex set (respectively a convex cone, a subspace). Then
the inverse image A-I(K) is a convex set (respectively a convex cone, a
subspace).
3. If C and D are closed convex cones in the Hilbert space X, show that C c D
if and only if Co ~ DO. More generally, show that for any two sets Sand
T in X, con (S) C con (T) if and only if So ~ TO.
4. (a) If {Ci } is any collection of convex cones, show that

In other words, the conical hull of the union is the same as the convex hull
of the union.
(b) Show that if {Cd is any collection of complete convex cones, then
278 CONSTRAINED INTERPOLATION FROM A CONVEX SET

[Hint: Part (a).]


5. Let K be a nonempty convex set in X and Xo EX. Prove the equivalence
of the following three statements.
(a) Xo E K;
(b) (x,xo)::; sup(x,y);
yEK

(c) (K - xo)O = {x E X I sup(x,y) = (x,xo)}.


yEK
6. Let Sand T be convex subsets of the Hilbert space X with 0 E SnT. Show
that the following four statements are equivalent.
(a) (S n T)O = (So + TO);
(b) (S n T)O c (So + TO);
(c) con (S) n con (T) c ---'-co-n--;(=S-n-=T"');
(d) con (S) n con (T) = con (S n T).
[Hint: Look at the proof of Lemma 10.3.]
7. One consequence of Theorem 4.6(4) is that if C 1 and C 2 are closed convex
cones in the Hilbert space X, then

(7.1)

Show that (7.1) fails in general if the convex cones C i are not closed.
[Hint: Consider the sets C 1 = {(a,;3) I a =;3 = 0 or a > 0 and;3 2: O} and
C2 = {(a,;3) I a =;3 = 0 or a < 0 and;3 2: O} in £2(2).] This proves that
the equivalent conditions of Exercise 6 do not always hold.
8. (a) If C is a convex set and a and ;3 are positive real numbers, then

(a + ;3)C = aC + ;3C.

(b) If C 1 and C2 are convex sets and a E R, then

9. If K is a convex subset of X, a function f :K ~ R U {oo} is called convex


if
f(>-.x + (1 - >-.)y) ::; )..j(x) + (1 - >-.)f(y)
for all x, y E K and 0 < >-. < L
(a) Show that f(x) := Ilxll is convex on X.
(b) If {Ij I j E J} is an indexed collection of convex functions on K, show
that f := SUPjEJ Ii is convex on K. (Here f(x) := sup{Jj(x) I j E J}, x E
K.)
(c) Show that f(x) := SUPYEK(X, y) is convex on X. (f is called the support
function of K.)
(d) Every linear functional on X is a convex function.
(e) The function

if x E K,
f(x) = { :
if x tJ. K,
EXERCISES 279

is convex on X. (J is called the indicator function of K.)


(f) If J : K -+ lRU{ oo} is convex on K, then the function F : X -+ lRU{ oo},
defined by
if X E K,
F(x):= { ~X)
ifx r:J.K,
is also convex. In other words, every convex Junction has an extension to a
convex function defined on the whole space.
10. Let {Cb C2, ... , Cm} be closed convex subsets of the Hilbert space X and
suppose L:~ C i =1= 0. Verify that {Cb C2, ... , Cm} has the strong CHIP if
and only if

(10.1)

for every x E nrCi , where Ie denotes the indicator function for C:


if x E C,
Ie(x):= { :
if X r:J. C,
and where aJ denotes the subdifferential of J:

aJ(x) := {x* E X I (x*, z - x) + J(x) ~ J(z) for all z EX}.

11. Let X = f2(T) (or, more generally, X = L2(T, /-L», {Xl. X2, ... , xm} C X,
and define A on X by

xEX.

Prove that A is surjective (Le., A(X) = f2(m» if and only if {XI,X2, ... , xm}
is linearly independent.
[Hint: Use Lemma 8.33 to establish first that A is surjective if and only if
A * is injective.]
12. Prove Lemma 10.19 concerning affine hulls.
[Hint: Lemma 10.17.]
13. Prove Theorem 10.23, which shows, in particular, that the relative interior
of a convex set is convex.
[Hint: Theorem 2.8 and Lemma 10.22.]
14. Prove that a nonempty set S has dimension n if and only if there exists a set
ofn+l points {XO, Xl, ... ,xn } in S such that {XI-XO,X2-XO, ... ,xn-xo} is
linearly independent, and no larger number of points in S has this property.
15. Let Ci be a convex subset of X with ri Ci =1= 0 for i = 1,2, ... , n (for example,
if dim X < 00 and C i =1= 0 for each i). Prove that

16. Let X and Y be inner product spaces, A : X -+ Y be linear, and C C X be


convex with ri C =1= 0. Show that

riA(C) = A(riC).
280 CONSTRAINED INTERPOLATION FROM A CONVEX SET

17. If {el, e2, ... , en} is an orthonormal basis for Y and r > 1/Vii, prove that

(This shows that the radius II Vii is largest possible in Theorem 10.30.)
[Hint: Choose any p with lin < p < rlVii and show that the vector
x = p L~ ei is in B(O, r), but not in co {±el, ±e2,· .. , ±e n }.]
18. Let 0 be a convex subset of X. Show that:
(a) 0 is extremal in O.
(b) The intersection of any collection of extremal subsets of 0 is either
empty or extremal.
(c) The union of any collection of extremal subsets of C is an extremal
subset of C.
(d) If El is extremal in C and E2 is extremal in E l , then E2 is extremal in
O.
19. If E is a convex extremal subset of the convex cone C and ri En ri C 0/= 0,
then E = C.
[Hint: Proposition 10.46.]
20. Let K be a convex set and x E K. Prove the equivalence of the following
three statements.
(a) x is an extreme point of K;
(b) y, z in K and x = ~(y + z) implies that y = z;
(c) K\ {x} is convex.
21. Let 0 be a convex cone in X. Prove that 0 is generating in X (Le., spanC =
X) if and only if 0 - C = X.
22. Consider the mapping A: £2(2) -t £2(2) defined by A((o:,,B)) = (0,,8), and
let

Verify the following statements.


(a) C is a closed convex set.
(b) A is a self-adjoint bounded linear operator with IIAII = 1.
(c) N(A) = {(0:,0) 10: E R}.
(d) A(C) = {(0,j3) 11J3I <S; I}.
(e) If b = ±(O, 1) and K(b) := C n A-l(b), then K(b)= {b} and C b = {b}.
(f) If b = (0, j3) with -1 < j3 < 1, then Cb = O.
(g) If b = (0,1), then
(i) (C - b) nN(A) = {O}.
(ii) con (C - b) nN(A) = N(A).
(iii) con[(C - b) nN(A)] = {O} 0/= con (0 - b) nN(A).
(h) Theorem 10.40 fails in general if Ob is replaced by C in (10.40.1) and
(10.40.2).
[Hint: Lemma 10.12 and part (g).]
*23. Let X = L2[0, 3], Xi(t) = [1 - It - il]+ (i = 1,2), and

K={xEXlx~O, (X,xll=l, (x,x21=0}.

Find PK(x), when x(t) = sin (~t).


[Hint: Theorem 10.52.]
EXERCISES 281

*24. Let X = L2[0, 1], Xl = X[O,l]' X2 = L~=l nX[2-n,2-n+'], and

K={XEXlx;::O, (x,xl)=I, (x,x2)=I}.

Find PK(X), when x(t) = sin (~t).


[Hint: Theorem 10.52.]
*25. Fix any integer m ;:: 1 and let X = L2 [0, m + 1]. Consider the functions
{Xl,X2, ... ,Xm } in X defined by

for i = 1,2, ... ,m. Let C:= {x E X I x;:: O} and define A: X -+ £2(m) by

Ax := (x, Xl)' (x, X2)'" ., (x, x m )), x E X.

Verify the following statements.


(a) {Xl, X2, ... , x m } is a linearly independent set of nonnegative functions.
(b) A(C) = {(al,a2, ... ,am ) I ai;:: 0 for all i}.
(c) riA(C) = int A(C) = {(aI, 002, .. . , am) I ai > 0 for all i}.
(d) bd A(C) = {(aI, 002, ... , am) I ai ;:: 0 for all i and ai = 0 for some i}.
*26. With the notation of Exercise 25, b = (1, E) for 0 < E < 1, and K(b) :=
C n A-l(b), compute PK(b) (0), and compare this with PK(b) (0) when E = 0
or 1.
*27. Let X = LdO,3], C = {x E X I x;:: O}, Xi(t) = [I-It - il]+ (i = 1,2), and
define A on X by

Ax := (x, Xl), (x, X2)), x E X.

Verify the following statements.


(a) A is a bounded linear operator from X onto £2(2).
(b) A* : £2(2) -+ X is given by

A*y = y(1)Xl + y(2)X2, Y = (y(I), y(2)) E £2(2).

(c) A(C) = A(C) = {(aI, (02) I ai ;:: 0 (i = 1,2)}.


(d) riA(C) = intA(C) = {(aI, (02) I ai > 0 (i = 1,2)}.
(e) bdA(C) = {(aI, (02) I ai;:: 0 (i = 1,2) and ai = 0 for some i}.
*28. Assume the hypothesis of Exercise 27, let b = (1,0), and let K = K(b) =
C n A-l(b). Verify that if x E X satisfies x;:: 0 and fol tx(t) dt:S: 1, then

where Al := 3 [1- fol tx(t)dt].


[Hint: Theorem 10.50.]
Using this fact, verify the following formulas:
(a) PK(O)(t) = 3tX[0,1] (t).
(b) PK(tO!)(t) = [to! + 3 (~t~) t] X[O,l](t).
282 CONSTRAINED INTERPOLATION FROM A CONVEX SET

(c) PK(et)(t) = etX[O,lj(t).

(d) PK(e t2)(t) = [2


et + "2(3
3 - e)t] X[O,lj(t).
*29. Assume the hypothesis of Exercise 27, let b = (1,1), and K = K(b) =
C n A-I(b). Verify that

[Hint: Show that b E int A( C) and apply Theorem 10.50.]


*30. Let X = L 2 [0, 1], C = {x E X I x 2 O},

L n X[2-
00

Xl = X[O,lj, X2 = n ,2- n +1j,


n=l

and define A on X by

Verify the following statements.


(a) A is a bounded linear operator from X onto £2(2).
(b) A* : £2(S) -+ X is given by

A*y = y(1)XI + y(2)X2' Y = (y(1), y(2)) E £2(2).

(c) A(C) = {(al,a2) I al = a2 = 0 or 0 < al:S ad·


(d) A(C) = {(al,Q2) 10 < Ql :S Q2}' (In particular, A(C) is not closed.)
(e) riA(C) = intA(C) = {(al,a2) 10 < al < a2}.
(f) bdA(C) = {(al,a2) 10 = al:S a2 or 0 < Ql = a2}.
*31. Assume the hypothesis of Exercise 30, let b = (1,1), and K = C n A-l(b).
If x E X satisfies

then show that

where

[Hint: Theorem 10.50.]


Using this fact, verify the following formulas.
(a) PK(O) = 2Xr!,lj'
(b) PK(t<') (t) = [ta + 12;:;(1~:)1] X[!,lj(t) for any a 2 O.
(c) PK(e t ) (t) = let + 2(1- e + ye))X[!,lj(t).
HISTORICAL NOTES 283

*32. In Example 10.53, it was shown that if Xi(t) = [1 -It - il]+ (i = 1,2) and

K = {x E L2[0, 3] I x 2': 0, (X,XI) = 1, (X,X2) = O},

then

In this exercise you are asked to use Dykstra's algorithm (see the descrip-
tion following Example 10.53) to "compute" PK(O). Specifically, what is
the result of applying Dykstra's algorithm to this problem after 10 itera-
tions, 100 iterations, and 1000 iterations? (Here, of course, we are assuming
that you have some familiarity with software like Mathematica or Maple to
automate these computations.)

Historical Notes
The main problem studied in this chapter (see Example 10.1) evolved from the
following "shape-preserving interpolation" problem. In the Sobolev space of real
functions f on [a, b] having absolutely continuous (k - l)st derivative f(k-I) with
f(k) in L2 [a, b], find a solution of the problem to minimize

{l b
If(k)(tW dt I f(k) 2': 0, f(td = Ci (i = 1,2, ... , n + k)},

where ti E [a, b] and Ci E IR are prescribed.


Numerous authors have studied this problem. For a survey of results up to 1986,
see Ward (1986). Using an integration by parts technique, Favard (1940) and, more
generally, de Boor (1976) showed that this probem has an equivalent reformulation
as follows: Minimize

{llylll y E L2[a,b], y 2': 0, (y,Xi) = bi (i = 1,2, ... ,n)},

where the functions Xi E L2 [a, b] and the numbers bi E IR can be expressed in terms
of the original data ti and Ci'
The fundamental property, the strong CHIP, was introduced by Deutsch, Li, and
Ward (1997) as a strengthening of the CHIP that had been studied earlier by Chui,
Deutsch, and Ward (1990). Lemma 10.3, Examples lO.7, 10.9, and 10.10, Lemma
10.12, and Theorem 10.13 are from Deutsch, Li, and Ward (1997). Theorems
10.23 and 10.24 are taken from Rockafellar (1970; Theorems 6.1 and 6.2). The
separation theorem (Theorem 10.25) is a special case of a result of Eidelheit (1936)
who generalized a result of Mazur (1933; p. 73) (who had proved the special case
where one of the convex sets is a singleton). The equivalence of (1)-(3) in Theorem
10.26 is from Rockafellar (1970; Theorem 6.4), while the equivalence of (1) and (4) is
from Deutsch, Li, Ward (1997; Lemma 3.9). Lemma lO.28(1) (respectively Lemma
10.28(2)) is from Rockafellar (1970; Theorem 6.3) (respectively (1970; Corollary
6.3.1)). Theorem 10.29 is from Deutsch, Li, and Ward (1997; Theorem 3.11).
Theorem 10.31 on dual cones of images can be found in Luenberger (1969). Theorem
10.32 is from Deutsch, Li, and Ward (1997; proof of Theorem 3.12). Special cases
of Theorem 10.32 were known earlier. For example, when C is a closed convex cone,
284 CONSTRAINED INTERPOLATION FROM A CONVEX SET

bE intA(C), and x = 0, see Micchelli and Utreras (1988). Also, when C is a closed
convex cone, see Chui, Deutsch, and Ward (1992; Lemma 2.1).
The notion of an extreme point is due to Krein and Milman (1940), who proved
the important result that any compact convex subset of a locally convex linear
topological space is the closed convex hull of its set of extreme points. This result
has many far-reaching ramifications.
Proposition 10.36 was suggested by Wu Li (1998; unpublished). The original
definition of Cb given in Deutsch, Li, and Ward (1997) (and in Chui, Deutsch, and
Ward (1992) for the case where C is a convex cone) was that C b was the smallest
convex extremal subset of C that contains C n A-l(b). Proposition 10.38 shows
that the two definitions of Cb agree. The perturbation theorem (Theorem 10.40)
(respectively Proposition 10.39) is from Deutsch, Li, and Ward (1997; Theorem 4.5
(respectively Proposition 4.6)).
The notion of a generating set in a Hilbert space, as well as Lemma 10.44, was
given in Deutsch, Ubhaya, Ward, and Xu (1996).
The section on constrained interpolation by positive functions essentially follows
Deutsch, Ubhaya, Ward, and Xu (1996) (who worked in the space L 2 [a, b]). The
main results of that section, Theorems 10.50 and 10.52, are both special cases of
the following theorem.
10.54 Theorem. Let (T,S,/-l) be a measure space and L 2 (/-l) = L 2 (T,S,/-l).
Let C = {x E L 2(/-l) I x ?: 0 a.e. (/-l)}, {Xl, ... ,X m } C L 2(/-l) \ {O}, b E £2(m),
Ax:= ((X,Xl), ... ,(x,xm )) for all x E L 2(/-l), K(b):= CnA-l(b) =1= 0, and Sb:=
U{suppx I x E K(b)}. Then, for every x E L 2 (/-l), there exists y E £2(m) that
satisfies

(10.54.1) ( [x + ~ y(i)xi] + XS b , Xj) = b(j) (j = 1, ... ,m).

Moreover,

(10.54.2)

for any y E £2(m) that satisfies (10.54.1). Finally, the characteristic function XS b
may be deleted from both (10.54.1) and (10.54.2) if {XlXSb'.·.' xmXsJ is linearly
independent.
Theorem 10.52, excluding its last sentence, was first proved by Micchelli, Smith,
Swetits, and Ward (1985) using variational methods. It was probably this paper,
more than any other, that motivated the general study presented in this chapter.
Other work related to the Micchelli, Smith, Swetits, and Ward paper (1985) include
Illiev and Pollul (1984), (1984a), Dontchev, Illiev, and Konstantinov (1985), Irvine
(1985), Irvine, Marin, and Smith (1986), Borwein and Wolkowicz (1986), Smith and
Wolkowicz (1986), Anderson and Ivert (1987), Dontchev (1987), Smith and Ward
(1988), Swetits, Weinstein, and Xu (1990), and Li, Pardalos, and Han (1992). A
related question as to the feasibility of the problem, i.e., when is C n A-I (b) =1= 0,
was studied by Mulansky and Neamtu (1998).
The problem of minimizing {llxlll x E C, Ax = b} can be easily cast as a certain
optimization problem whose "optimal"solution can be studied via Fenchel dual-
ity. This was done by Rockafellar (1967), Massam (1979), Borwein and Wolkowicz
HISTORICAL NOTES 285

(1981), Gowda and Teboulle (1990), and Borwein and Lewis (1992), (1992a). In-
terpolation problems have also been considered from a control theory perspective,
see Dontchev, Illiev, and Konstantinov (1985) and Dontchev (1987). A practical
motivation for the study of such constrained problems arises in connection with L2
spectral estimation (see Goodrich and Steinhardt (1986)).
Some of the results of this chapter have been extended and generalized. For
example, results analogous to Theorem 10.52, in the case where C is the cone of
increasing functions or the cone of convex functions, was given in Chui, Deutsch,
and Ward (1992); more generally, results analogous to Theorem 10.52 when C is
the cone of n-convex functions were established in Deutsch, Ubhaya, and Xu (1995)
and Deutsch, Ubhaya, Ward, and Xu (1996). In this situation, it is necesary to
know how to recognize best approximations from C. One noteworthy result along
these lines is the following.
10.55 Theorem. Let C denote the set of nondecreasing functions on [a, b],
f E L 2 [a, b], ftt C, and 9 E C. Then 9 = Pc(f) if and only if 9 is equal

J:
(almost everywhere on [a, bJ) to the derivative of the greatest convex minorant of
the indefinite integral s >--+ f(t) dt of f·
Recall that the greatest convex minomnt k of a function h is the largest convex
function bounded above by h; that is,

k(t) := sup{ c(t) I c convex on [a, b], c ~ h} for all t E [a, b].

This result was established by Reid (1968) (in the case that f is bounded) using
methods of optimal control, and by Deutsch, Ubhaya, and Xu (1995; Theorem 5.2)
in the general case by the methods of this chapter.
Deutsch, Li, and Ward (2000) extended the main results of this chapter to in-
clude inequality constraints as well as equality ones. That is, they considered the
approximating set to be of the form

K = C n {x E X I Ax ~ b}.

Li and Swetits (1997) developed an algorithm to handle inequality constraints in


finite-dimensional spaces.
Deutsch (1998) showed that the strong CHIP is the precise condition that al-
lowed a Karush-Kuhn-Tucker-type characterization of optimal solutions in convex
optimization. Bauschke, Borwein, and Li (1999) studied the relationship between
the strong CHIP, various kinds of "regularity," and Jameson's property (G).
It should be mentioned that while the Dykstra algorithm can be used for com-
puting PK(b)(X), there are descent methods for directly solving the equation

A[Pc(x + A*y)] = b
for y E £2(m), and hence for computing PK(b)(X) = Pc(x + A*y) when C is a
polyhedron; see Deutsch, Li, and Ward (1997; section 6). Moreover, algorithms
that apply to the problem considered here when C is a convex cone were developed
in Irvine, Marin, and Smith (1986) and Micchelli and Utreras (1988).
CHAPTER 11

INTERPOLATION AND APPROXIMATION

Interpolation
In this chapter we first consider the general problem of finite interpolation. Then
we study the related problems of simultaneous approximation and interpolation
(SAl), simultaneous approximation and norm-preservation (SAN), simultaneous
interpolation and norm-preservation (SIN), and simultaneous approximation and
interpolation with norm-preservation (SAIN).
It is a well known basic fact in analysis that it is always possible to find a (unique)
polynomial of degree n - 1 to interpolate any prescribed set of n real numbers at n
prescribed points (see Corollary 11.4 below). That is, if {tl, t2, ... , tn} is a set of n
distinct real numbers and {CI' C2, ... ,cn } is any set of n real numbers, then there
is a unique polynomial y E P n - l such that y(t i ) = Ci (i = 1,2, ... , n).
Setting a = miniti' b = maxiti, X = C 2 [a,b] nPn - l , and defining the point
evaluation functionals xi on X by

xEX,

we see that X is n-dimensional, xi E X*, and the problem of polynomial interpo-


lation can be reworded as follows: For any given set of n real numbers Ci, there
exists a unique y E X such that

x;(y) = Ci (i = 1,2, ... ,n).

Thus polynomial interpolation is a special case of the general problem of finite


interpolation: Let M be an n-dimensional subspace of the inner product space X
and let {xi, x 2, ... , x~} be a set of n (not necessarily continuous) linear functionals
on X. Given any set of n scalars {CI, C2, ... , cn } in JR, determine whether there
exists an element y E M such that

(i = 1,2, ... ,n).

The next theorem governs this situation. But first we state a definition.
11.1 Definition. Let x, y in X and suppose r is a set of linear functionals on
X. We say that y interpolates x relative to r if

(11.1.1) x*(y) = x*(x) for every x* E r.

Note that if r = {xi, X2' ... ,x~}, then y interpolates x relative to r if and only if
y interpolates x relative to span r. Now we describe when interpolation can always
be carried out.

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
288 INTERPOLATION AND APPROXIMATION

11.2 Characterization of When Finite Interpolation Is Possible. Let M


be an n-dimensional subspace ofthe inner product space X and let xi, X2' ... ,x~ be
n (not necessarily bounded) linear functionals on X. Then the following statements
are equivalent:
(1) For each set of n scalars {C1' C2, ... , cn } in JR, there exists y E M such that

(11.2.1) xi(Y) = Ci (i=1,2, ... ,n);

(2) For each set of n scalars {C1, C2, .. . , cn } in JR, there exists a unique y E M
such that (11.2.1) holds;
(3) For each x E X, there exists a unique Y E M that interpolates x relative to
r:= {xi,x 2, ... ,x~}; i.e.,
(11.2.2) (i = 1,2, ... ,n);

(4) The only solution y E M to the equations

(11.2.3) (i=1,2, ... ,n)

is Y = 0;
(5) For every basis {Xl, X2, . .. , x n } of M,

(11.2.4)

(6) For some basis {X1,X2, ... ,xn } of M, (11.2.4) holds;


(7) There exists a set {Xl, X2, ... , xn} in M (necessarily a basis) such that

(11.2.5) (i,j=1,2, ... ,n);

(8) The set of restrictions {xiIM, x2IM, ... , X~IM} is linearly independent (i.e.,
ifL~ !Xixi(y) = 0 for all y E M, then !Xi = 0 for i = 1,2, ... , n).
Proof. (1) ==? (2). Suppose (1) holds but (2) fails. Then there exist scalars
C1, C2, ... , Cn and distinct elements yl, Y2 in M such that xi (Yj) = C; (i = 1, 2, ... , n)
for j = 1,2. Then the nonzero vector y = Y1 - Y2 is in M and satisfies xi(Y) = 0
(i = 1,2, ... , n). Letting {Xl, X2, . .. , xn} be a basis for M, it follows that y =
L~ !XjXj for some aj, not all zero, and
n
~:>jxi(xj) = 0 (i=1,2, ... ,n).
j=l

It follows that det [xi(Xj)] = 0, and hence that there exist scalars d 1,d2, ... ,dn
such that the system of equations

(i=1,2, ... ,n)

fails to have any solution ((31, (32, ... , (3n). That is, there is no element z E M such
that x;(z) = d i (i = 1,2, ... , n). This contradicts (1), so that (2) must hold.
(2) ==? (3). This follows by taking Ci = xi(x) (i = 1,2, ... , n).
INTERPOLATION 289

(3) ==} (4). This follows by taking x = 0 in (3).


(4) ==} (5). Let {Xl, X2, ... , xn} be a basis of M. If (5) fails, then det [xi(xj)] =
0, and the homogeneous system of equations '£';'=1 ajxi(xj) = 0 (i = 1,2, ... , n)
has a nontrivial solution (a1, a2, ... , an). Thus the element Y = '£~ aiXi is in M,
Y of 0, and xi(y) = 0 (i = 1,2, ... , n). This proves that the equations (11.2.3) have
a nonzero solution in M, so that (4) fails.
(5) ==} (6). This is obvious.
(6) ==} (7). Let {Xl, X2, ... , xn} be a basis for M such that (11.2.4) holds. Then,
for each k E {I, 2, ... , n}, the equations

n
Lakjxi(xj) = bik (i=1,2, ... ,n)
j=l

have a unique solution (ak1, ak2,···, akn). Setting Yk = ,£7=1 akjXj, it follows
that Yk E M and
(i=1,2, ... ,n).
To see that {Y1, Y2, . .. , Yn} is a basis for M, it suffices to show that {Y1, Y2,···, Yn}
is linearly independent. If '£~ aiYi = 0, then for j = 1,2, ... , n,

(7) ==} (8). Let {Xl, X2, . .. , xn} be a set in M such that (11.2.5) holds. If
'£~ aixi (y) = 0 for every Y EM, then in particular,

n
0= Laixi(xj) = aj (j = 1,2, ... ,n).
i=l

Thus {xi 1M, X21 M, ... , X~ 1M} is linearly independent.


(8) ==} (1). Let {Xl, X2, ... , xn} be a basis for M. If {xiIM, x 2IM, ... , X~IM} is
linearly independent, then the only solution to the system of equations

n
Laixi(xj) = 0 (j=1,2, ... ,n)
i=l

is (a1, a2, ... , an) = (0,0, ... ,0). Hence det [xi(Xj)] of O. Thus if{ C1, C2, ... , cn} C
JR., there is a (unique) solution (/31, /32, ... , /3n) to the system of equations

n
L/3jxi(Xj) = Ci (i=1,2, ... ,n).
i=l

Then Y = '£~ /3jXj E M satisfies (11.2.1). •


In particular, if X is n-dimensional, the following corollary is an immediate
consequence of Theorem 11.2.
290 INTERPOLATION AND APPROXIMATION

11.3 Corollary. If X is an n-dimensional inner product space and xi, X2'" ., x~


are n linear functionals on X, then the following statements are equivalent:
(1) For each set of n scalars {CI, C2, ... , en} in lR, there exists an x E X such that

(11.3.1) xi(x) = Ci (i=1,2, ... ,n);

(2) For each set of n scalars {CI, C2, ., . , cn} in lR, there exists a unique x E X
such that (11.3.1) holds;
(3) The only solution x E X of the equations

(11.3.2) (i=1,2, ... ,n)

is x = 0;
(4) For every basis {Xl, X2, ... , xn} of X,

(11.3.3)

(5) For some basis {Xl, X2, ... , Xn} of X, (11.3.3) holds;
(6) There exists a set {XI, X2, ... , xn} in X (necessarily a basis) such that

(11.3.4) (i,j=1,2, ... ,n);

(7) {xi, x2, ... , x~} is linearly independent.


As one application of this corollary, we prove the statement about polynomial
interpolation made at the beginning of the chapter.
11.4 Polynomial Interpolation. Let t l , t 2, .. . , tn be n distinct real numbers
and {CI, C2, ... , cn } be any set of n real numbers. Then there is a unique polynomial
p E P n - l of degree at most n - 1 such that

(11.4.1 ) P(ti) = Ci (i=1,2, ... ,n).

In particular, if two polynomials in Pn-l coincide at n distinct points, they must


be identical.
Proof. Let [a, b] be any finite interval in lR containing all the ti's and let
X = C2 [a, b] n P n - l . Then X is n-dimensional. Define n linear functionals xi on
Xby
Xi(x) := X(ti) (i=1,2, ... ,n).
By Corollary 11.3, it suffices to show that X contains a set {Xl,X2, ... ,Xn } such
that Xi (tj) =8ij (i,j=1,2, ... ,n). Set

Xi(t) : = rr
n (t-t·)
( J)
t - t·J
j=l'
(i=1,2, ... ,n).
Hi

Clearly, Xi E P n- l and Xi(t j ) = 8ij . •


We should observe that interpolation can fail even if "the number of unknowns
equals the number of conditions."
INTERPOLATION 291

11.5 Example. Let X = C2 [-I, 1] and M = span{xI,x2}, where XI(t) = 1


ami X2(t) = t 2. Then, for any given pair of distinct points t l , t2 in [-1,1], define
the linear functionals xi on X by xi(x) = X(ti) (i = 1,2). Clearly,

In particular, det[xi(xj)) = 0 if t2 = -tl, and so by Theorem 11.2, interpolation


fails in general for this case. Specifically, if CI and C2 are distinct real scalars, it is
not possible to choose coefficients Ci!l, Ci!2 such that the polynomial p( t) = Ci!l + Ci!2 t 2
satisfies P(tl) = CI and p( -tr) = C2. This is also obvious from the fact that no
matter how the coefficients Ci!i are chosen, p(t) = Ci!l + Ci!2 t2 is an even function and
hence p( -t) = p(t) for all t.
If any of the equivalent statements of Theorem 11.2 hold, then (by Theorem
11.2(8)) the set {Xi,x2, ... ,x;;J must be linearly independent. There is an essential
converse to this fact. Namely, if {xi, x2, ... , x~} is a linearly independent set of
linear functionals on X, then there is an n-dimensional subspace M of X such that
interpolation can be carried out in M. That is, all the statements of Theorem 11.2
are r91id. This result was established earlier in Theorem 6.36, and we state it here
again for convenience.
11.6 Interpolation Theorem. Let {xi, X2, ... , x~} be a linearly independent
set of linear functionals on X. Then there exists a set {Xl, ... , x n } in X such that

(11.6.1) (i,j=I,2, ... ,n).

In particular, for each set of n real scalars {Cl, C2, ... , en}, the element y
I:~ CiXi satisfies

(11.6.2) (i = 1,2, ... , n).

11. 7 Definition. Let Xl, ... , Xn be n elements in X and xi, ... , x~ be n linear
functionals on X. The pair of sets {Xl, X2, ... , xn} and {xi, X2' ... , x~} is called
biorthogonal if

(11. 7.1) (i,j=I,2, ... ,n).

Theorem 11.6 shows that if {xi, x2, ... , x~} is linearly independent, there exists
a set {Xl, X2, ... , xn} in X such that the pair of sets is biorthogonal.
The usefulness of biorthogonal sets stems from the fact (Theorem 11.6) that one
can easily construct solutions to interpolation problems. That is, if {Xl, ... , xn}
and {xi, ... , x~} are biorthogonal, then the linear combination y = L~ CiXi solves
the interpolation problem

(i=I,2, ... ,n).


292 INTERPOLATION AND APPROXIMATION

11.8 Examples of Biorthogonal Sets.


(1) (Lagrange interpolation) Let a < tl < t2 < ... < tn :s:: b, wet)
(t - tl)(t - t2) ... (t - t n ), and

(j = 1,2, ... ,n).

Then Xj E Pn - 1, and defining xi on C2 [a, b] by

(i=I,2, ... ,n),

we see that
(i,j=I,2, ... ,n).
That is, {Xl,X2, ... ,xn } and {xi,x~, ... ,x~} are biorthogonal. Thus the polyno-
mial p = ~~ CiXi E P n - 1 satisfies P(ti) = Ci (i = 1,2, ... , n).
(2) (Taylor interpolation) In the inner product space X = {x E C2 [-I, 1] I
x(n-l) exists on [-1, Ij}, consider the polynomials Xi(t) = ti-l/(i - I)! and linear
functionals xi on X defined by xi(x) := x(i-l)(O) (i = 1,2, ... ,n). It is easy to
verify that the sets {Xl,X2, ... ,Xn } and {xi,x2""'x~} are biorthogonal. Hence
the polynomial p = ~~ CiXi E Pn - l satisfies

(i=I,2, ... ,n).

Additional examples of biorthogonal sets are given in the exercises (namely,


Exercises 11.14.5-11.14.7).

Simultaneous Approximation and Interpolation


If Y and Z are subsets of X, we recall that Y is dense in Z if for each z E Z
and E > 0, there exists y E Y such that liz - yll < E. That is, Y ~ Z.
In modeling the behavior of a system, we often try to approximate the system
by a simpler one that preserves certain characteristics of the original system. For
example, the next result shows that it is always possible to simultaneously approx-
imate and interpolate from a dense subspace.
11.9 Simultaneous Approximation and Interpolation (SAl). Let Y be
a dense subspace of X and let {xi, x 2 , ... , x~}
C X*. Then, for every x E X and
E > 0, there exists y E Y such that

(11.9.1) Ilx - yll < E


and

(11.9.2) x;(y) = x;(x) (i=I,2, ... ,n).

Proof. It is no loss of generality to assume that {xi, X2, ... , x~} is linearly
independent. Further, we may even assume that the set of restrictions
{xi Iy, x2ly, ... , x~ Iy} is linearly independent. [For if ~~ ClOiXi (y) = 0 for every
SIMULTANEOUS APPROXIMATION AND INTERPOLATION 293

Y E Y, then since Y is dense in X, 2::~ QiXi(X) = 0 for all x E X; that is,


2::~ QiXi = O. By linear independence of the xi's, Q1 = Q2 = ... = Q n = 0.] The
linear functionals Yi = xi Iy (i = 1, 2, ... , n) thus form a linearly independent set
in Y*. By Theorem 11.6, there exists a set {Xl, X2, ... , xn} in Y such that

Let c = 2::~ Ilx; II Ilxj I and E > 0 be given. Choose YI E Y such that Ilx - YIII <
E(1 + c)-I. Set Y2 = 2::~ X; (x - YI)Xj and Y = YI + Y2. Then Y E Y,

(i=I,2, ... ,n),

and
n
Ilx - yll :e::: Ilx - yIiI + IIY211 < E(1 + c)-l + 2) x; (x - YI)IIIXjll
I
n
:e::: E(1 + C)-l + E(1 + C)-l IJx; II IIXj II = E. •

This theorem implies that any result that concludes that a certain subspace is
dense in X is actually a result about simultaneous approximation and interpolation,
and not just approximation alone. As an application, we consider the following.

11.10 Corollary (Approximation by polynomials with matching mo-


ments). Let XI,X2, ... ,Xn be in C 2[a,b]. Then for each x E C 2[a,b] and E > 0,
there exists a polynomial p such that

(11.10.1) Ilx-pll < E


and

(11.10.2) lb p(t)xi(t)dt = lb x(t)xi(t)dt (i = 1,2, ... , n).

Proof. Let X = C2 [a, b], Y be the subspace of all polynomials, and define xi
onX by

xi(z) := lb z(t)xi(t)dt (i=I,2, ... ,n).

Then Y is dense by the Weierstrass approximation theorem (Corollary 7.20), and


xi E X* by Theorem 5.1S. The result now follows by Theorem 11.9. •

I:
In some problems (see, e.g., Problem 4 of Chapter 1), the energy of a system
is proportional to x2(t)dt for a certain function x. Using this terminology, the
next result shows that it is always possible to approximate from a dense subspace
while simultaneously preserving energy.
294 INTERPOLATION AND APPROXIMATION

11.11 Simultaneous Approximation and Norm-preservation (SAN).


Let Y be a dense subspace of X. Then, for each x E X and E > 0, there exists
y E Y such that

(11.11.1) Ilx-yll < E

and

(11.11.2) Ilyll = Ilxll·

Proof. Ifx = 0, takey = o. Ifx I 0, choose z E Y\{O} such that Ilx-zll < E/2.
Then y = Ilxll(llzll-1}z E Y, lIyll = Ilxll, and

Ilx - yll = Ilx -lIxll(llzll-1}zll ~ Ilx - zll + liz -llxll(lIzll-1)zll


~ Ilx - zll + Illzll-llxlll ~ 211x - zll < E. •

Simultaneous Approximation, Interpolation, and Norm-preservation


In contrast to Theorems 11.9 and 11.11, it is not always possible to simultane-
ously interpolate and preserve norm (SIN), and a fortiori, it is not always possible
to simultaneously approximate, interpolate, and preserve norm (SAIN). These two
latter properties turn out to be equivalent and can be characterized. First we state
an essential lemma.
11.12 Lemma. Let Y be a dense subspace of X, {Xi,X2, ... ,x;'} a subset of
X*, and
n
Z = n{x E X I xi(x) = OJ.
1

Then Y n Z is dense in Z.
Proof. By Theorem 11.9, for each z E Z and E > 0, there exists y E Y such
that liz - yll < E and xi(y) = xi(z) = 0 for i = 1,2, ... , n. Thus y E Y n Z and
IIz-YIl < E. •

11.13 Theorem. Let Y be a dense subspace of the Hilbert space X and let
{xi, x2, ... , x;'} be a subset of X*. Then the following four statements are equiva-
lent:
(I) (SAIN) For each x E X and E > 0, there exists y E Y such that

(11.13.1) Ilx-yll < E,

(11.13.2) (i = 1,2, ... , n),


and

(11.13.3) lIyll = Ilxll;


(2) (SIN) For each x E X, there exists y E Y such that (11.13.2) and (11.13.3)
hold;
EXERCISES 295

(3) Each xi attains its norm at a point in Y;


(4) Each xi has its representer in Y.
Proof. First observe that since X is a Hilbert space, Theorem 6.10 implies that
each xi has a representer Xi EX. Further, it is no loss of generality to assume that
{Xl, X2,· .. , xn} is linearly independent.
The implication (1) =?- (2) is trivial.
(2) =?- (3). Assume (2) holds, fix any i E {I, 2, ... , n}, and let x = xilllxiII-
Choose Y E Y such that xi(Y) = xi (x) and Ilyll = Ilxll = 1. Then

so xi attains its norm at y.


(3) ~ (4). This is just Theorem 6.12.
(4) =?- (1). If (4) holds, then the representers Xi are all in Y. We must show that
for each x E X and 10 > 0, there is ayE Y such that Ilx - YII < 10, (y, Xi) = (x, Xi)
(i = 1,2, ... , n), and IIYII = Ilxll. Let {Yl, Y2,···, Yn} be an orthonormal basis for
M := span{ Xl, X2, ... , x n }. Now,
n
Ml. = n{y E X I (Y,Yi) = O}
1

and X = M EB Ml. by Theorem 5.9. Thus we can write

for some z E Ml. and Qi


= (X,Yi). Since YnMl. is dense in Ml. by Lemma 11.12,
we can use Theorem 11.11 to obtain w E Y n Ml. such that

Ilw - zll < 10 and Ilwll = IlzII.


Setting Y = L~ QiYi + w, we see that Y E Y, Ilx - YII = liz - wll < 10,

(i=I,2, ... ,n),

and

IIyI1 2 = IItQiYil12 + IIwl1 2 = IltQiYir + IIzl12 = Ilx11 2.


Thus (1) holds. •

Exercises

1. (a) Let X be n-dimensional, {Xl, X2, ... ,xn} be in X, and {XiX2, ... , x~}
be in X*. Show that any two of the following conditions implies the third:
(i) {Xl, X2, . .. ,xn } is linearly independent;
(ii) {xi, x2, ... ,x~} is linearly independent;
(iii) det[xi(xj)) =I O.
(b) More generally than (a), let {Xl, X2, ... , xn } be in X, let M denote
296 INTERPOLATION AND APPROXIMATION

span{XI,X2, ... ,Xn }, and let {Xi,X2, ... ,x;'} be in X*. Show that any two
of the following conditions implies the third:
(i) {Xl, X2, . .. , xn} is linearly independent;
(ii) {xi 1M, x21M, ... , X;. 1M } is linearly independent;
(iii) det[xi(xj)J =I 0 .
2. Exercise 1 can be used to construct a basis for X* if a basis for the n-
dimensional space X is given; it suffices to choose a set of linear functionals
{xi,X2, ... ,x;'} on X such that det[xi(xj)J =I o. As an application, let X =
span{xb X2, ... , X6}, where the functions Xi are defined on [-1, 1J x [-1, 1J
by XI(S,t) = 1, X2(S,t) = s, X3(S,t) = t, X4(S,t) = s2, X5(S,t) = st, and
X6(S, t) = t 2. Find a basis for X*.
3. Let X be the 6-dimensional space of functions defined as in Exercise 2.
Show that it is not always possible to find an x E X that assumes arbitrarily
prescribed values at six distinct points Zi = (Si, ti) in [-1, 1J x [-l,lJ.
4. True or False? For each of the statements below, determine whether it is
possible to construct a polynomial p E P2 satisfying the stated conditions.
Determine also whether the solution is unique when it exists. Justify your
answers.
(a) p(O) = 1, p'(l) = 2, p"(2) = -l.
(b) p(O) = 0, J: p(t)dt = 1,J: t p(t)dt = 2.
(c) p(O) = 1, p"(O) = 0, plll(O) = -l.
(d) p(O) = 1, pll(O) = 0, plll(O) = o.
5. Show that iftl and t2 are distinct real numbers, then the functions XI(t) :=
1, X2(t) := t and the linear functionals

are biorthogonal.
6. (Hermite or osculatory interpolation). Let X = P 2n - 1 (regarded as a
subspace of O2 [a, bJ) and a ::; tl ::; t2 ::; ... ::; tn ::; b.
(a) Show that for any set of 2n real numbers {el, ... , en, d l , ... , dn }, there
exists a unique p E X such that

(b) Verify that the sets offunctions {Xl. X2, ... , X2n} and functionals
{xi,X2, ... ,X2n} are biorthogonal, where
n
w(t) = II(t - ti),
I

(c) Construct the polynomial p E P2n-1 that solves part (a).


[Hint: Part (b).J
EXERCISES 297

7. Let [a, b] be any interval with b - a < 21f. The set of trigonometric
polynomials of degree n is the subspace Tn of C2 [a, b] defined by

Tn = span{l, cos t, cos 2t, . .. ,cos nt, sin t, sin 2t, ... ,sin nt}.

(a) Show that if a ~ tl < t2 < ... < t2n+l ~ b, the functions
{Xl, X2,.··, X2n+l} and the functionals {xi, x 2,··· ,x2n +d are biorthogonal,
where

and xj(x):= x(tj ).

Of course, it must also be verified that Xj E Tn. [Hint: Induction on n.]


(b) Verify that dim Tn = 2n + 1.
(c) Show that for any set of 2n + 1 real scalars Ci, there exists a unique
X E Tn such that

(i=1,2, ... ,2n+1).

Construct X explicitly.
8. Given any n real numbers ti, define the n-th order Vandermonde deter-
minant by

1 1 1
tl t2 tn
Vn(t l , t2,···, tn) :=
t n1 - l t n2 - l n- l
tn

(a) Prove that

Vn(tl, t2,···, tn) = II


l:Sj<i:Sn
(ti-tj).

[Hint: Induct on n using the fact that Vn+l (tl' t2, ... ,tn, t) is a polynomial
of degree n in t that vanishes for t = ti (i = 1,2, ... , n).]
(b) Give another proof of Theorem 11.4 based on Theorem 11.2 and part
(a).
9. (a) Prove the following "converse" to Theorem 11.2. Let {xi,x2, ... ,x~}
be n linear functionals on X. Then the following statements are equivalent.
(i) {xi, X2' ... ,x~} is linearly independent;
(ii) There exist n vectors Xl,X2, ... ,X n in X such that xj(Xj) = Dij
(i,j = 1,2, ... ,n);
(iii) For each set of n real scalars {Cl' C2, ... , Cn }, there exists y E X
such that xi(y) = Ci (i = 1,2, ... ,n);
(iv) There exists an n-dimensional subspace M of X such that
{xi 1M, x 21M, ... , x~ 1M} is linearly independent.
(b) When is the element y of (iii) in part (a) unique?
[Hint: When does nJ'{x E x. I xi(x) = O} consist of only the zero element?]
298 INTERPOLATION AND APPROXIMATION

10. If {Xl,X2, ... ,Xn} is linearly independent in X, then there exists


{Xi,X2, ... ,x~} in X* such that xi(xj) = Oij (i,j = 1,2, ... ,n).
[Hint: Regard the Xi as linear functionals Xi on X* via the definition
Xi(X*) := X*(Xi) for every x* E X*. Apply Theorem 11.6.]
11. Let {x*, xi, ... , x~} be in X*. Show that x* E con{xi, x2, ... , x~} if and
only if x* (x) :::: 0 whenever xi (x) :::: 0 for all i. (Recall that con (A) denotes
the conical hull of A: the set of all nonnegative linear combinations of
elements of A.)
12. Show that Theorem 11.13 is also valid in a complex Hilbert space.
13. There are several results in this chapter that are valid in any (real or com-
plex) normed linear space. Verify that each of the following is of this type.
(a) Theorem 11.2, where the scalars in (2) and (3) are in C.
(b) Corollary 11.3, where the scalars in (1) and (2) are in C.
(c) Theorem 11.6, where the scalars c; are in C.
(d) Theorem 11.9.
(e) Theorem 11.11.
(f) Lemma 11.12.
(g) Exercise 11.14.1.
(h) Exercise 11.14.9.
(i) Exercise 11.14.10.

Historical Notes
The classical results concerning polynomial interpolation, Taylor's theorem, di-
vided differences, etc., can be found in the books of Davis (1963), Walsh (1956),
and Steffensen (1950). For a more modern approach to polynomial interpolation of
functions, with the emphasis on what happens when the number of interpolation
points becomes infinitely large, see Szabados and Vertesi (1990). The equivalence
of (1), (5), and (7) of Corollary 11.3 can be found in Davis (1963; p. 26). The
first example of a "SAl" (simultaneous approximation and interpolation) theorem
seems to be that of Walsh in 1928 (see Walsh (1956; p. 310)). He proved that a
continuous function f on a compact subset of the complex plane can be uniformly
approximated by polynomials that also interpolate f on a prescribed finite subset
of the domain. The abstract SAl Theorem 11.9, under the more general normed
linear space setting with Y dense and convex (and not necessarily a subspace), was
established by Yamabe (1950). Singer (1959) further generalized Yamabe's theorem
to a linear topological space setting. Unaware of Singer's result, Deutsch (1966)
generalized Theorem 11.9 to the linear topological space setting.
Theorem 11.11 is a special case of a result of Deutsch and Morris (1969; Theorem
3.1). The first "SAIN" (simultaneous approximation and interpolation with norm-
preservation) theorem was given by Wolibner (1951), who proved the following. Let
a :s: tl < t2 < ... < tn :s: b and let f be a real continuous function on [a, b]. Then for
each E > 0, there exists a polynomial p such that Ilf - plloo < E, p(ti) = f(ti) (i =
1,2, ... ,n), and Ilplloo = Ilflloo. (Here Ilglloo := max{lg(t) I I t E [a, b]}.) Paszkowski
(1957) proved the first "SIN" (simultaneous interpolation and norm-preservation)
theorem when he proved the following. If f is a continuous function on an interval
[a, b] and a :s: tl < t2 < ... < tn :s: b, then there exists a polynomial p such that
p(ti) = f(ti) for i = 1,2, ... , nand Ilplloo = Ilflloo. (In fact, he even showed that
the degree of the polynomial that works is independent of the function, but depends
HISTORlCAL NOTES 299

only on the points til) The equivalence of (1), (3), and (4) of Theorem 11.13 was
proved by Deutsch and Morris (1969). The equivalence of (1) and (2) in Theorem
11.13 was established by Deutsch (1985). Results of SAIN (respectively SIN) type,
in a more general normed linear space setting, can be found in Deutsch and Morris
(1969) (respectively Deutsch (1985)).
CHAPTER 12

CONVEXITY OF CHEBYSHEV SETS

Is Every Chebyshev Set Convex?


In Theorem 3.5 we saw that every closed convex subset of a Hilbert space is a
Chebyshev set. In this chapter we will study the converse problem of whether or
not every Chebyshev subset of a Hilbert space must be convex.
The main result is this: Every boundedly compact Chebyshev set must be con-
vex (Theorem 12.6). In particular, every Chebyshev subset of a finite-dimensional
Hilbert space is convex. Whether this result is also true when the Hilbert space is
infinite-dimensional is perhaps the major unsolved problem in (abstract) approxi-
mation theory today.
It is convenient to introduce notation for line segments or intervals in X. The
line segment or interval joining the two points x and y in X is the set

[x, y] := {Ax + (1 - A)y I 0 :S A :S I}.

That is, [x, y] = co{x, y} is the convex hull of x and y.


Our first observation is that every point on the line segment joining x to its best
approximation PK(x) also has PK(X) as its best approximation (see Figure 12.0.1).

, 0 X

Figure 12.0.1

12.1 Lemma. Let K be a Chebyshev set in X and x E X. Then PK(y) = PK(X)


for every y E [x, PK(x)].
Proof. If the result were false, then there would exist y E [x, PK(X)] and z E K
such that Ily - zll < Ily - PK(x)ll . We can write y = AX + (1- A)PK(x) for some

F. Deutsch, Best Approximation in Inner Product Spaces


© Springer-Verlag New York, Inc. 2001
302 CONVEXITY OF CHEBYSHEV SETS

0< A < 1. Then

Ilx - zll ::; Ilx - yll + Ily - zll < Ilx - yll + Ily - PK(x)11
= (1- A)llx - PK(x)11 + Allx - PK(x)11 = Ilx - PK(x)11 = d(x,K),
which is absurd. •

Chebyshev Suns
A concept that will prove useful in studying Chebyshev sets is that of a "sun."
12.2 Definition. A Chebyshev set K in X is called a sun if for each x E X,

for every A 2: O.

In other words (using Lemma 12.1), K is a sun if and only if each point on
the ray from PK(x) through x also has PK(X) as its best approximation in K (see
Figure 12.2.1).

Figure 12.2.1

The next result shows that not only is a sun equivalent to a convex set, but
that the contractive property of the metric projection onto a convex Chebyshev set
actually characterizes those Chebyshev sets that are convex.
12.3 Suns and Convex Sets are Equivalent. Let K be a Chebyshev set in
an inner product space X. Then the following statements are equivalent:
(1) K is convex;
(2) K is a sun;
(3) For every x E X, PK(X) = P[Y,h(x)] (x) for every y E K;
(4) PK is nonexpansive:
(12.3.1)

Proof. (1) =? (2). If K is not a sun, then there exist x E X and A > 0 such
that z:= x + A[x - PK (x) 1satisfies PK (z) =1= PK (x). Thus there exists y E K such
that liz - yll < liz - PK(x)ll· Substituting for z and dividing both sides of this
inequality by 1 + A, we obtain
CHEBYSHEV SUNS 303

This proves that K cannot be convex, since otherwise >"(1+>..)-1 PK(x)+ (1 +>..)-l y
would be an element of K closer to x than PK(X).
(2) =} (3). If (3) fails, there exist x E X and y E K such that PK(X) f
P[Y,h(x)] (x). Thus there exists 0 < >.. ~ 1 such that z := >..y + (1 - >")PK(X) =
P[Y,h(x)](X) and
Ilx - zll < Ilx - PK(x)lI·
Dividing by>.., we get

Ilx + 1 ~ >"[x - PK(x)]- yll < II~x - ~PK(x)1I


= IIx+ 1 ~ >"[x- PK(x)] - PK (X)II.
This proves that y is a better approximation to x + (1 - >..).-1 [x - PK(X)] than
PK(X). Hence K is not a sun.
(3) =} (4). Assume (3) holds and let x and Y be in X. Since both PK(x) =
P[h(x),h(Y)](X) and PK(Y) = P[pK(x),h(Y)] (y), then by the characterization the-
orem (Theorem 4.1) (applied to the convex set [PK(x), PK (y)]) , we deduce (x -
PK(X), PK(y) - PK(x)) ~ 0 and (y - PK(y), PK(x) - PK(y)) ~ O. Equivalently,
(x - PK(X), Pdx) - PK(y)) ~ 0 and (PK(Y) - y, PK(X) - PK(y)) ~ O.
Adding these, we deduce that
o~ (x - Y + PK(Y) - PK(X) , PK(x) - PK(y))
= (x - y,PK(x) - PK(y)) -IIPK(X) - PK (y)11 2
~ Ilx - yIIIIPK(x) - PK(Y)II-IIPK(X) - PK(y)11 2 ,
which proves (4).
(4) =} (1). If K is not convex, then there exist points x and y in K such that
z := ~(x + y) rt K (see Exercise 12.8.1). Next we observe that
(12.3.2) max{llx - PK(z)ll, Ily - PK(z)ll} > Ilx - yll/2.
For otherwise, Ilx - PK(z)11 ~ Ilx - YII/2 and Ily - PK(z)11 ~ IIx - YII/2. Then
Ilx - yll ~ Ilx - PK(z) II + IIPK(z) - yll ~ Ilx - yll
implies that equality must hold for all these inequalities. In particular,
Il[x - PK(Z)] + [PK(Z) - Y]II = Ilx - PK(z)11 + IIPK(Z) - yll·
By the condition of equality in the triangle inequality, we must have x - PK(z)=
p[PK(Z) - y] for some p ~ O. Since Ilx - PK(Z) II = Ilx - YII/2 = Ily - PK(z)lI, it
follows that p = 1 and

which is absurd. Thus (12.3.2) holds.


Without loss of generality, we may assume that Ilx - PK(z)11 > Ilx - yll/2. But
since x E K, we have that PK(x) = x and Ilx - YII/2 = IIx - zll implies that
IIPK(x) - PK(z)11 > Ilx - zII.
That is, P K is not nonexpansive. •
The metric projection onto an approximately compact Chebyshev set is always
continuous.
304 CONVEXITY OF CHEBYSHEV SETS

12.4 Lemma. If K is an approximatively compact Chebyshev set, then PK is


continuous.
Proof. If not, there is a point x E X and a sequence {xn} converging to x such
that

(12.4.1)

Now by Theorem 5.3,

d(x, K) :S Ilx - PK(xn) I :S Ilx - Xn II + Ilxn - PK(xn) II


= Ilx - xnll + d(xn, K) -t d(x, K).
Thus {PK(x n)} is a minimizing sequence. By approximative compactness, there is
a subsequence {PK(X ni )} that converges to a point y E K. Then

Ilx - yll = lim Ilx - PK(xnJII = d(x, K)

implies that y = PK(x) and PK(xnJ -t PK(x). But this contradicts (12.4.1). •
For proving the main result of this chapter we shall need the following well-known
fixed-point theorem, which we state without proof.
12.5 Schauder Fixed-Point Theorem. Let B be a closed convex subset of
the inner product space X and let F be a continuous mapping from B into a
compact subset of B. Then F has a "fixed-point"; that is, there is Xo E B such
that F(xo) = Xo.

Convexity of Boundedly Compact Chebyshev Sets


Recall that a subset K of X is boundedly compact if each bounded sequence in
K has a subsequence converging to a point in K. The main result of this chapter
can be stated as follows.
12.6 Boundedly Compact Chebyshev Sets are Convex. If K is a bound-
edly compact Chebyshev set in an inner product space, then K is convex.
Proof. By Theorem 12.3, it suffices to verify that K is a sun. We proceed to
prove this by contradiction.
Suppose K were not a sun. Then there would exist x E X\K such that not
every element of the form X(A) := x + A(X - PK(X)), A =2: 0, has PK(x) as its best
approximation in K. Let

Using Lemma 12.1 we see that PK(X(A)) = PK(x) for every :S A < A1 and
PK(X(A)) f PK(X) for every A > A1. If A1 > 0, the continuity of the distance
°
function (Theorem 5.3) implies that

d(X(Al), K) = lim d(x(A), K) = lim Ilx(A) - PK(x)11


.:I.-t.:l.;- .:I.-t.:l.;-

= lim (1 + A)llx - PK(x)11 = (1 + A1)llx - PK(x)11


.:I.-t.:l.;-
CONVEXITY OF BOUNDEDLY COMPACT CHEBYSHEV SETS 305

If Al = 0, then X(Al) = x and the above equality holds trivially. Thus PK(X(Al)) =
PK(X) and X(Al) is the farthest point on the ray from PK(X) through x that has
PK(x) as its best approximation. For simplicity, set Xl = X(Al)'
Now choose any r > 0 such that the closed ball B = {y E X I Ilxl - YII ::; r} is
disjoint from K. Define a mapping F on B by

(12.6.1)

Figure 12.6.3

Clearly, IIF(y) - xIII = r, so F maps B into itself (see Figure 12.6.3).


Since K is boundedly compact, it is approximatively compact (by Lemma 3.11),
so P K is continuous by Lemma 12.4. Since the norm is continuous and II Xl -
PK(y)11 > r for all y E B, it follows that F is continuous.
Claim: The range of F is contained in a compact subset of B.
It suffices to show that F(B), the closure of the range of F, is compact. Let
{zn} be any sequence in F(B) (n :::: 2). For each n :::: 2, choose Yn E F(B) such
that Ilzn - Ynll < lin. Then choose Xn E B such that Yn = F(xn); that is,

By the nonexpansiveness of the distance function d(x, K) (Theorem 5.3), it follows


that for every ko E K,

and so {PK(X n )} is a bounded sequence in K. Hence there exists a subsequence


{PK(XnJ} converging to a point Zo E K. Thus
r
Yni -+ Xl + II Xl - Zo
II [Xl - zo] =: Yo·

Since Yni E F(B) for all i, Yo E F(B). Also,

Ilzni - Yoll ::; Ilzni - YnJ + IIYni - Yoll-+ O.


306 CONVEXITY OF CHEBYSHEV SETS

Hence {zn} has a subsequence converging to a point in F(B). Thus F(B) is com-
pact, and this proves the claim.
By the Schauder fixed-point theorem (Theorem 12.5), F has a fixed point Xo E B:

(12.6.2)

Thus
II Xl - PK(xo)11 + r P ( )
Xl = r + Ilxl _ PK(Xo) 1(0 r + II Xl _ PK(xo)11 K Xo
is a proper convex combination of Xo and PK(xo). By Lemma 12.1, PK(XI)
PK(xo). Substituting this into (12.6.2), we obtain

But this contradicts the assumption that Xl was the farthest point on the ray from
PK(XI) through Xl that has PK(XI) as its best approximation. This completes the
proof. •
12.7 Convexity of Chebyshev Sets in Finite-Dilllensional Spaces. Let
X be a finite-dimensional inner product space and K be a nonempty subset of X.
Then K is a Chebyshev set if and only if K is closed and convex.
Proof. The "if' part follows from Theorem 3.5 using the fact that finite-
dimensional spaces are complete (Theorem 3.7(3)). Conversely, if K is a Chebyshev
set, then K is closed by Theorem 3.1, and K is boundedly compact by Lemma
3.7(1). The result now follows by Theorem 12.6. •
The main unanswered question concerning these results is the following:
Question. Must every Chebyshev set in (an infinite-dimensional) Hilbert space
be convex?
We believe that the answer is no.

Exercises

1. A set K in X is called midpoint convex if ~(x + y) E K whenever X, y E


K.
(a) Show that every convex set is midpoint convex; and every midpoint
convex set that is closed must be convex.
(b) Give an example of a set in lR that is midpoint convex, but not convex.
(c) Show that if K is midpoint convex, then K is dense in its closed convex
hull, co(K).
2. Let K be a convex Chebyshev set in the inner product space X. Show that
for each pair X, y E X, either IIPK(x)-PK(y)11 < Ilx-yll or PK(X)-PK(Y) =
x-Yo
3. Prove that for any pair of vectors X, Z in X,

(1) {y E X I (y - Z,X - z) > O} = U.>.>oB(z + A(X - z), Allx - zll).


In words, the open half-space containing X that is determined by the hy-
perplane through z and orthogonal to X - z is the union of all open balls
HISTORICAL NOTES 307

with centers on the ray from z through x whose radius is the distance from
the center to z. [Hint: Let Land R denote the sets on the left and right
sides of (1). Note that y E R if and only if liz + A(X - z) - yll < Allx - zll
for some A > O. Square both sides of this inequality and simplify to deduce
that y E L.]
4. Show that a Chebyshev set K is a sun if and only if for every x E X \ K,

Kn {y E X I (y- PK(X),X- PK(X)) > O} = 0.


[Hint: Exercise 3 with z = PK(X),]
5. Give a proof that a Chebyshev set in 12 (2) is convex without appealing to
any fixed-point theorem.
6. Let K be a Chebyshev set in any normed linear space X. For the statements
below, prove that (a) =? (b) =? (c).
(a) K is convex;
(b) K is a sun;
(c) For every x EX, PK(X) = P[Y,h(x)] (x) for every y E K.
7. Prove the following generalization of Theorem 12.6 to any normed linear
space. Every boundedly compact Chebyshev set in a normed linear space
X is a sun. [Hint: Schauder's fixed-point theorem is valid in any normed
linear space.]
8. Let X denote the 2-dimensional space ]R2 with the norm Ilxll
maxl<i<2{lx(i)I}. Show that the set K = {x E X I x(2) 2: -x(1)/3 or x(2)
2: -3;;(1)} is a nonconvex Chebyshev set.
Historical Notes
The idea of a sun was first developed and used by Klee (1953), but the termi-
nology "sun" was first proposed a little later by Efimov and Stechkin (1958).
Phelps (1957) showed that (4) implies (1) in Theorem 12.3. That is, he showed
that if PK is nonexpansive, then K must be convex. The proof that a Chebyshev
set in a finite-dimensional Hilbert space must be convex was first given in the Dutch
doctoral thesis of Bunt (1934). Motzkin (1935) gave one proof of the convexity of
Chebyshev sets in 12 (2), and another one (1935a) for bounded Chebyshev sets in
12(2). Kritikos (1938), apparently unaware of the earlier proofs, gave another proof
of the convexity of Chebyshev sets in 12(n). Jessen (1940), aware of Kritikos's proof,
gave yet another one. Busemann (1947) and Valentine (1964) generalized these
results by showing the convexity of Chebyshev sets in any finite-dimensional smooth
and strictly convex normed linear space. Klee (1953) showed that strict convexity
was not essential for the Busemann-Valentine result. Klee (1961) generalized this
further by showing the convexity of a Chebyshev set in a smooth reflexive Banach
space if its metric projection satisfies certain continuity criteria. Vlasov (1961)
showed that in a smooth Banach space, every boundedly compact Chebyshev set is
convex. In particular, this yields Theorem 12.6 as a corollary. Theorem 12.6 seems
to be the strongest (positive) result known about the convexity of Chebyshev sets
in a Hilbert space. However, every item of the following theorem is a stronger result
in the sense that it actually characterizes those Chebyshev sets that are convex.
12.8 TheoreIn. Let C be a Chebyshev set in the Hilbert space X. Each of the
following statements is equivalent to C being convex.
(1) C is weakly closed (Klee (1961));
308 CONVEXITY OF CHEBYSHEV SETS

(2) C is approximatively compact (Efimov-Stechkin (1961));


(3) Pc is continuous (Vlasov (1967) and Asplund (1969));
(4) Pc is "radially" continuous (Vlasov (1967a)), i.e., for any x E X, the re-
striction of Pc to the ray {x + >..(x - Pc(x)] I >.. :::0: O} is continuous at
x·,
(5)
lim d(x" C) - d(x, C) = 1
<->0+ Ilx, - xii
for every x E X \ C, where x, = x + E(X - Pc(x)] (Vlasov (1967a);
(6) Pc is continuous except possibly at a countable number of points
(Balaganskii (1982));
(7) Pc is continuous except possibly on a subset of a hyperplane (Vlasov and
Balaganskii (1996; Corollary 2.8));
(8) C is a sun (Efimov-Stechkin (1961));
(9) C intersects each closed half-space in a proximinal set (Asplund (1969)).

Asplund (1969) showed that if a nonconvex Chebyshev set exists in a Hilbert


space X, then one also exists that has a bounded convex complement. Johnson
(1987) gave an example of a nonconvex Chebyshev set in an incomplete inner prod-
uct space. Jiang (1990) found some errors in Johnson's example, but showed that
all were correctable. Johnson's example does not answer the question, since the
space was not complete, but it tends to support the conjecture of Klee (1966) that
nonconvex Chebyshev sets exist in (some) infinite-dimensional Hilbert space. By
modifying his construction in (Johnson, 1987), Johnson (1998) showed that the non-
convex Chebyshev set could be constructed so as also to be bounded. Berens and
Westphal (1978) have shown that the "maximal monotonicity" of Pc is equivalent
to the convexity of C. See also Berens (1980) for an exposition of the monotonicity
aspect of the metric projection in Hilbert space. An excellent (unpublished) expo-
sition of the convexity of Chebyshev sets in finite-dimensional Hilbert spaces, along
with proofs, was given by Kelly (1978). Surveys on the convexity of Chebyshev
sets problem were given by Vlasov (1973), Narang (1977), Singer (1977), Deutsch
(1993), and Balaganskii and Vlasov (1996). The latter survey also contains a geo-
metric proof of an example of Vlasov, similar to Johnson's example. In addition,
the Balaganskii and Vlasov survey (1996) contains a large amount of material re-
lating the geometry of Banach spaces with the approximative properties of certain
subsets.
Related to this circle of ideas is the problem of characterizing geometrically those
finite-dimensional normed linear spaces X in which each Chebyshev set is convex.
When n := dimX S 4, it is known that each Chebyshev set in X is convex if and
only if each "exposed point" of the unit sphere S(X) is a "smooth point." (The
case n = 1 is trivial; the case n = 2 follows because if each exposed point of S(X)
is a smooth point, then X is smooth; the proof when n = 3 is due to Brondsted
(1966); the proof when n = 4 was given by Brown (1980), who developed quite an
elaborate topological machinery for this purpose.) Whether the same result holds
true when n > 4 is still open, but I suspect it is. In particular, there are finite-
dimensional normed linear spaces (e.g., the "smooth spaces") that are not Hilbert
spaces, but all of whose Chebyshev subsets are convex (Klee (1961)).
However, by restricting the Chebyshev sets being considered, a characterization
HISTORICAL NOTES 309

can be given. Tsar'kov (1984) showed that every bounded Chebyshev set in a finite-
dimensional normed linear space X is convex if and only if the extreme points of
S(X*) are dense in S(X*). See also Brown (1986) for related material.
Finally, we note that the question of the convexity of Chebyshev sets in an
infinite-dimensional Hilbert space was first posed explicitly by Klee (1961) and im-
plicitly by Efimov and Stechkin (1961), since they asked the question (which is
equivalent by Theorem 12.8(2)) whether each Chebyshev set must be approxima-
tively compact.
ZORN'S LEMMA 311

APPENDIX 1

ZORN'S LEMMA

Definition. A partially ordered set is a pair (S, -<), where S is a nonempty


set and -< is a relation on S that satisfies the following three properties:
(1) x -< x for all XES.
(2) If x -< y and y -< z, then x -< z.
(3) Ifx -< y and y -< x, then x = y.
The relation -< is called a partial order on S. In other words, a partial order
is (1) "reflexive," (2) "transitive," and (3) "anti-symmetric." For example, the
inclusion relation C is a partial order on the collection S = 2x of all subsets of a
:s
set X, and the relation is a partial order on the set of real numbers R.
We use terminology suggested by these two examples by saying that x is greater
than y when we write y -< x. A totally ordered (or linearly ordered) subset of the
partially ordered set (S, -<) is a subset T of S such that for each pair of elements
x and y in T, either x -< y or y -< x. For example, any subset T of JR is totally
ordered in (JR, :S), but 2x is not totally ordered in (2X, C).
Let (S, -<) be a partially ordered set and let T C S. An element b E S is called
an upper bound for T if t -< b for every t E T. An element m E S is called
maximal if m -< x implies x -< m. That is, the only element of S that is greater
than m is m itself.
Zorn's Lemma. A partially ordered set has a maximal element if each totally
ordered subset has an upper bound.
It is known that Zorn's lemma is logically equivalent to the axiom of choice,
which asserts the possibility of forming a new set by choosing exactly one element
from each set in an arbitrary collection of nonempty sets.
312 APPENDICES

APPENDIX 2

EVERY HILBERT SPACE IS £2(1)

In this appendix we prove that every Hilbert space is equivalent to the Hilbert
space £2(I) for some index set I.
We recall (see Exercise 4 of Chapter 6) that two Hilbert spaces X and Yare
called equivalent if there is a bijective mapping U from X onto Y that preserves
linearity,

U(o:x + (3y) = 0: U(x) + (3 U(y) for all x, y EX, 0:, (3 E JR.,

preserves the inner product,

(U(x), U(y)) = (x,y) for all x, y EX,

and preserves the norm,

IIU(x)11 = IIxll for all x EX.

That is, two equivalent Hilbert spaces are essentially the same. The mapping U
amounts to nothing more than a relabeling of the elements.
In Exercise 4 of Chapter 6 it was outlined that X and Yare equivalent if and
only if there exists a linear mapping U from X onto Y that is norm-preserving:
lIU(x) II = Ilxll for all x E X.
Theorem. Every Hilbert space X i= {O} is equivalent to the Hilbert space £2 (I)
for some index set I.
Proof. By Corollary 4.20 and Theorem 4.21, X has an orthonormal basis
{ei 1 i E I}. Thus each x E X has the representation

1/2

x= L (x, ei)ei and IIxll = ~


[
1 (x, ei) 12
]

iEI

Define the map U on X into the space of functions on I by

(Ux)(i):= (x,ei), i E I.

Note that since

iEI iEI

by Corollary 4.20, Ux E £2(I). Thus U : X -> £2(I). It is easy to see that U is


linear. Moreover, the last equation above shows that U is norm-preserving. To
complete the proof, it suffices to show that U is onto.
EVERY HILBERT SPACE IS £2(1) 313

First note that functions r5i : 1 -> lR defined for each i E 1 by

(Kronecker's delta),

form an orthonormal basis for £2(1). In fact, if y E £2(1), then

y = L>(i)r5i , L ly(i)12 = IIyl12 .


iEI iEI

Given any y E £2(1), define x by

x = Ly(i)ei.
iEI

Then x E X and (Ux)(i) = y(i) for each i E 1. That is, Ux = y, so U is onto. •


Implicit in the proof was the fact that if X is equivalent to £2(1), then the
cardinality of 1 is equal to the cardinality of an orthonormal basis for X. In other
words, the dimensions of X and £2(1) are the same. (The dimension of a Hilbert
space is the cardinality of an arbitrary orthonormal basis.)
REFERENCES
N. I. Achieser
[lJ Theory of Approximation, Ungar, New York, 1956.

N. I. Achieser and M. G. Krein


[lJ Some Questions in the Theory of Moments, Kharkov, 1938 (Russian). [Translated
in: Translations of Mathematical Monographs, Vol. 2, Amer. Math. Soc., Providence,
RI, 1962.J

P. Alexandroff and H. Hopf


[lJ Topologie, Springer, Berlin, 1935. (Reprinted by Chelsea Publ. Co., New York, 1965.)

D. Amir
[lJ Characterizations of Inner Product Spaces, Birkhiiuser Verlag, Basel, 1986.

D. Amir, F. Deutsch, and W. Li


[lJ Characterizing the Chebyshev polyhedrons in an inner product space, preprint, 2000.

V. E. Anderson and P. A. Ivert


[lJ Constrained interpolants with minimal Wk'P-norm, J. Approx. Theory, 49(1987),
283-288.

W. N. Anderson, Jr. and R. J. Duffin


[lJ Series and parallel addition of matrices, J. Math. Anal. Appl., 26(1969), 576-594.

N. Aronszajn
[lJ Theory of reproducing kernels, Trans. Amer. Math. Soc., 68(1950), 337-403.
[2J Introduction to the Theory of Hilbert Spaces, Vol. 1, Research Foundation of
Oklahoma A & M College, Stillwater, OK, 1950a.

N. Aronszajn and K. T. Smith


[lJ Invariant subspaces of completely continuous operators, Annals Math., 60(1954),
345-350.

G. Ascoli
[lJ Sugli spazi lineari metrici e Ie loro varieta lineari, Ann. Mat. Pura Appl., (4) 10(1932),
33-81, 203-232.

E. Asplund
[lJ Cebysev sets in Hilbert space, Trans. Amer. Math. Soc., 144(1969),235-240.

F. V. Atkinson
[lJ The normal solubility of linear equations in normed spaces (Russian), Mat. Sbornik
N.S., 28 (70) (1951),3-14.
[2J On relatively regular operators, Acta Sci. Math. Szeged, 15(1953), 38-56.

B. Atlestam and F. Sullivan


[lJ Iteration with best approximation operators, Rev. Roum. Math. Pures et Appl., 21(1976),
125-131.

J.-P. Aubin
[lJ Applied Functional Analysis, Wiley-Interscience, New York, 1979.

R. Baire
[lJ Sur les f~nctions de variables n\elles, Ann. Mat. Pura Appl., 3(1899), 1-123.

V. S. Balaganskii
[lJ Approximative properties of sets in Hilbert space, Math. Notes, 31(1982), 397-404.

V. S. Balaganskii and L. P. Vlasov


[lJ The problem of convexity of Chebyshev sets, Russian Math. Surveys, 51(1996),
1127-1190. [English translation of: Uspekhi Mat. Nauk, 51(1996), 125-188.J
316 REFERENCES

S. Banach
[1] Sur les fonctionnelles Iineaires I, Studia Math., 1(1929a), 211-216.
[2] Sur les fonctionnelles Iineaires II, Studia Math., 1(1929b), 223-239.
[3] Theoremes sur les ensembles de premieres categorie, Fund. Math., 16(1930), 395-398.
[4] Theorie des operations Iineaires, Monografje Matematyczne, Warsaw, 1932. (Reprinted
by Chelsea Pub!. Co., New York, 1955.)

S. Banach and H. Steinhaus


[1] Sur Ie principe de la condensation de singularites, Fund. Math., 9(1927), 50-61.

V. Barbu and Th. Precupanu


[1] Convexity and Optimization in Banach Spaces, Sijthoff & Noordhoff, the
Netherlands, 1978.

H. H. Bauschke
[1] Projection Algorithms and Monotone Operators, Ph.D. thesis, Simon Fraser
University, 1996.

H. H. Bauschke and J. M. Borwein


[1] On the convergence of von Neumann's alternating projection algorithm for two sets,
Set-Valued Analysis, 1(1993), 185-212.
[2] On projection algorithms for solving convex feasibility problems, SIAM Review, 38(1996)
367-426.

H. H. Bauschke, J. M. Borwein, and A. S. Lewis


[1] The method of cyclic projections for closed convex sets in Hilbert space, in
Optimization and Nonlinear Analysis (edited by Y. Censor and S. Reich),
Contemporary Mathematics, Amer. Math. Soc., 1996.

H. H. Bauschke, J. M. Borwein, and W. Li


[1] Strong conical hull intersection property, bounded linear regularity, Jameson's property
(G), and error bounds in convex optimization, Math. Program., 86(1999), 135-160.

H. H. Bauschke, F. Deutsch, H. Hundal, and S.-H. Park


[1] Accelerating the convergence of the method of alternating projections, preprint, July
23, 1999.

A. Ben-Israel and T. N. E. Greville


[1] Generalized Inverses: Theory and Applications, John Wiley & Sons, New York,
1974.

H. Berens
[1] Vorlesung liber Nichtlineare Approximation, Universitiit Erlangen-Niirnberg
Lecture Notes, 1977/78.
[2] Best approximation in Hilbert space, in Approximation Theory III (edited by
E. W. Cheney), Academic Press, New York, 1980, 1-20.

H. Berens and U. Westphal


[1] Kodissipative metrische Projektionen in normierten linearen Riiume, in Linear Spaces
and Approximation (edited by P. L. Butzer and B. Sz.-Nagy), ISNM, 40(1978),
120-130.

S. N. Bernstein
[1] Demonstration du theoreme de Weierstrass fardee sur Ie caleul de probabilites,
Communications of the Kharkov Mathematical Society, 13(1912), 1-2.

S. Bernstein and C. de la Vallee Poussin


[1] L'Approximation, Chelsea Pub!. Co., New York, 1970. [This is a reprint in one volume
of two books "Le<;ons sur les Proprietes Extremales et la Meilleure Approximation des
Fonctions Analytiques d'une Variable Reele", by S. Bernstein, Paris, 1926, and "Le<;ons
sur L'Approximation des Fonctions d'une Variable Reelle" by C. de la Vallee Poussin,
Paris, 1919. Text was unaltered except for correction of errata.]
REFERENCES 317

A. Bjerhammar
[1] Rectangular reciprocal matrices, with special reference to geodetic calculations, Bull.
Geodesique, 1951, 188-220.

A. Bjorck
[1] Numerical Methods for Least Squares Problems, SIAM, Philadelphia, 1996.

H. Bohman
[1] On approximation of continuous and analytic functions, Arkiv for Matematik, 2(1952),
43-56.

H. F. Bohnenblust and A. Sobczyk


[1] Extensions of functionals on complex linear spaces, Bull. Amer. Math. Soc., 44(1938),
91-93.

B. P. J. N. Bolzano
[1] Rein analytischer Beweis des Lehrsatzes ... , Prague, 1817; Abh. Gesell. Wiss. Prague,
(3) 5 (1818), 1-60; Ostwald's Klassiker, 153 (edited by P. E. B. Jourdain), Leipzig,
1905,3-43.

C. de Boor
[1] On "best" interpolation, J. Approx. Theory, 16(1976), 28-42.

E. Borel
[1] These, Annales Scientifiques de I'Ecole Normal Superieure, (3), 12(1895),9-55.

J. M. Borwein and A. S. Lewis


[1] Partially finite convex programming, Part I: Quasi relative interior and duality theory,
Math. Programming, 57(1992), 15-48.
[2] Partially finite convex programming, Part II: Explicit lattice models, Math.
Programming, 57(1992a), 49-83.

J. M. Borwein and H. Wolkowicz


[1] Facial reduction for a cone-convex programming problem, J. Austral. Math. Soc., Ser.
A, 30(1981), 369-380.
[2] A simple constraint qualification in infinite-dimensional programming, Math.
Programming, 35(1986), 83-96.

T. L. Boullion and P. L. Odell


[1] Proceedings of the Symposium on Theory and Applications of Generalized
Inverses of Matrices, Texas Tech Press, Lubbock, Texas, 1968.
[2] Generalized Inverse Matrices, Wiley-Interscience, New York, 1971.

J. P. Boyle and R. L. Dykstra


[1] A method for finding projections onto the intersection of convex sets in Hilbert spaces,
in Advances in Order Restricted Statistical Inference, Lecture Notes in
Statistics, Springer-Verlag, 1985, 28-47.

A. P. Boznay
[1] A remark on the alternating algorithm, Periodica Math. Hungarica, 17(1986), 241-244.

D. Braess
[1] Nonlinear Approximation Theory, Springer-Verlag, New York, 1986.

A. Br~mdsted
[1] Convex ~ets and Chebyshev sets II, Math. Scand., 18(1966), 5-15.

B. Brosowski and F. Deutsch


[1] An elementary proof of the Stone-Weierstrass theorem, Proe. Amer. Math. Soc., 81(1981),
89-92.
318 REFERENCES

A. L. Brown
[IJ Abstract Approximation Theory, Seminar in Analysis 4, MATSCIENCE, Madras,
1969/70.
[2J Chebyshev sets and facial systems of convex sets in finite-dimensional spaces, Proc.
London Math. Soc., 41(1980), 297-339.
[3J Chebyshev sets and the shapes of convex bodies, in Methods of Functional Analysis
in Approximation Theory (edited by C. A. Micchelli, D. V. Pai, and B. V. Limaye),
Birkhauser Verlag, Boston, 1986, 97-121.

V. Buniakowsky
[1 J Sur quelques inegalites concernant les integrales ordinaires et les integrales aux differences
finies, Mem. Acad. St. Petersburg (7)1, no. 9(1859).

L. N. H. Bunt
[IJ Bitdrage tot de theorie der konvekse puntverzamelingen, thesis, Univ. of Groningen,
Amsterdam (Dutch), 1934.

R. B. Burckel and S. Saeki


[1J An elementary proof of the Miintz-Sz8sz theorem, Expo. Math., 4(1983), 335-341.

H. Busemann
[1J Note on a theorem on convex sets, Mat. Tidsskrift B, 1947,32-34.

R. Cakon
[1J Alternative proofs of Weierstrass theorem of approximation: an expository paper,
masters paper, The Pennsylvania State University, 1987.

A. L. Cauchy
[1 J Cours d'analyse algebrique, Oeuvres, (2) III, 1821.
[2J Oeuvres, ser. II, t. 3, Gauthier-Villars, Paris, 1900.

E. W. Cheney
[1J Introduction to Approximation Theory, McGraw-Hill, New York, 1966.

W. Cheney and W. Light


[1J A Course in Approximation Theory, Brooks/Cole Pub. Co., Pacific Grove,
CA,2000.

E. W. Cheney and A. A. Goldstein


[IJ Proximity maps for convex sets, Proc. Amer. Math. Soc., 10(1959), 448-450.

C. K. Chui, F. Deutsch, and J. D. Ward


[1J Constrained best approximation in Hilbert space, Constr. Approx., 6(1990), 35-64.
[2J Constrained best approximation in Hilbert space II, J. Approx. Theory, 71(1992),
213-238.

J. B. Conway
[1J A Course in Functional Analysis, Second ed., Springer-Verlag, New York, 1990.

P. Cousin
[1J Sur les fonctions de n variables complexes, Acta Math., 19(1895), 1---{i1.

G. Davis, S. Mallat, and Z. Zhang


[1J Adaptive time-frequency approximations with matching pursuits, in Wavelets:
Theory, Algorithms, and Applications (edited by C. L. Chui, L. Montefusco,
and L. Puccio), Academic Press, New York, 1994, pp. 271-293.

P.J. Davis
[1J Interpolation and Approximation, Blaisdell, New York, 1963. (Reprinted by Dover,
New York, 1975.)
REFERENCES 319

M. M. Day
[IJ Reflexive Banach spaces not isomorphic to uniformly convex spaces, Bull. Amer. Math.
Soc., 47(1941), 313-317.

L. Debnath and P. Mikusinski


[IJ Introduction to Hilbert Spaces with Applications, Academic Press, Inc., Boston,
1990.

C. A. Desoer and B. H. Whalen


[IJ A note on pseudoinverses, J. Soc. Industr. Appl. Math., 11(1963),442-447.

F. Deutsch
[IJ Applications of Functional Analysis to Approximation Theory, Ph.D. thesis, Brown
University, 1966.
[2J Simultaneous interpolation and approximation in topological linear spaces, J. SIAM
Appl. Math., 14(1966), 1180-1190.
[3J The alternating method of von Neumann, in Multivariate Approximation Theory
(edited by W. Schempp and K. Zeller), ISNM 51, Birkhauser Verlag, Basel, 1979,83-96.
[4J Representers of linear functionals, norm-attaining functionals, and best approximation
by cones and linear varieties in inner product spaces, J. Approx. Theory, 36(1982),
226-236.
[5J Von Neumann's alternating method: the rate of convergence, in Approximation
Theory IV, Academic Press, New York, 1983,427-434.
[6J Rate of convergence of the method of alternating projections, ISNM, Vol. 72, Birkhauser
Verlag, Basel, 1984, 96-107.
[7J Simultaneous interpolation and norm-preservation, ISNM, Vol. 74, Birkhauser Verlag,
Basel, 1985, 122-132.
[8J The method of alternating orthogonal projections, in Approximation Theory, Spline
Functions and Applications (edited by S .. P. Singh), Kluwer Academic Publishers, the
Netherlands, 1992, 105-12l.
[9J The convexity of Chebyshev sets in Hilbert space, in Topics in Polynomials of One
and Several Variables and their Applications (edited by Th. M. Rassias,
H. M. Srivastava, and A. Yanushauskas), World Scientific, 1993, 143-150.
[lOJ Dykstra's cyclic projections algorithm: the rate of convergence, in Approximation
Theory, Wavelets and Applications (edited by S. P. Singh), Kluwer Academic
Publishers, the Netherlands, 1995, 87-94.
[l1J The role of the stong conical hull intersection property in convex optimization and
approximation, in Aproximation Theory IX (edited by C. K. Chui and L. L.
Schumaker), Vanderbilt University Press, Nashville, TN, 1998, 105-112.
[12J Accelerating the convergence of the method of alternating projections via a line search:
a brief survey, in the proceedings of the research workshop on parallel algorithms (edited
by D. Butnariu, Y. Censor, and S. Reich), 2001, to appear.

F. Deutsch and H. Hundal


[IJ The rate of convergence of Dykstra's cyclic projections algorithm: the polyhedral case,
Numer. Funct. Anal. and Optimiz., 15 (1994), 537-565.
[2J The rate of convergence for the method of alternating projections II, J. Math. Anal.
Appl., 205(1997), 381-405.

F. Deutsch, W. Li, and J. D. Ward


[I] A dual approach to constrained interpolation from a convex subset of Hilbert space, J.
Approx. Theory, 90(1997), 385-414.
[2] Best approximation from the intersection of a closed convex set and a polyhedron in
Hilbert space, weak Slater conditions, and the strong conical hull intersection property,
SIAM J. Optimiz. 10(2000), 252-268.

F. R. Deutsch and P. H. Maserick


[IJ Applications of the Hahn-Banach theorem in approximation theory, SIAM Review,
9(1967), 516-530.
320 REFERENCES

F. Deutsch, J. H. McCabe, and G. M. Phillips


[1] Some algorithms for computing best approximations from convex cones, SIAM J. Numer.
Anal., 12(1975), 390-403.

F. Deutsch and P. D. Morris


[1] On simultaneous approximation and interpolation which preserves the norm, J. Approx.
Theory, 2(1969), 355-373.

F. Deutsch, V. A. Ubhaya, J. D. Ward, and Y. Xu


[1] Constrained best aproximation in Hilbert space III. Applications to n-convex functions,
Constr. Approx., 12(1996),361-384.

F. Deutsch, V. A. Ubhaya, and Y. Xu


[1] Dual cones, constrained n-convex Lp-approximation, and perfect splines, J. Approx.
Theory, 80(1995), 180--203.

R. A. DeVore and G. G. Lorentz


[1] Constructive Approximation, Springer-Verlag, New York, 1993.

J. Dieudonne
[1] La dualite dans les espaces vectoriels topologiques, Ann. Sci. Ecole Norm. Sup., 59(1942),
107-139.
[2] History of Functional Analysis, North Holland Math. Studies 49, Elsevier, New
York, 1981.

J. Dixmier
[1] Etude sur les varietes et les operaterus de Julia avec quelques application, Bull. Soc.
Math. France, 77(1949), 11-101.

A. L. Dontchev
[1] Duality methods for constrained best interpolation, Math. Balkanica, 1(1987),96--105.

A. L. Dontchev, G. Illiev, and M. M. Konstantinov


[1] Constrained interpolation via optimal control, in Proceedings of the Fourteenth Spring
Conference of the Union of Bulgarian Mathematicians, Sunny Beach, April 6-9, 1985,
385-392.

N. Dunford and J. T. Schwartz


[IJ Linear Operators Part I: General Theory, Interscience Publ., New York, 1958.

R. L. Dykstra
[1J An algorithm for restricted least squares regression, J. Amer. Statist. Assoc., 78(1983),
837-842.

N. V. Efimov and S. B. Stechkin


[lJ Some properties of Chebyshev sets, Dok!. Akad. Nauk SSSR, 118(1958), 17-19.
[2J Approximative compactness and Chebyshev sets, SOy. Math. Dokl., 2(1961), 1226--1228.

M. Eidelheit
[1J Zur Theorie der konvexen Mengen in linearen normierten Riiumen, Studia Math.,
6(1936),104-111.

G. Faber
[1J Uber die interpolatorische Darstellung stetiger Funktionen, Deutsche Mathematiker-
Vereinigung Jahresbericht, 23(1914), 192--210.

J. Favard
[1] Sur l'interpolation, J. Math. Pures Appl., 19(1940), 281-306.

L. Fejer
[1] Sur les fonctions bornees et integrables, Comptes Rendus, 131(1900), 984-987.
REFERENCES 321

[2] Uber Weierstrassche Approximation, besonderes durch Hermitesche Interpolation, Math.


Annal., 102(1930), 707-725.

W. Fenchel
[1] Convex Cones, Sets, and Functions, Lecture notes, Princeton University, 1953.

G. B. Folland
[1] Real Analysis, John Wiley & Sons, New York, 1984.

C. Franchetti
[1] On the alternating approximation method, Sezione Sci., 7(1973), 169-175.

C. Franchetti and W. Light


[1] The alternating algorithm in uniformly convex spaces, J. London Math. Soc., 29(1984),
545-555.

M. Frechet
[1] Sur les operations lineaires I, Trans. Amer. Math. Soc., 5(1904), 493-499.
[2] Sur les operations lineaires II, Trans. Amer. Math. Soc., 6(1905), 134-140.
[3] Sur quelques points du Calcul fonctionnel, Rend. Circ. mat. Palermo, 22(1906), 1-74.
[4] Sur les operations lineaires III, Trans. Amer. Math. Soc., 8(1907), 433-446.
[5] Sur les ensembles de fonctions et les operations lineaires, C. R. Acad. Sci. Paris,
144(1907a), 1414-1416.

1. Fredholm
[1] Sur une nouvelle methode pour la resolution du probleme de Dirichlet, Kong. Vetenskaps-
Akademiens Forh. Stockholm, (1900),39-46.
[2] Sur une classe d'equations fonctionnelles, Acta math., 27(1903), 365-390.

K. Friedrichs
[1] On certain inequalities and characteristic value problems for analytic functions and for
functions of two variables, Trans. Amer. Math. Soc., 41(1937), 321-364.

D. Gaier
[1] Lectures on Complex Approximation, translated by R. McLaughlin, Birkhauser,
Boston, 1987.

W. B. Gearhart and M. Koshy


[1] Acceleration schemes for the method of alternating projections, J. Compo Appl. Math.,
26(1989), 235-249.

M. Golomb
[1] Lectures on Theory of Approximation, Argonne National Laboratory, Applied
Mathematics Division, 1962.

B. K. Goodrich and A. Steinhardt


[1] £2 spectral estimation, SIAM J. Appl. Math., 46(1986),417-428.

R. Gordon, R. Bender, and G. T. Herman


[1] Algebraic reconstruction techniques (ART) for three-dimensional electron microscopy
and X-ray photography, J.Theoretical BioI., 29(1970), 471-48l.

M. S. Gowda and M. Teboulle


[1] A comparison of constraint qualifications in infinite-dimensional convex programming,
SIAM J. Control Optimiz., 28(1990), 925-935.

J. P. Gram
[1] Om Rackkendvilklinger bestemte ved Hjaelp af de mindste Kvadraters Methode,
Copenhagen, 1879. [Translated in: Uber die Entwicklung reeler Funktionen in Reihen
mittels der Methode der kleinsten Quadrate, J. Reine Angewandte Mathematik, 94
(1883),41-73.]
322 REFERENCES

1. Grattan-Guiness
[1] The Development of the Foundations of Mathematical Analysis from Euler
to Riemann, M.LT. Press, Cambridge, MA, 1970.

C. W. Groetsch
[1] Generalized Inverses of Linear Operators, Marcel Dekker, New York, 1977.

L. G. Gubin, B. T. Polyak, and E. V. Raik


[1] The method of projections for finding the common point of convex sets, USSR
Computational Mathematics and Mathematical Physics, 7(5) (1957), 1-24.

H. Hahn
[1] Uber lineare Gleichungssysteme in linearen Riiurnen, J. Reine Angew. Math., 157(1927),
214-229.

P. R. Halmos
[1] Introduction to Hilbert Space and the Theory of Spectral Multiplicity,
Chelsea Publ. Co., New York, 1951.
[2] A Hilbert Space Problem Book, Second ed., Springer-Verlag, New York, 1982.

1. Halperin
[1] The product of projection operators, Acta Sci. Math. (Szeged), 23(1952), 95--99.

F. Hausdorff
[1] Zur Theorie der linearen metrischen Riiume, J. Reine Angew. Math., 157(1932),294-311.

H. E. Heine
[1] Uber trigonometrische Reihen, Journal fUr die Reine und Angewandte Mathematik, 71
(1870), 353-365.
[2] Die Elemente der Functionenlehre, Journal fUr dei Reine und Angewandte Mathematik,
74(1872), 172-188.

M. R. Hestenes
[1] Relative self-adjoint operators in Hilbert space, Pacific J. Math., 11(1961), 1315-1357.

D. Hilbert
[1] Grundziige einer allgemeinen Theorie der linearen Integralgleichungen, I, Nachrichten
Akad. Wiss. Gottingen. Math.-Phys. Kl., 1904, 49-91.
[2] Grundziige einer allgemeinen Theorie der linearen Integralgleichungen, II, Nachrichten
Akad. Wiss. Gottingen. Math.-Phys. Kl., 1905, 213-259.
[3] Grundziige einer allgemeinen Theorie der linearen Integralgleichungen, III, Nachrichten
Akad. Wiss. Gottingen. Math.-Phys. Kl., 1905,307-338.
[4] Grundziige einer allgemeinen Theorie der linearen Integralgleichungen, IV, Nachrichten
Akad. Wiss. Gottingen. Math.-Phys. Kl., 1906, 157-227.
[5] Grundziige einer allgemeinen Theorie der linearen Integralgleichungen, V, Nachrichten
Akad. Wiss. Gottingen. Math.-Phys. Kl., 1906, 439-480.
[6] Grundziige einer allgemeinen Theorie der linearen Integralgleichungen, VI, Nachrichten
Akad. Wiss. Gottingen. Math.-Phys. Kl., 1910,355-417.
(These six articles were published in book form by Teubner, Leipzig, 1912. They were
reprinted by Chelsea Pub. Co., New York, 1953.)

T. H. Hildebrandt
[lJ On uniform limitedness of sets of functional equations, Bull. Amer. Math. Soc., 29(1923),
309-315.

R. A. Hirschfeld
[lJ On best approximation in normed vector spaces, II, Nieuw Archief voor Wiskunde (3),
VI (1958), 99-107.

R. B. Holmes
[lJ A Course on Optimization and Best Approximation, Lecture Notes in
Mathematics, #257, Springer-Verlag, New York, 1972.
REFERENCES 323

G. N. Hounsfield
[IJ Computerized transverse axial scanning (tomography): Part I Description of system,
British J. Radio!., 46(1973), 1016-1022.

H. S. Hundal
[IJ Generalizations of Dykstra's Algorithm, doctoral dissertation, The Pennsylvania State
University, 1995.

H. Hundal and F. Deutsch


[IJ Two generalizations of Dykstra's cyclic projections algorithm, Math. Programming, 77
(1997), 335-355.

W. A. Hurwitz
[IJ On the pseudo-resolvent to the kernel of an integral equation, Trans. Amer. Math. Soc.,
13(1912), 405-418.

G. Illiev and W. Pollul


[IJ Convex interpolation with minimal Loo-norm of the second derivative, Math Z., 186(1984),
49-56.
[2J Convex interpolation by functions with minimal Lp-norm (1 < p < ()()) of the kth
derivative, in Proceedings of the Thirteenth Spring Conference of the Union of
Bulgarian Mathematicians, Sunny Beach, April 6-9, 1984a.

L. Irvine
[IJ Constrained Interpolation, master's thesis, Old Dominion University, 1985.

L. D. Irvine, S. P. Marin, and P. W. Smith


[IJ Constrained interpolation and smoothing, Constr. Approx., 2(1986), 129-151.

D. Jackson
[IJ Uber die Genauigkeit der Annaherung Stetiger Funktionen Durch Ganze Rationale
Funktionen Gegebenen Grades und Trigonometrische Summen Gegebener Ordnung,
dissertation, Giittingen, 1911.
[2J The Theory of Approximation, Amer. Math. Soc. Colloq. Publications, Vo!' XI,
New York, 1930.

B. Jessen
[IJ To Saetninger om konvekse Punktmaengder, Mat. Tidsskrift B, 1940, 66-70.

M. Jiang
[IJ On Johnson's example of a nonconvex Chebyshev set, J. Approx. Theory, 74(1993),
152-158.

G. G. Johnson
[IJ A nonconvex set which has the unique nearest point property, J. Approx. Theory,
51(1987),289-332.
[2J Chebyshev foam, Topology and its Applications, 20(1998), 1-9.

P. Jordan and J. von Neumann


[IJ On inner products in linear metric spaces, Ann. Math., 36(1935), 719-723.

L. V. Kantorovich and V. 1. Krylov


[IJ Approximate Methods of Higher Analysis, translated by C. D. Benster,
Interscience Pub!., New York, 1958.

T. Kato
[IJ Perturbation Theory for Linear Operators, second edition, Springer, New York,
1984.

S. Kayalar and H. Weinert


[IJ Error bounds for the method of alternating projections, Math. Control Signal Systems,
1 (1988), 43-59.
324 REFERENCES

B. F. Kelly, III
[1 J The convexity of Chebyshev sets in finite dimensional normed linear spaces, master's
paper, The Pennsylvania State University, 1978.

V. Klee
[IJ Convex bodies and periodic homeomorphisms in Hilbert space, Trans. Amer. Math.
Soc., 74 (1953), 10-43.
[2J Convexity of Chebyshev sets, Math. Annalen, 142(1961), 292-304.
[3J On a problem of Hirschfeld, Nieuw Archief voor Wiskunde (3), XI (1963), 22-26.
[4J Remarks on nearest point in normed linear spaces, Proc. Colloq. Convexity, Univ. of
Copenhagen, 1966, 168-176.

M. Kline
[IJ Mathematical Thought from Ancient to Modern Times, Oxford Univ. Press,
New York, 1972.

A. N. Kolmogorov and S. V. Fomin


[IJ Elements of the Theory of Functions and Functional Analysis, Volume 1
Metric and Normed Spaces, translated by L. F. Boron, Graylock Press, Rochester,
NY, 1957.

P. P. Korovkin
[1 J On the convergence of linear positive operators in the space of continuous functions,
Doklady Akademii Nauk SSSR, 90(1953), 961-964 (Russian).
[2J Linear Operators and Approximation Theory, Hindustan Pub!. Corp.,
Delhi, 1960.

M. Krein and D. Milman


[IJ On extreme points of regular convex sets, Studia Math., 9(1940), 133-138.

M. Kritikos
[IJ Sur quelques proprietes des ensembles convexes, Bull. Math. de la Soc. Roumaine des
Sciences, 40(1938), 87-92.

H. Kuhn
[IJ Ein elementarer Beweis des Weierstrasschen Approximationssatzes, Archiv der
Mathematik, 15(1964), 316-317.

C. Kuratowski
[IJ La propriete de Baire dans les espaces metriques, Fund. Math., 16(1930),390-394.
[2J Topologie, Vo!' I, Warzawa, 1958.
[3J Topologie, Vol. II, Warzawa, 1961.

J. L. Lagrange
[1 J Oeuvres, t. 3, Gauthier-Villars, Paris, 1869.

E. Landau
[IJ Uber die Approximation einer stetigen Funktionen durch eine ganze rationale Funktion,
Rendiconti del Circolo Matematico di Palermo, 25(1908), 337-345.

R. E. Langer (editor)
[IJ On Numerical Approximation, Univ. Wisconsin Press, Madison, 1959.

P.-J. Laurent
[IJ Approximation et Optimisation, Hermann, Paris, 1972.

C. L. Lawson and R. J. Hanson


[IJ Soving Least Squares Problems, Prentice-Hall, Inc., Englewood Cliffs, NJ, 1974.

H. Lebesgue
[IJ Sur l'approximation des fonctions, Bull. Sci. MatMmatique, 22(1898), 278-287.
REFERENCES 325

[2[ Le<;ons sur l'integration et la recherche des fonctions primitives, Gauthier-Villars, Paris,
1904 (second edition 1928).

M. Lerch
[1[ On the main theorem of the theory of generating functions (in Czech), Rozpravy Ceske
Akademie, no. 33, Volume I, (1892) 681-685.
[2[ Sur un point de la tMorie des fonctions generatrices d'Abel, Acta Mathematica, 27(1903),
339-352.

B. Levi
[1[ Sui principio de Dirichlet, Rend. Circ. Math. Palermo, 22(1906), 293-360.

W.Li
[1[ Private communication, 1998.

W. Li, P. M. Pardalos, and C. G. Han


[1[ Gauss-Seidel method for least-distance problems, J. Optim. Theory Appl., 75(1992),
487-500.

W. Li and J. Swetits
[1[ A new algorithm for solving strictly convex quadratic programs, SIAM J. Optim.,
7(1997), 595-619.

W. A. Light and E. W. Cheney


[1[ Approximation Theory in Tensor Product Spaces, Lecture Notes in Mathematics,
# 1169, Springer-Verlag, New York, 1985.
G. G. Lorentz, M. von Golitschek, and Y. Makovoz
[1[ Constructive Approximation Advanced Problems, Springer, New York, 1996.

H. Liiwig
[1[ Komplexe euklidische Raume von beliebiger endlicher oder unendlicher Dimensionszahl,
Acta Sci. Math. Szeged, 7(1934), 1-33.

D. G. Luenberger
[1[ Optimization by Vector Space Methods, John Wiley & Sons, New York, 1969.

H. Massam
[1[ Optimality conditions for a cone-convex programming problem, J. Australian Math.
Soc., Ser. A, 27(1979), 141-162.

S. Mazur
[1[ Uber konvexe Mengen in linearen normierten Raumen, Studia Math., 4(1933), 70-84.
G. Meinardus
[1[ Approximation of Functions: Theory and Numerical Methods, translated by
L. L. Schumaker, Springer-Verlag, New York, 1967.

C. Micchelli, P. Smith, J. Swetits, and J. Ward


[1[ Constrained Lp-approximation, Constr. Approx., 1(1985),93-102.

C. Micchelli and F. Utreras


[1[ Smoothing and interpolation in a convex subset of Hilbert space, SIAM J. Sci. Statist.
Computing, 9(1988), 728-746.

G. Mittag-Leffler
[1[ Sur la representation analytique des fonctions d'une variable reelle, Rend. Cire. Math.
Palermo, 14(1900), 217-224.

E. H. Moore
[1[ On the reciprocal of the general algebraic matrix, Bull. Amer. Math. Soc., 26(1920),
394-395.
[2[ General Analysis, Memoirs Amer. Philos. Soc., 1(1935), 147-209.
326 REFERENCES

J. J. Moreau
[1] Decomposition orthogonale dans un espace hilbertien selon deux cones mutuellement
polaires, C. R. Acad. Sci., Paris, 255(1962), 238-240.

T. S. Motzkin
[1] Sur quelques proprietes caracteristiques des ensembles convexes, Rend. Acad. dei Lincei
(Roma), 21(1935), 562-567.
[2] Sur quelques proprietes caracteristiques des ensembles bornes non convexes, Rend. Acad.
dei Lincei (Roma), 21(1935a), 773-779.

B. Mulansky and M. Neamtu


[1] Interpolation and approximation from convex sets, J. Approx. Theory, 92(1998), 82-100.

C. Muntz
[1] Uber den Approximationssatz von Weierstrass, H. A. Schwarz Festschrift,
Mathematische Abhandlungen, Berlin, 1914, 303-312.

F. J. Murray and J. von Neumann


[1] On rings of operators. I, Ann. Math., 37(1936), 116--229.

H.Nakano
[1] Spectral theory in the Hilbert space, Japan Soc. Promotion Sc., 1953, Tokyo.

T. D. Narang
[1] Convexity of Chebyshev sets, Nieuw Archief Voor Wiskunde, 25(1977), 377-402.

M. Z. Nashed
[1] A decomposition relative to convex sets, Proc. Amer. Math. Soc., 19(1968), 782-786.
[2] Generalized Inverses and Applications, Acad. Press, New York, 1976.

M. Z. Nashed and G. F. Votruba


[1] A unified approach to generalized inverses of linear operators: I. Algebraic, topological
and projectional properties, Bull. Amer. Math. Soc., 80(1974), 825-830.
[2] A unified approach to generalized inverses of linear operators: II. Extremal and proximal
properties, Bull. Amer. Math. Soc., 80(1974), 831-835.

J. von Neumann
[1] Allegemeine Eigenwerttheorie Hermitescher Funktionaloperatoren, Math. Annalen, 102
(1929a),49--131. (In collected works: Vol. II, pp. 3-85.)
[2] Sur Algebra der Funktionaloperatoren und Theorie der normalen Operatoren, Mat.
Annalen, 102(1929b), 370-427. (In collected works: Vol. II, pp. 86-143.)
[3] Uber adjungierte Funktionaloperatoren, Ann. Math., 33(1932), 294-310. (In collected
works: Vol. II, pp. 242-258.)
[4] Functional Operators-Vol. II. The Geometry of Orthogonal Spaces, Annals
of Math. Studies #22, Princeton University Press, Princeton, NJ, 1950. [This is a reprint
of mimeographed lecture notes first distributed in 1933.]
[5] Collected Works, Vol. II, Operators, Ergodic Theory and Almost Periodic Functions in
a Group, Permagon Press, New York, 1961.

L. Nirenerg
[1] Functional Analysis, lecture notes by L. Sibner, New York University, 1961.

A.Ocneanu
[1] personal correspondence, March, 1982.

W. Osgood
[1] Non uniform convergence and the integration of series term by term, Amer. Jour. Math.,
19(1897), 155-190.

G. Pantelidis
[1] Konvergente Iterationsverfahren fUr fiach konvexe Banachriiume, Rhein.-Westf. Institut
fUr Instrumentelle Mathematik Bonn (lIM), Serie A, 23(1968), 4-8.
REFERENCES 327

S. Paszkowski
[lJ On approximation with nodes, Rozprawy Mathematyczne XIV, Warszawa, 1957.

R. Penrose
[lJ A generalized inverse for matrices, Proc. Cambridge Philos. Soc., 51(1955), 406---413.

R. R. Phelps
[lJ Convex sets and nearest points, Proc. Amer. Math. Soc., 8(1957), 790-797.

E. Picard
[lJ Sur la representation approchee des fonctions, Comptes Rendus, 112(1891), 183-186.

A. Pinkus
[lJ Weierstrass and approximation theory, J. Approx. Theory (to appear).

M. J. D. Powell
[lJ A new algorithm for unconstrained optimization, in Nonlinear Programming (edited
by J. B. Rosen, o. L. Mangasarian, and K. Ritter), Academic Press, New York, 1970.
[2J Approximation Theory and Methods, Cambridge Univ. Press, Cambridge, 1981.

J. B. Prolla
[lJ Weierstrass-Stone, the Theorem, Vol. 5 in the series "Approximation & Optimization,"
Peter Lang, Frankfurt, 1993.

R. Rado
[lJ Note on generalized inverses of matrices, Proc. Cambridge Philos. Soc., 52(1956),
600-601.

T. J. Ransford
[lJ A short elementary proof of the Bishop-Stone-Weierstrass theorem, Math. Proc.
Cambridge Philos. Soc., 96(1984), 309-311.

C. R. Rao
[lJ A note on a generalized inverse of a matrix with applications to problems in mathematical
statistics, J. Royal Statist. Soc. Ser. B., 24(1962), 152-158.

C. R. Rao and S. K. Mitra


[lJ Generalized Inverse of Matrices and Its Applications, John Wiley & Sons, New
York, 1971.

W. T. Reid
[lJ A simple optimal control problem involving approximation by monotone functions, J.
Optim. Theory Appl., 2(1968), 365-377.
[2J Generalized inverses of differential and intergral operators, in Proceedings of the
Symposium on Theory and Applications of Generalized Inverses of Matrices
(edited by T. L. Boullion and P .. L. Odell), Texas Tech Press, Lubbock, Texas, 1968,
pp.1-25.

F. Rellich
[lJ Spektraltheorie in nichtseparabeln Riiumen, Math. Ann., 110(1935),342-356.

J. R. Rice
[lJ The Approximation of Functions, Volume 1 Linear Theory, Addison-Wesley,
Reading, MA, 1964.
[2J The Approximation of Functions, Volume 2 Nonlinear and Multivariate
Theory, Addison-Wesley, Reading, MA, 1969.

F. Riesz
[lJ Sur une espece de geometrie analytiques des systemes de fonctions sommables, C. R. Acad.
Sci. Paris, 144(1907), 1409-1411.
[2J Untersuchungen liber Systeme integrierbarer Funktionen, Math. Annalen 69(1910),
449-497.
328 REFERENCES

[3] Les Systemes d'Equations Lineaires it une Infinite d'Inconnues,


Gauthier-Villars, Paris, 1913.
[4] Dber lineare Funktionalgleichungen, Acta Math., 41(1918), 71-98.
[5] Zur Theorie des Hilbertschen Raumes, Acta Sci. Math. Szeged, 7(1934),34-38.

F. Riesz and B. Sz.-Nagy


[1] Functional Analysis, translated by L. F. Boron, Ungar Pub!. Co., New York, 1955.

T. J. Rivlin
[IJ An Introduction to the Approximation of Functions, Blaisdell, Waltham, MA,
1969.

R. T. Rockafellar
[1J Duality and stability in extremum problems involving convex functions, Pacific J. Math.,
21(1967), 167-187.
[2] Convex Analysis, Princeton Univ. Press, Princeton, NJ, 1970.

C. Runge
[IJ Zur Theorie der eindeutigen analytischen Funktionen, Acta Mathematica, 6(1885),
229-245.
[2J Dber die Darstellung willkiirlicher Functionen, Acta Mathematica, 7(1885a), 387-392.

A. Sard
[IJ Linear Approximation, Math. Surveys, No.9, Amer. Math. Soc., Providence, 1963.

J. Schauder
[1] Dber die Umkehrung linearer, stetiger Funktionaloperationen, Studia Math., 2(1930),
1-6.

E. Schmidt
[IJ Dber die Auflosung linearer Gleichungen mit unendlich vielen Unbekannten, Rend. Circ.
mat. Palermo, 25(1908), 53-77.

A. Schonhage
[1J Approximationstheorie, W. de Gruyter & Co., New York, 1971.

H. A. Schwarz
[1] Gesammelte mathematische Abhandlungen, 2 vols., Springer, Berlin, 1890.

H. S. Shapiro
[IJ Topics in Approximation Theory, Lecture Notes in Math., # 187, Springer-Verlag,
New York, 1971.

C. L. Siegel
[IJ Dber die analytische Theorie der quadratischen Formen III, Ann. Math., 38(1937),
212-291.

1. Singer
[IJ Remarque sur un theoreme d'approximation de H. Yamabe, Atti Accad. Naz. Lincei
Rend. C!. Sci. Fis. Mat. Nat., 26(1959), 33-34.
[2J J Best Approximation in Normed Linear Spaces by Elements of Linear
Subspaces, Springer-Verlag, New York, 1970.
[3J The Theory of Best Approximation and Functional Analysis, CBMS #13,
SIAM, Philadelphia, 1974.

R. Smarzewski
[1] Iterative recovering of orthogonal projections, preprint, 1986.

K. T. Smith, D. C. Solmon, and S. L. Wagner


[1 J Practical and mathematical aspects of the problem of reconstructing objects from
radiographs, Bul!. Amer. Math. Soc., 83(1977), 1227-1270.
REFERENCES 329

P. W. Smith and J. D. Ward


[1] Distinguished solutions to an Loo minimization problem, Approx. Theory App!., 4(1988),
2(}-40.

P. W. Smith and H. Wolkowicz


[1] A nonlinear equation for linear programming, Math. Programming, 34(1986), 235-238.

G. A. Soukhomlinoff
[1] Uber Fortsetzung von linearen Funktionalen in linearen komplexen Riiumen und linearen
Quaternionriiumen, Mat. Sbornik N.S., 45(1938), 353-358.

J. F. Steffensen
[1] Interpolation, Second ed., Chelsea Pub!. Co., New York, 1950.

E. Steinitz
[1] Bedingt konvergente Reihen und konvexe Systeme I, J. Reine Angew. Math., 143(1913),
128-175.
[2] Bedingt konvergente Reihen und konvexe Systeme II, J. Reine Angew. Math., 144(1914),
1--40.
[3] Bedingt konvergente Reihen und konvexe Systeme III, J. Reine Angew. Math., 146(1916),
1-52.

W. J. Stiles
[11 Closest point maps and their products, Nieuw Archief voor Wiskunde (3), XIII (1965a),
19-29.
[2] Closest point maps and their products II, Nieuw Archief voor Wiskunde (3), XIII
(1965b), 212-225.
[3] A solution to Hirschfeld's problem, Nieuw Archief voor Wiskunde (3), XIII (1965c),
116-119.

M. H. Stone
[1] Applications of the theory of boolean rings to general topology, Trans. Amer. Math.
Soc., 41(1937), 375--48l.
[2] The generalized Weierstrass approximation theorem, Mathematics Magazine, 21(1948),
167-183, 237-254.

J. J. Swetits, S. E. Weinstein, and Y. Xu


[1] Approximation in Lp[O, 1] by n-convex functions, Numer. Funet. Ana!. Optimiz., 11(1990),
167-179.

J. Szabados and P. Vertesi


[1] Interpolation of Functions, World Scientific Pub!., New Jersey, 1990.

O. Szasz
[1] Uber die Approximation stetiger Funktionen durch lineare Aggregate von Potenzen,
Math. Anna!., 77(1916), 482-496.

1. G. Tsar'kov
[1] Bounded Chebyshev sets in finite-dimensional Banach spaces, Math. Notes, 36(1)(1984),
530--537.

Y. Y. Tseng
[1] The Characteristic Value Problem of Hermitian Functional Operators in a Non-Hilbert
Space, doctoral dissertation, Univ. Chicago, 1933.
[2] Sur les solutions des equations operatrices fonctionnelles entre les espaces unitaires.
Solutions extremales. Solutions virtuelles., C. R. Acad. Sci. Paris, 228(1949a), 64()-{l4l.
[3] Generalized inverses of unbounded operators between two unitary spaces, Dok!. Akad.
Nauk SSSR (N.S.), 67(1949b), 431-434.
[4] Properties and classification of generalized inverses of closed operators, Dok!. Akad.
Nauk SSSR (N.S.), 67(1949c), 607-610.
[5] Virtual solutions and general inversions, Uspehi. Mat. Nauk (N.S.), 11(1956),213-215.
330 REFERENCES

F. A. Valentine
[1] Convex Sets, McGraw-Hill, New York, 1964.

C. J. de la Vallee Poussin
[1] Sur l'approximation des fonctions d'une variable reelle et leurs dEirivees par de polynomes
et des suites limitees de Fourier, Bulletin de l'AcadEimie Royal de Belgique, 3(1908),
193-254.

L. P. Vlasov
[1] CebySev sets in Banach spaces, Sov. Math. Dokl., 2(1961), 1373-1374.
[2] Chebyshev sets and approximatively convex sets, Math. Notes, 2(1967), 600-605.
[3] On Cebysev sets, Sov. Math. Dokl., 8(1967a), 401-404.
[4] Approximative properties of sets in normed linear spaces, Russian Math. Surveys,
28(1973), 1-66.

J. L. Walsh
[1] Interpolation and Approximation by Rational Functions in the Complex
Domain, Amer. Math. Soc. Colloquium Publications, vol. 20, Providence, RI, 1956.

J. Ward
[1] Some constrained approximation problems, in Approximation Theory V (edited by
C. K. Chui, L. L. Schumaker, and J. D. Ward), Academic Press, New York, 1986,
211-229.

G. A. Watson
[1] Approximation Theory and Numerical Methods, Wiley, New York, 1980.

H. Weyl
[1] Raum . Zeit· Materie, Springer-Verlag, Berlin, 1923.

K. Weierstrass
[1] Uber die analytische darstellbarkeit sogenannter willkiirlicher Funktionen einer reelen
Veraenderlichen, Sitzungsberichte der Akademie zu Berlin, 1885, 633-639 and 789-805.

N. Wiener
[1] On the factorization of matrices, Comment. Math. Helv., 29(1955), 97-111.

S. Xu
[1] Estimation of the convergence rate of Dykstra's cyclic projections algorithm in polyhe-
dral case, Acta Math. Appl. Sinica, 16 (2000),217-220.

W. Wolibner
[1] Sur un polynome d'interpolation, Colloq. Math., 2(1951), 136-137.

H. Yamabe
[1] On an extension of the HeIly's theorem, Osaka Math. J., 2(1950), 15-17.

N. Young
[1] An Introduction to Hilbert Space, Cambridge Univ. Press, Cambridge, 1988.

E. H. Zarantonello
[1] Projections on convex sets in Hilbert space and spectral theory. Part I. Projections on
convex sets. Part II. Spectral theory, in Contributions to Nonlinear Functional
Analysis (edited by E. H. Zarantonello), Academic Press, New York, 1971, pp. 237-424.
INDEX

Achieser, 32, 152, 315 from hyperplanes, 97~101


active index set, 115, 241 from infinite-dimensional
adjoint subspaces, 60
of linear operator, 171~174, 191 uniqueness of, 23, 35
of matrix, 173 bicompact (= compact), 41
affine biorthogonal sets, 291, 292
hull, 31, 248 Bishop, 153
set, 31, 247 Bjerhammar, 192,317
Alexandroff, 40, 315 Bohman, 152, 153,317
alternating projections, method of Bohnenblust, 123, 317
(Chapter 9),195,216,217 Bolzano, 41, 317
rate of convergence of, 217 ~Weierstrass theorem, 41
rate of convergence of (for affine de Boor, 283, 317
sets), 222 Borel, 41, 317
rate of convergence of (for two Borwein, 234, 235, 284, 285, 316, 317
subspaces), 201, 234 Boullion, 192, 317
Amir, 19, 123, 315 boundary of a set, 27
Anderson, 192, 284, 315 bounded
angle linear functional, (Chapter 6), 90, 93
between subspaces, 197, 234 linear operator, 78
between vectors, 10 continuity of, 78
anti-proximinal, 100~ 101 set, 12
Aronszajn, 32, 69,87, 233, 234,235,315 bounded inverse theorem, 169
Ascoli, 123, 315 boundedly compact, 38, 41, 304
Asplund, 308, 315 Boyle, 213, 234, 317
Atkinson, 192, 315 Boyle-Dykstra theorem, 213, 234, 276
Atlestam, 233, 315 with affine sets, 216
Aubin, 87, 315 Boznay, 234, 317
Br0nsted, 308, 317
Baire, 191, 315 Brosowski, 153, 317
Baire's theorem, 164, 191 Brown, 308, 309, 318
Balaganskii, 308, 315 Buniakowsky, 19, 318
Banach, 123, 191, 234, 316 Bunt, 307, 318
Bauschke, 234, 235, 285, 316 Burckel, 153, 318
Bender, 233, 321 Busemann, 307, 318
Ben-Israel, 191, 192, 316
Berens, 308, 316 Cakon, 153, 318
Bernoulli's inequality, 139, 153 Cauchy, 41, 318
Bernstein, 152, 316 determinant formula, 66, 143
polynomials, 151, 152 sequence, 12, 93
Bessel's inequality, 55, 56 characteristic function, 271
best approximation, (Chapter 2), 21 characterization of best approximations,
characterization of, 43 (Chapter 4)
existence of, 33, 34 from affine sets, 215
from affine sets, 137, 215 from convex cones, 48, 49
from convex sets, 43 from convex sets, 43
from half-spaces, 225 from finite-dimensional subspaces, 50
332 INDEX

from infinite-dimensional at a point, 71


subspaces, 60 Lipschitz - , 71
from polyhedra, 115 uniformly - , 71
from subspaces, 50 control problem, 2, 28, 139
Chebyshev convergent sequence, 10
affine set, 138 weakly - , 203
approximatively compact - , 304 convex
boundedly compact - , 304 cone, 24, 48
closed convex, 35 hull, 31,
convex cone, 74 midpoint - , 306
half-space, 108 regression, 228
hyperplane, 99 set, 22
polyhedral set, 111 Cousin, 41, 318
polynomials, 53
set, 21, 25, 34, 35, 37 Davis, 32, 70, 152, 298, 318
convexity of, (Chapter 12) Day, 233, 319
in finite-dimensional spaces, 37, 306 dense, 92, 141, 146, 164, 294
subspace, determinant
finite-co dimensional - , 135 Gram - , 65, 129, 132, 151, 187
finite-dimensional - , 37 Vandermonde -,297
sun, 302 Desoer, 192, 319
Chene~ 32, 69, 70, 318, 325 Deutsch, 69, 87, 123, 151, 153, 233, 234,
Chui, 69, 283, 284, 285, 318 235, 283, 284, 285, 298, 299, 308, 315,
closed 316,317,318,319,320,323
ball, 11 Dieudonne, 19, 123, 320
graph, 170 dimension of a set, 248
range, 180 direct sum, 76, 97
set, 11 Dirichlet, 152, 233
sum of subspaces, 199, 222 discrete spaces, 17
closed graph theorem, 170 distance, 7
closed range theorem, 168 between vectors, 7
closure of a set, 12 to convex cone, 106, 127
codimension, 133 to convex set, (Chapter 7), 105, 125
compact, 38, 41 to finite-codimensional subspace, 136
approximatively - , 34, 35 to finite-dimensional subspace, 64,
boundedly -,38, 41,304 130, 144
complete to half-space, 108
inner product space, 14, 169, 174 to hyperplane, 98
subset, 15, 34, 36 to set, 72
complex inner product space, 17-18 to subspace, 106, 127
cone Dixmier, 234, 320
of positive functions, 85 Dontchev, 284, 285, 320
translates of, 64 dual cone, 44, 46-48
conical hull, 31, 45 characterization of best
constrained interpolation, (Chapter 10) approximations, 44
by positive cone, 273 of images, 256
continuous, 71 to a polyhedron, 115
INDEX 333

dual space, 90,158 injective - , or injection, 160, 187


completeness of, 158 inverse - , 161
Duffin, 192, 315 subdifferential - , 279
Dunford, 41, 123, 191,320 support - , 278
Dykstra, 234, 320 surjective -, or surjection, 160
Dykstra's algorithm, 207, 234, 276, 283,
285, 317 Gearhart, 235, 321
generalized inverse, 179, 191
Effimov, 32, 41, 307, 308, 309, 320 characterization of, 182
Eidelheit, 123, 283, 320 for metric projection, 184
equivalent inner product spaces, 118, formula for, 182
312 linearity and boundedness of, 181
error of approximation, (Chapter 7) for zero operator, 184
Euclidean space, 4, 19 generalized solution, (Chapter 8), 155
existence of best approximations, characterization of, 178
(Chapter 3), 33, 34 minimal norm -, 179
extremal subset, 262, 264, 265, 270 characterization of, 179
of cones, 269 generating set, 268, 284
extreme point, 262 Goldstein, 69, 318
Goodrich, 285, 321
Faber, 153, 320 Gordon, 233, 321
Farkas lemma, 120 Gowda, 285, 321
Favard, 283, 320 Gram, 51, 69, 151, 321
Fejer, 152, 153, 320 determinant, 65, 129, 132, 151, 187
Fenchel, 69,321 inequality, 132
finite-codimensional subspace, 133 matrix, 51, 65, 129
characterization of, 133 -Schmidt orthonormalization, 51, 52,
finite-dimensional subspace, 36 69
is Chebyshev, 37 graph of a mapping, 170
is complete, 36 Grattan-Guiness, 41, 322
is closed, 36 Greville, 191, 192, 316
finitely-generated, 109 Groetsch, 191, 192, 322
cones, 110 Gubin, 235, 322
fiat (see affine set) Guiness,41
Fourier, 152
series, 56, 59, 152 Hadamard
coefficients, 57, 59, 69 determinent inequality, 65
Franchetti, 233, 234, 321 Hahn, 123, 322
Frechet, 19, 89, 123, 321 Hahn-Banach theorem, 94, 123
-Riesz representation theorem, 89, 94 half-spaces, 102, 107
Fredholm, 19, 321 Chebyshev - , 108
Friedrichs, 234, 321 Halperin, 217, 234, 322
function (mapping, operator) Han, 284, 325
bijective -, or bijection, 160, 187 hat mapping, 91
characteristic - , 271 Hausdorff, 191
convex -, 278 Heine, 41, 322
indicator -, 279 -Borel theorem, 41
334 INDEX

Herman, 233, 321 Jessen, 307, 323


Hermite (or osculatory) interpolation, Jiang, 308, 323
296 Johnson, 308, 323
Hestenes, 191, 322 Jordan, 19, 323
Hilbert, 19, 191, 234, 322
matrix, 66 Kato, 69, 234, 323
space, 14-15, 19, 87 Kayalar, 234, 235, 323
Hildebrandt, 191, 322 Kelly, 308, 324
Hirschfeld, 233,322 kernel,97
Hopf, 40, 315 Killgrove, 32
Hounsfield, 233, 323 Klee, 41, 233, 307, 308, 309, 324
Hundal, 234, 235, 316, 319, 323 Kline, 41, 323
hyperplane, 97 Konstantinov, 284, 285,320
Chebyshev - , 99 Korovkin, 152, 324
proximinal -, 99 Koshy, 235
separating - , 102 Krein, 152, 284, 315, 324
Kritikos, 307, 324
idempotent, 193, 218 Kuhn, 153, 324
identity operator, 159, 173 Kuratowski, 40, 191, 324
Illiev, 284, 285, 320, 323
indicator function, 279 Lagrange, 19, 324
infinite sum of sets, 68 interpolation, 292
inner product, 2 Landau, 152, 324
complex, 17 Laurent, 32, 87, 324
spaces (complex), 17-18 Lebesgue, 41, 152, 324
spaces (real), 3, 4-7 Legendre polynomials, 53
integral operator, 174 length of vector, 7
interior of a set, 12 Lerch, 152, 325
relative - , 248, 249, 250, 251, 253, Levi, 41, 325
254, 257, 265, 266 Lewis, 235, 285, 316, 317
interpolates, 287 Li, 123, 283,284,285,315, 316,319,325
interpolation, 287 Light, 70, 234, 318, 321, 325
general problem of, 287 limit of sequence, 10
Hermite (or osculatory) -, 296 line segment (or interval), 301
polynomial - , 290 linear
possibility of, 288 equations, 1, 28, 62, 226
Taylor - , 292 functional, 78, 90
interpolation theorem, 111, 291 boundedness of, 84
invariance unbounded -, 84
by scalar multiplication, 25 inequalities, 227
by translation, 25 isometry, 118
inverse operator, 161, 176 operator equations, 155
Irvine, 284, 285, 323 variety (see affine set)
isotone regression, 85, 228 linear operator (linear mapping, linear
Ivert, 284, 315 function), 77-78
boundedness below of, 163
Jackson, 152, 323 bounded ness of, 78, 82
INDEX 335

equations, 155 Nakano, 233, 326


nonnegative, 79, 188 Narang, 308, 326
norm of, 78 Nashed, 87, 192, 326
representation of, 175 Neamtu, 284, 326
L6wig, 19, 70, 325 von Neumann, 19, 70, 87, 192, 233,
Luenberger, 32, 283, 325 323, 326
Nirenberg, 123, 151, 326
Mallat, 70, 318 nonexpansive function, 71
nonnegative linear operator, 79, 188,
mapping (see function)
231
Marin, 284, 285, 323
norm, 3, 90
Maserick, 123, 319
attaining, 95, 97, 111, 295
Massam, 284, 325
properties of, 7
matrix of linear operator, 158
normal
maximal element, 311
equations, 50-51
maximal orthonormal set, 59, 60
linear operator, 176
Mazur, 283, 325 normed linear space, 7, 90
McCabe, 123, 320 null space of mapping, 160
metric projection (Chapter 5), 87
continuity of, 73, 304
Ocneanu, 122, 326
firm nonexpansivenesss of, 73
Odell, 192, 317
idempotency of, 72, 79
open
linearity of, 79
ball, 11
monotonicity of, 72 cover, 41
nonexpansiveness of, 72, 302 set, 11
nonnegativity of, 79 open mapping theorem, 166
onto convex cones, 74 operator (see function)
positive homogeneity of, 74 Opial's condition, 232
onto convex sets, 71 orthogonal, 8, 52
onto subspaces, 76 complement, 44, 233
self-adjointness of, 79, 173 projection, 80, 87, 188, 194
strict nonexpansiveness of, 73 set, 8
uniform continuity of, 73 sum, 76
Micchelli, 284, 285, 325 orthonormal, 8
midpoint convex, 306 basis, 60, 61
Milman, 284, 324 set, 9, 52, 55, 57
minimizing sequence, 34 maximal ~, 59, 60
Mitra, 192, 327 Osgood, 191, 326
Mittag-Leffler, 152, 325
Moore, 191, 325 Pantelidis, 233, 326
Moreau, 69, 87, 326 parallelogram law, 8, 103
Morris, 298, 320 Pardalos, 284, 325
Motzkin, 307, 326 Park, 234, 235, 316
Mulansky, 284, 326 Parseval identity, 59
Muntz, 153, 326 extended ~, 59
Muntz's theorem, 147, 153 partial order, 311
Murray, 192, 326 partially ordered set, 311
336 INDEX

Paszkowski, 298, 327 Runge, 152, 328


Penrose, 192, 327
perturbation theorem, 267 Saeki, 153, 318
Phelps, 32, 307, 327 Schauder, 191, 328
Phillips, 123, 320 fixed-point theorem, 304
Picard, 152, 327 Schmidt, 19,51,328
Pinkus, 153, 327 Schwartz, 41, 123, 191, 320
Pollul, 284, 323 Schwarz, 19, 328
Polyak, 235, 322 inequality, 3, 19
polyhedral set (polyhedron), 109 self-adjoint, 173, 176, 193, 217
Chebyshev -, 111 separable space, 69, 142
dual cone to -, 115 separate sets by hyperplane, 102
polynomial approximation, 1 separation theorem, 103, 123, 251
to data, 1, 27, 61 for subspaces and cones, 105
to function, 2, 28, 63 sequence
polynomial interpolation, 290 Cauchy -,12
polynomials, 24 convergent -, 10
of Bernstein, 151 minimizing -,34
of Chebyshev, 53 shape-preserving interpolation, 229,
of Legendre, 53 237, 283
on discrete sets, 54 Siegel, 192, 328
trigonometric -, 297 Simonic, 235
positive constrained interpolation, 2, 30 simultaneous approximation and
Powell,233,327 interpolation (SAl), 292
pre-Hilbert space, 3 simultaneous approximation and
product space, 169, 190 norm-preservation (SAN), 294
projection, 86 simultaneous approximation, interpola-
theorem, 77, 87 tion, and norm-preservation (SAIN),
Prolla, 153, 327 294,298
proximinal set, 21, 25, 32-33, 37 simultaneous interpolation and norm-
Pythagorean theorem, 9 preservation (SIN), 294, 298
Singer, 32, 298, 308, 328
Rado,192,327 Smarzewski, 234, 328
Raik, 235, 322 Smith, 32, 87, 235, 284, 285, 315, 323,
range of operator, 160 325, 328, 329
Ransford, 153, 327 Sobczyk, 123, 317
Rao, 192, 327 Solmon, 235, 328
reduction principle, 80 Soukhomlinoff, 123, 329
regression Stechkin, 32, 41, 307, 308, 309, 320
convex, 228 Steffensen, 298, 329
isotone, 85, 228 Steinhardt, 285, 321
Reid, 191, 285, 327 Steinhaus, 191, 316
Rellich, 19, 70, 327 Steinitz, 32, 69, 329
representer, 91, 95, 97, 111, 295 Stiles, 233, 234, 329
Riesz, 41, 123, 191, 233, 327, 328 Stone, 153, 329
Rivlin, 32, 328 strong CHIP, 239, 240, 241, 242, 243,
Rockafellar, 32, 283, 284, 328 244, 245, 246, 257, 262
INDEX 337

strong conical hull intersection property open mapping -, 166


(see strong CHIP) parallelogram law, 8, 103
strong separation theorem, 103, 123 perturbation -, 267
for subspaces and cones, 105 Pythagorean -, 9
strong uniqueness, 64 Schauder fixed-point -, 304
sub differential, 279 Schwarz's inequality, 3, 19
subspace, 24, 64 separation -, 103, 123
maximal - , 97 Smarzewski -,234
spanned by, 24 Smith-Solmon-Wagner -,220,235
Sullivan, 233 Stiles -, 233
sum Stone-Weierstrass -, 153
of sets, 255 strong separation -, 103, 123
of subspaces, 199, 222, 231 uniform boundedness -, 165, 191
sun, 302 uniqueness of best approximations, 23
characterization of, 302 Vlasov's -, 304, 307, 308
support function, 278 weak compactness - , 205, 234
support of a function, 15, 270 Weierstrass approximation-,
symmetric matrix, 174 140, 152
Swetits, 284, 285, 325, 325, 329 Wolibner's -, 298
Szabados, 298,329 Yamabe's -,292,298
Szasz, 153, 329 three-term recurrence, 54
totally ordered set, 311
Teboulle, 285, 321 Tsar'kov, 309, 329
theorem Tseng, 192, 329
Aronszajn -,235
Bauschke-Borwein-Lewis -,235
Berens-Westphal - , 308 Ubhaya, 284, 285, 320
Bolzano-Weierstrass - , 41 uniform boundedness theorem, 165
bounded inverse - , 169 uniqueness of best approximations,
Boyle-Dykstra -, 213, 234 (Chapter 3), 23, 35
Bunt's -, 306 upper bound, 311
closed graph -, 170 Utreras, 284, 285,325
closed range -, 168, 191
Deutsch -,234
Deutsch-Morris -, 294, 299 Valentine, 307, 330
Efimov-Stechkin - , 308 de la Vallee-Poussin, 152, 316, 330
F'rechet-lliesz representation -, Vandermonde determinant, 297
94, 123 Vertesi, 298, 329
Gram-Schmidt orthogonalization -, Vlasov, 307, 308, 315, 330
52 Votruba, 192, 326
Hadamard determinant inequality, 65
Hahn-Banach extension -, 94, 123 Wagner, 235, 328
Heine-Borel -, 41 Walsh, 298, 330
interpolation -, 111 Ward, 69, 283, 284, 285, 318, 319, 320,
Klee's -, 307 325, 329, 330
Muntz's - , 147, 153 weak compactness theorem, 205, 234
von Ntumann -, 195, 233 weak convergence, 203, 206
338 INDEX

not equal to strong, 204 Wolibner, 298, 330


versus strong, 204 Wolkowicz, 284, 317, 329
weakly closed, 207
Weierstrass, 41, 152, 153, 330 Xu, 284, 285,320, 329, 330
approximation theorem, 140, 152
weighted spaces, 17 Yamabe, 298, 330
Weinert, 234, 235, 323
Weinstein, 284, 329
Zarantonello, 87, 330
Westphal, 308,316
zero operator, 173
Weyl, 19, 330
Zhang, 70, 318
Whalen, 192, 319
Zorn's lemma, 311
Wiener, 233,330

Das könnte Ihnen auch gefallen