Author: Ponnusamy S.  

Tags: mathematics   mathematical analysis  

ISBN: 1-84265-079-3

Year: 2002

Text
                    ou nations
Function I
na Sl



S PONNUSAMY


\,/
Alpha Science





Foundations of FUNCTIONAL ANALYSIS
Otlaer Books of Interest Advanced Engineering Mathematics (1-84265-086-6) R.K. Jain and S.R.K. Iyengar Calculus for Scientists and Engineers (1-84265-048-3) KD. Joshi Complex Analysis (1-84265-030-0) v: Karunakaran A Course in Ordinary Differential Equations (1-84265-068-8) B. Rai et al An Elementary Course in Partial Differential Equations (81-7319-170-0) T. Amarnath A First Course in Mathematical Analysis (81-7319-064-X) D. Somasundaram and B. Choudhary Foundations of Complex Analysis (81-7319-040-2) S. Ponnusamy Function Spaces and Applications (1-84265-002-5) D.E. Edmunds Functional Analysis (1-84265-109-9) Chandrasekhara Rao Functional Analysis (81-7319-199-9) Pawan K. Jain An Introduction'to Measure and Integration (81-7319-120-4) Inder K. Rana Mathematical Analysis and Applications (81-7319-306-1) A.R Dwivedi Mathematical Applitions in Social and Industrial Sectors (81-7319-357-6) NC. Mahanti 
Foundations of FUNCTIONAL ANALYSIS s. Ponnusam.y a Alpha Science International Ltd. Pangbourne England 
s. Ponnusamy Associate Professor Department of Mathematics Indian Institute of Technology, Madras Chennai-600 036, India Copyright @ 2002 Alpha Science International Ltd. P.o. Box 4067, Pangbourne RG8 BUT, UK All rights reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the prior written permission of the publishers. ISBN 1-84265-079-3 Printed in India. 
TO BOOMA 
Preface This book is a first course on functional analysis and therefore not much of prerequisite is assumed here. In fact, only some knowledge of elementary linear algebra, real and complex analysis is essential. However, I have tried my best to include, in Part I, the necessary basic results and the ideas of the aforesaid topics. From some point of view, the reader will find that they are the essential beginning for the course on functional analysis. Part I of this text concerns with two fundamental chapters. Chapter 1 is preliminary i nature and deals with elements of basic concepts in calculus, lgebra, real and complex analysis. One of the interesting parts in the sub- ject of analysis is the concept of metric spaces and Chapter 2 concentrates more on this topic. Besides metric spaces, the most important parts in functional analysis are Banach spaces, Hilbert spaces and linear functionals on these spaces. Thus, the main goal of this book is to give a rigorous analysis of the ba- sics of functional analysis and therefore, the material found in Parts II and III forms the core of the book. Part II is devoted to the theory of normed space.s, Banach spaces, principle of contraction mappings and lin- ear operators in normed spaces, while Part III is focused more on inner product spaces, Hilbert spaces and the representation of linear functionals with some applications. Examples, remarks, observations and figures of this book are used to illustrate the important points, the concepts and the motivation at every suitable opportunity. Each chapter is provided with a fairly extensive set of useful exercises. In general, these exercises are not difficult and should be worked out to master the subject. Some of the problems are challenging, but they should not be beyond the range of the talented students. The numbering system followed in the text is self-evident and needs no elucidation here. Now, I have one last comment on the notation. For the sake of convenience, the sign _ signals the end of the proofs of Theorem, Corollary, Lemma and Proposition whereas the sign. indicates the end of Remark, Observation and Example. In summary, I have endeavoured to produce a text which is useful for the class room, as well as for self-study. In addition, I hope that through this book the reader will gain sufficient mathematical maturity to be able to pursue any advanced course in Functional Analysis with greater ease and understanding. I believe that the text can be comfortably used by the en- gineering students because of the inclusion of a large number of motivating examples and exercises. Certainly there will be much room for improve- 
. . . Vlll Preface ments, and I welcome comments and suggestions from anyone who reads the book. In writing this book, I received lot of encouragement from several of my teachers, friends and collaborators with whom I learnt lot of mathemat- ics. I am particularly grateful to Professors R.Balasubramanian (Chen- nai), O.P.Juneja (Kanpur), R.Ramachandran, R.K.S.Rathore (Kanpur), M.S.Rangachari (Chennai), F.Ronning (Norway), St.Ruscheweyh (Germany), C.S.Seshadri (Chennai), and V.Singh (Kanpur) who have been the source of inspiration. Also, I take this opportunity to thank Professor Matti Vuori- nen (Finland) for the hospitality during several of my visits to the Uni- versity of Helsinki, Finland, resulting in many improvements at various stages of the book manuscript. I also thank Prof. Hans-Olav J.Tylli and Prof. G.P.Youvaraj for providing many useful suggestions. My special thanks to Dr. Susma N.Agrawal, who spared her invaluable time in reading the entire manuscript and made many criticisms. It is my duty to thank Dr. Manju Rani Agrawal and Mrs. P.Vasundhra for their careful reading of the entire manuscript, Mr. V.Ashok Kumar and Ms. V.Sunitha for preparing the figures. I enjoyed the company of all my colleagues in the Mathematics Department at lIT Madras, especially,. Professors A.Avudainayagam, S.Sundar, P.Veeramani and V.Vetrivel for their encouragement and I thank all of them. I started this project while I was at the Indian Institute of Technol- ogy, Guwahati and I thank the institute for its support. I wish to express my thanks to the Centre for Continuing Education at the Indian Institute of Technology, Madras, India, for its support in the preparation of the manuscript. Finally, it has been a pleasure in working with the Narosa Publishing House and I am indebted to Shri N.K. Mehra for his confidence and continued enthusiasm during all stages of the writing process. I must also record my appreciation due to my daughter Abirami and son Ashwin for their understanding and love during the long period that I have taken to complete this book. Above all, my deepest gratitude goes to my wife Booma (alias Geetha), to whom the book is dedicated, for her infinite patience, continued support, and the loving encouragement in all walks of my life. S. Ponnusamy 
Symbol o aES aftS {a} {x:...} XuY XnY XCY X C Y or X S; Y XxY X \Y or X - Y XC => <==} or -/-t or -It  N No z+ z- z Q I R Index of Special Notation Meaning for empty set a is an element of the set S a is not an element of S set having a as unique element set of all elements with the property set of all elements in X or Y; Le., union of the sets X and Y set of all elements in X as well as in Y; Le., intersection of the sets X and Y set X is contained in the set Y; Le X is a subset of Y X C Y and X :F Y; Le., set X is a proper subset of Y Cartesian product of sets X and Y, {(x,y): x EX, Y E Y} set of all elements that live in X but not in Y complement of X implies (gives) if and only if, or briefly 'iff' converges (approaches) to; into does not converge does not imply set of all natural numbers, {I, 2, · . .} N U {OJ = {O, 1,2, · . .} set of positive integers, {1,2,...} set of all negative integers, {..., -2, -I} set of all integers (positive, negative and zero) set of all rational numbers, {p/q: p, q E Z, q :F O} set of all irrational number set of all real numbers 
x 1R c C IRQ IR- IRt IR+ IRoo Coo IRn en i]R 1F Z Izi Rez Imz argz Argz limsup IZnl lim inf IZnl lim IZnl supS f : D  Dl f(x) f(D) f-l (D) f-l(y) fog SUPxED f(x) Index of Special Notations IR U {-oo, oo}, extended real line set of all complex numbers, complex plane extended complex plane, C U {oo} set of all nonpositive real numbers, {x E IR : x < O} set of all negative real numbers, {x E ]R : x < O} set of all nonnegative real numbers, {x E ]R: x > O} set of all positive real numbers, {x E ]R: x > O} set of all infinite sequence of real numbers set of all infinite sequence of complex numbers n-dimensional real Euclidean space, the set of all n-tuples x = (Xl, X2, · · · , x n ), Xj E IR, j = 1,2, . . . , n n-dimensional complex Euclidean space, the set of all complex n-tuples z = (Zl, Z2, . . . , zn) set of all purely imaginary numbers, imaginary axis field C or IR z := x - i y, complex conjugate of z = x + iy V X2 + y2, modulus of z = x + iy, x, y E ]R real part x of z = x + iy imaginary part y of z = x + iy set of real values of () such that z = Izle iB argument () E arg z such that -1r < () < 1r upper limit of the real sequence {Iznl} lower limit of the real sequence {Iznl} limit of the real sequence {Iznl} least upper bound, or the supremum, of the set S C IR f is a function from D into Dl the value of the function at x set of all values f(x) with xED; Le., y E f(D) <==} 3 xED such that f(x) = y {x : f(x) ED}, the preimage of D w.r.t f the preimage of one element {y} composition mapping of f and g supremum of f in D 
Index of Special Notation inf S infzED f(x) maxS minS dist (A, B) d(A) [Xl, X2] (Xl, X2) (a; r) (a; r) o(a; r) r  ad e Z Logz logz o oz o {)z fz fz: 1'1 + 1'2 £(1') f(n) (a) 1l(D) f(x) = O(g(x)) } asx-+a Xl greatest lower bound, or the infimum, of the set S C ]R infimum of f in D the maximum of the set S C JR; the largest element in S the minimum of the set S C JR; the smallest element in S distance from A to B diameter of A {x = (1- t)X1 + tX2 : 0 < t < I} {x = (1 - t)X1 + tX2 : 0 < t < I} open disc {z E C: Iz - al < r} (a E C, r > 0) closed disc {ZEC: Iz-al < r}(aEC, r>O) the circle {z E C: Iz - al = r} d(O; r) d (0; 1), unit disc {z E C: I z I < I} unit circle {z E C: Izl = I} exp(z) = Eno  , exponential function In Izl + iArgz, -1r < Argz < 1r In Izi + i argz := Logz + 2k1ri, k E Z 1 ( 0 .0 ) . - - - - z = x + y 2 ox oy' 1 (  +i ) 2 ox oy of oz of {)z sum of the curves 1'1, 1'2 length of the curve l' n-th derivative of f evaluated at a family of analytic (holomorphic) functions on D. If(x)1 < Klg(x)1 for all x in a neighbourhood of a 
. . Xll f(x) = o(g(x)) } asx-+a lim X n = X, } n-+oo or X n -+ x, or d(xn, x) -+ 0 C(X) I" (n ) } (l < p<oo) 1 00 (n) I" (1 < p < 00) 1 00 c Co Coo A = (aij) A-l At r n (IF) Mmxn (IF) 6 ij det (A) trace (A) Cc[a, b] C[a, b] [ ']B : V -+ r Iv := I Bv := (}, 0 L(V, W) B(V, W) Ker (T) or NT 1m (T) or RT Sl. Index of Special Notations Hm f(x) = 0 z-+a g(x) sequence {xn} converges to x with a metric d set of all continuous scalar valued functions on X {Z = (Zl, Z2, . · · , zn) E r : L  1 I Z k I" < oo} {z = (Zl, Z2,..., zn) E r : maxl<k<n IZkl < oo} {z = {Zn}nl : L  llzkl" < oo} {z = {Zn}nl : SUPlk<oo IZkl < oo} {z = {Zn}nl E 1 00 : lim n -+ oo Zn exists and is finite} {z = {Zn}nl E c: lim n -+ oo Zn = O} {z = {Zn}nl E 1 00 : support [{Zn}nl] = {n : Zn  O} is finite} matrix A whose (i,j)-th entry is aij inverse of a non-singular matrix A transpose of a matrix A set of all polynomials of degree at most n with coefficients over the field IF set of all m x n matrices with entries in IF Kronecker delta function determinant of a matrix A = (aij)nxn L au, trace of A set of all continuous complex valued functions on [a, b] set of all continuous real valued functions on [a, b] coordinate map v I-t [V]B w.r.t the basis B identity map on the set V zero element on the vector space V set of all linear transformations T : V -+ W set of all bounded linear transformations T : V  W kernel of T or the nullspace of T image of T or range space of T orthogonal complement of S 
Index of Special Notation Xl,ll, d( ., .) d x (.,.) IIxll IIxlix II/lip (., .) distance function in a metric space distance function in a metric space X norm or length of a vector x norm of a vector x E X LP-norm of / inner product 
Contents Preface ........... Index of Special Notations. I BASIC ANALYSIS . . . . . . . . 1 Analysis and Linear Algebra. . 1.1 Review of Complex Numbers 1.2 Functions and Countability. .... 1.3 Review of Differentiability in Real Line 1.4 Concept of Derivative in Complex Plane. 1.5 Concept of Riemann and Riemann Stieltjes Integrals 1.6 Vector Spaces . . . . . . . . . . . . . . . . . . . . 1.7 Linear Transformations between Vector Spaces 1.8 Inequalities..... 1.9 Exercises.................. Concepts in Metric Spaces . . . . . . . . . . . . . 2.1 Metric Spaces: Definitions and Examples 2.2 Holder and Minkowski Inequalities . . . . 2.3 Metric spaces IP(n), IP and C[a, b] 2.4 Basic Topology . . . . . . . . . . . . . . 2.5 Continuity and Equivalent Metrics 2.6 Compactness............. 2.7 Cauchy Sequences and Completeness . . 2.8 Completion of Metric Spaces . . . . 2.9 Exercises II BANACH SPACES . . . . . . . 3 Normed Spaces . . . . . . . 3.1 Properties of Norm. 3.2 Convexity and Completeness 3.3 The Banach Spaces IP(n) (1 < p < 00) 3.4 The Sequence Spaces lP (1 < p < 00) . 3.5 The Function Space C(X) . . 3.6 Basic Results on LP-Spaces 3.7 Norms on C[a, b] 3.8 Exercises........... 2 iii v 1 3 3 10 15 19 22 27 35 42 53 59 59 67 75 82 94 103 120 .. 127 132 141 143 143 152 162 164 172 179 185 191 
XVl CONTENTS Contraction Mappings and Applications . . . 4.1 Discussion on Fixed Point Problems 4.2 Contraction Mapping Principle . . . 4.3 Applications to Differential Equations 4.4 Exercises............... Linear Operators on Normed Spaces . . . . . . 5.1 Finite Dimensional Normed Space . . . 5.2 Direct Sums and Complementary Subspaces . 5.3 Riesz Theorems. . . . . . . . . . . . . . . . 5.4 Approximation in Function Spaces 5.5 Schauder Basis . . . . . . . . . . . . . . 5.6 Bounded Linear Operators ........ 5.7 Inverse Operators. . . . . . . . . 5.8 Completion of Normed Spaces ...... 5.9 Quotient Spaces ...... . . . . . 5.10 Baire Category Theorem. . 5.11 Open Mapping Theorem. . . . . . . . . . . 5.12 Closed Graph Theorem ... ..... 5.13 Uniform Boundedness Principle. . . . . . 5.14 Extension of Continuous Functionals . . . 5.15 Embedding of Normed Spaces. . 5.16 Exercises .... . . . . . III HILBERT SPACES. . . . . . . . . 6 Inner Product Spaces. . . . . . 6.1 Inner Product. . . . . . . . 6.2 Examples of Hilbert Spaces . . . . . 6.3 Applications of Polarization Identity 6.4 Completion of Inner Product Spaces 6.5 Orthogonal Family of Vectors . . . . . . 6.6 Projections on Finite Dimensional Spaces 6.7 Orthogonal Projections on Hilbert Spaces 6.8 Orthonormal Basis and Bessel Inequality 6.9 Cardinality Theorems for Orthonormal Bases 6.10 Applications of Uniform Boundedness Principle. 6.11 Exercises............. Representation of Linear Functionals . . . . 7.1 Riesz Representation Theorem . ... 7.2 Adjoint Operators on Hilbert Spaces . . 7.3 Exercises............... 4 5 7 Bibliography Index ............. 197 197 200 212 217 221 222 231 234 247 253 255 274 277 285 290 294 299 305 307 321 332 339 343 343 353 361 365 367 377 382 399 413 416 422 429 429 436 448 449 451 
Part I BASIC ANALYSIS In this part we shall first review in Chapter 1 some basic facts from real and complex analysis and linear algebra. Apart from certain additional materials for the foundation of functional analysis, Chapter 1 also contains the mathematical background for most of the chapters and it is expected that they will really be useful for the understanding of the rest of the book. The material is grouped under several sections, each of which contains a set of theorems, lemmas, propositions, basic mathematical examples, remarks and certain interesting observations. Inequalities play an important role in several branches of science and technology, and therefore for an easy access to the reader, we provide a separate section on Inequalities particularly to represent several standard inequalities associated with sums and integrals of real/ complex valued func- tions. Some of these inequalities are important in other branches of math- ematics as well and, in fact, most of the students might have learnt the preliminary part of the material of this chapter in the usual analysis course. Chapter 2 reviews the concept of metric spaces, certain function spaces and develops point-set topology of metric spaces. These are done through concepts such as neighbourhoods, open sets, closed sets, sequences, etc. In Section 2.4, we introduce the concept of topology. These topological concepts are used to define the concept of continuity on metric spaces. In- deed, in Section 2.5, we investigate some topological properties of continuity which, in fact, can be applied in a more general setting. The concept of completeness plays a central role in the theory of metric spaces and is discussed in Sections 2.7 and 2.8. In particular, prior to the discussion on the completeness property, we analyze some basic examples of function spaces that appear frequently in Analysis. For example, the space CF[a, b] of all F(C or IR)-valued continuous functions on the closed interval [a, b] and the space BV[a, b] of all functions of bounded variation on the closed interval [a, b]. The sequence spaces lP(n) and IP, 1 < p < 00, and the properties such as inequalities between them are also discussed in Section 2.3. In particular, we shall see in the subsequent chapters that how the completeness property is important in the study of Banach spaces and Hilbert spaces. Completeness of some other spaces will be considered in Part II as well. 
Chapter 1 Analysis and Linear Algebra The purpose of this chapter is to give a brief review of the basic concepts from the theory of real and complex analysis which will be needed in what follows. In Section 1.2, we introduce the basic definitions such as one-to- oneness and ontoness. In Section 1.6, we deal with vector spaces, subspaces, linear transformations, and some of the basic properties of certain spaces such as IP(n) and IP, 1 < p < 00. Several basic inequalities will be proved in Section 1.8. These inequalities are required to show that certain spaces are metric spaces and they are also important in the other branches of analysis. 1.1 Review of Complex Numbers We briefly recall some notation and a few facts from the algebra of sets. If A is a set of objects (Le. numbers, vectors, or functions), and x is an element (or member) of A, we write x E A. Likewise, the expression x  A refers to "x is not a member of A". IT B is another set and each element of B is also an element of A, then we say that B is a subset of A, and we write B C A. Equivalently, B C A {:::::} x E B implies that x E A. For instance, if A and B are two sets then A = B {:::::} B C A and A C B. We use the symbol '0' to denote the empty set, the set with no elements. Clearly, each set is a subset of itself, and therefore to distinguish the subsets that do not coincide with the set in question, we say that A is a proper subset of B if A C B and in addition, B also contains at least one element that does not belong to A. We denote by the symbol A S; B, a proper subset A of B. However, if one wishes to indicate that A is a subset of B which is 
4 Chapter 1: Analysis and Linear Algebra possibly the set A itself, then we write A C B. When A is not a subset of B, then we can indicate this by the notation Acj.B meaning that there is at least one element x E A but not in B. 1.1. Relation. A relation < on a set F is a strict total order whenever a f:. a, a < b and b < c =} a < c, a < b or a = b or b > a for all a, b and c in F. We write a < b for a < b or a = b, and note that in a total order a < b {:} b f:. a. Familiar ordered fields are Q and III In an ordered field we define the absolute value lal of a as: lal = { a -a for a > 0 for a < O. 1.2. Concept of equivalence relation. Now, we consider a useful class of mappings called equivalence relation. Later we shall discus a variety of classes of mappings such as one-to-one, onto, and continuous. We use the notation "a f'J b" to indicate the relationship between two elements a, b of a set X rather than the ordered pair notation (a, b). Given a non empty subset X, an equivalence relation on X is a relation between the elements of X, denoted by the symbol f'J (or R), which satisfies the following three rules for all x, y, z EX: (El) x f'J x, Le. every element 01 X has a relation with itself [Reflexivity] (E2) x f'J y => y f'J x, Le. if x is related to y, then y is related to x [Symmetry] (E3) x f'J y, Y f'J Z => X f'J z, Le. if x is related to y and y is related to z, then x is related to z. [Transitivity] A relation f'J on a set X is called a partial ordering if it is reflexive, an- tisymmetric (Le. for x, y EX, X f'J Y and y f'J X imply that x = y), and transitive. A partially ordered set is a pair (X, ""), where X is a set and f'J is a partial ordering on that set. Clearly, the standard operation " < " on IR defines a partial ordering on III In consistent with this fact, it is a standard practice to denote the partial ordering by the more suggestive symbol " < " rather than the symbol "f'J". A partially ordered set (X, < ) is called linearly/totally ordered if < sat- isfies the condition x < y or y < x for every x, y EX. 
1.1. Review of Complex Numbers 5 In this case, < is called a linear ordering. If Y e x, then an element m E X is an upper bound for Y (with respect to the ordering < ) if Y < m for all Y E Y. An element m E X is called the maximal provided m < x, x E X implies m = x. 1.3. Examples. (1) Ordinary equality '=' is a trivial example of an equivalence relation on IR whereas each of ' < ' and ' > ' is not an equivalence relation. Note that, each of ' < ' and ' > ' is a partial ordering on III Moreover, the relation "<" does not satisfy the reflexivity condition since x < x is not true for any x E III It is also not symmetric, although it is transitive. (2) Assume that x = y iff x - y is even, where x, y E III Then ' = ' is an equivalence relation on the set III In particular, IR is not a partially ordered set with respect to this relation (eg. choose x = 3 and y = 5). (3) Let X = Z, the set of all integers. On X, define x f"oJ y iff y - x is divisible by 2. Then, it can be checked that Z is an equivalence relation with respect to the defined relation. (4) Let Y be the family of all subsets of a set X, and assume that A < B iff A C B, for A, BEY. Then Y is a partially ordered set with respect to the set inclusion C as our partial ordering on Y. In particular, for each A, BEY, it follows that A C Band B C A imply that A = B. If S C Y, then UAS A is an upper bound for S. We shall now state without proof the axiom of choice in one of its equivalent form, namely the Zorn's lemma: 1.4. Lemma. Let E be a nonempty partially ordered set. If every totally ordered subset of E has an upper bound, then E has a maximal element. Now, we start with the discussion on the set of complex numbers. The starting point for the introduction of the complex numbers, which all al- ready familiar to us from high schools, arises when we need to solve certain equations such as x 2 + 1 = o. Complex numbers may be introduced in the following way. A complex number is an ordered pair of real numbers: Z = (x, y). The word 'ordered' means that (x, y), (y, x) are distinct unless x = y. If Zl = (Xl,Yl) and Z2 = (X2,Y2) then we say that Zl = Z2 <==> Xl = X2 and Yl = Y2. 
6 Chapter 1: Analysis and Linear Algebra In particular, Z = (x, y) = (0,0) <==} x = 0 and y = o. In C, the set of all ordered pairs of real numbers, we define the arithmetic operations for Zl = (Xl, Yl) and Z2 = (X2, Y2) E C: Zl  Z2 - (Xl, Yl)  (X2, Y2) - (Xl X2,Yl Y2) ZlZ2 - (Xl, Yl)(X2, Y2) (XlX2 - YlY2, XlY2 + YlX2) AZ - (AX, AY), A E IR Zl ( X1 X 2 + 1}11}2 X2Y1 - X1Y2 ) if Z2  (0,0). - 2 2 ' 2 1A ' Z2 x 2 + Y2 x2 + 2 The notation commonly used for a complex number is not (x,y) but x+iy, x, y real. Following Euler, we define i in the complex number system C of ordered pairs: i := (0, 1). From here it follows that complex numbers (x, 0) may be identified with real numbers x. From the multiplication rule we can write (in an informal way) i 2 = i x i = (-1,0) = -1 + iO which makes it possible to express a complex number Z - (x, y) in the following useful algebraic form: Z = (x, y) = (x, 0) + (0, y) = (x, 0) + (0, l)(y, 0). Thus, the set C of complex numbers is defined to be the set of numbers of the form Z = x + iy, where i = A and x, yare real numbers. 1.5. Elementary properties of complex numbers. Complex num- bers of the form (x, 0) are said to be purely real or just real. Those of the form (0, y) are said to be purely imaginary. 'Zero' viz. (0,0) = 0 + iO is the only complex number at once real and purely imaginary. The complex conjugate of Z = x + iy is the complex number Z := x - iy. Note that Z = Z iff l x + iy = x - iy, i.e. Z is purely real. Geometrically, the complex conjugate of Z is obtained by reflecting Z in the real axis. For any two complex numbers Zl and Z2, the following simple properties of the complex conjugate are easy to verify by a straightforward calculation: · Z'1:EZ2 = Z l  Z2 · ZlZ2 = Z l Z2 1 The shorter form 'iff' is to be read as 'if and only if'. 
1.1. Review of Complex Numbers 7 · Zl = Zl . Zl/Z2 = Z l/ Z2 for Z2  O. We also define Izi := vx 2 + y2 which is called the modulus of the complex number z. We observe that Izl 2 = z z = x 2 + y2 = (Rez)2 + (Imz)2. With the standard notation, it follows that z+ z z- z Re z = x = and 1m z = Y = . 2 2i For any pair Zl, Z2 E C, the following properties are easy to verify: . Re (Zl :i: Z2) = Re Zl :i: Re Z2 . 1m (Zl :i: Z2) = 1m Zl :i: 1m Z2 · I Z IZ21 = I Z lllz21 · IZ1/ z21 = I Z ll/l z 21 for Z2 :F O. We know that the ordered pairs of real numbers represent points in the geometric plane with respect to a pair of rectangular axes. We then call the collection of ordered pairs as the Cartesian product }R x ]R = ]R2 and the two axes as x-axis, y-axis. Because (x,O) E }R2 corresponds to real numbers, x-axis is called the real axis and since iy = (0, y) E ]R2 is purely imaginary for y real, y-axis is called the imaginary axis. Now, we can conveniently visualize C as a plane with x + iy as points in }R2 and we simply refer to it as the complex plane. Depending on the problems on hand, we use x + iy, or (x, y), to represent a complex number. Thus, we see that a complex number Z = x+iy can be viewed geometrically as the point (x, y) in a coordinate plane (complex plane C): x+iy I-t (x, y). The distance from the origin to a complex number Z = x + iy is then V X 2 + y2 = Izi. For Z = x + iy, it is appropriate to include some useful elementary inequal- ities: . IRezl < Izl . IImzl < Izl · ()(Ixl + Iyl) < Izl < Ixl + Iyl . Iz:i: wi < Izi + Iwl, for z, w E C . Ilzl - Iwll < Iz - wi. 
8 Chapter 1: Analysis and Linear Algebra y y = r sin 8 z = re i9 = x + iy o x = r cos 8 x Figure 1.1: Polar representation of a complex number These inequalities are easy to verify. We define the argument of z, denoted by arg z, as the angle 8 made by the vector 0 z with the positive x-axis. Clearly z has an infinite number of distinct arguments. Any two distinct arguments of z differ from each other by an integral multiple of 21r. (Since z = 0 <==} Izl = 0, argz in this case is indeterminate). Thus, 8 is unique up to addition of a multiple of 21r radians. In order to specify a unique value of arg z, we may restrict its value to some interval of length 21r. To do this we introduce the concept of "principal value" of arg z as follows: For an arbitrary z :j:. 0, the particular argument of z lying in the range -1r < 8 < 1r is called the principal argument of z and is denoted by Arg z: 8 = Argz = arctan (y/x), -1r < 8 < 1r, where the last condition ensures that 8 is uniquely defined. Thus, the relation between arg z and Arg z is given by argz = Argz + 2k1r, k = 0, :i:1, :i:2, ... . If we let Izl = r then the complex number z = x + iy can be expressed in the so-called trigonometric form (or Euler form) z = r( cos 8 + i sin 8) =: re i9 and this representation is called the polar representation or modulus argu- ment form of z (see Figure 1.1). Then, we have x = rcos8 and y = r sin 8. By induction it is simple to prove De Moivre's formula: (cos 8 + i sin 8)n = (cos n8 + i sin n8), i.e e in9 = (e i9 )n, where n is a positive integer. Now, we note that the Euler exponent behaves as the exponential function. Returning to the starting equation x 2 + 1 = 0, we find that it has two complex roots Xl = i and X2 = -i. 
1.1. Review of Complex Numbers 9 1.6. Field. A field is a set F which possesses two binary operations namely, addition (+) and multiplication (.) such that F is closed with respect to these two operations (meaning that a, b E F implies a + b E F and a · b E F) and satisfies the familiar rules of rational arithmetic: . addition is associative, i.e. (a + b) + c = a + (b + c) for each a, b, c E F . addition is commutative, Le. a + b = b + a for each a, b E F . there exists an element 0 E F such that 0 + a = a for all a E F (0 is called additive identity) . to every a E F, there corresponds an additive inverse -a E F such that a + (-a = 0 for all a E F . multiplication is associative, Le. (a · b) · c = a. (b. c) for each a, b, c E F . multiplication is commutative, Le. a. b = b · a for each a, b E F. . there exists an element 1 E F, 1 :j:. 0, such that 1 . a = a for all a E F (1 is called multiplicative identity) . to every 0 :j:. a E F, there corresponds a multiplicative inverse a -1 E F such that a . a -1 = 1 for all a E F . multiplication distributes over addition, i.e. a. 0 = 0 and a · (b + c) = a · b + a · c The most familiar examples of fields are the set of rational numbers and the set of real numbers, for which the notation Q and IR are used, respecti vely. 1.7. The field C. First we show that C is a field. For this it suffices to prove the existence of the multiplicative inverse of each nonzero complex number Zl :j:. 0, as the remaining axioms are easy to verify. Thus, for Zl :j:. 0, we need to solve for Z2 the equation ZlZ2 = 1. By the multiplication rule, namely Zl Z 2 = (Xl, Y1)(X2, Y2) = (X1 X 2 - Y1Y2, X1Y2 + Y1 X 2), it follows that Zl Z 2 = (1,0) <==> { X1X2 - Y1Y2 = 1 X1Y2 + Y1 X 2 = 0 
10 Chapter 1: Analysis and Linear Algebra <==} ( Xl l ) (::) = () Yl <==} ( X2 ) 1 (Xl YI) () Y2 - x + X -Yl Xl <==} ( X2 ) (Xl YI) Y2 = X + X ' - X + X . Therefore, the multiplicative inverse (or simply the inverse or reciprocal) Z-l of a complex number z = X + iy :j:. 0 is then given by 1 x - iy ( x ) . ( y ) z = x 2 + y2 = x 2 + y2 -  x2 + y2 . In particular, in view of the fact that IR is field, the above discussion shows that C is a field too. Further, writing a real number x as (x, 0), as pointed out earlier, and noting that (Xl,O) + (X2,0) = (Xl + X2,0) and (Xl, 0)(X2, 0) = (XlX2, 0), IR turns out to be a subfield of C. If z = x + iy, then we use the notation x = Re z for the real part of z, and y = 1m z for the imaginary part of z. It is important to note that C is not an ordered field. 1.2 Functions and Countability Let X and Y be two nonempty subsets of C or III A function/mappinrr I from X to Y is a rule, which associates with each x E X a unique element y E Y. We write (1.8) I : X -+ Y to denote the mapping of X into Y. We call the sets X and Y the domain and codomain of the function I, respectively. When we describe a mapping by describing its effect on the individual elements, we use the symbolt-t; thus "the mapping x t-t I(x) of X into Y" means that I is a mapping of X into Y taking each element x of X into the element I(x) of Y. IT y = I(x), we say that y is the image of x. If S e x, we can have I : S  Y and we call this new function the restriction of I in (1.8) to S and denote it by lis' If I is defined on X and S e x, then 1(8) = {/(s) : s E S} 2The terms mapping, function and transformation are used synonymously. 
1.2. Functions and Countability 11 is called the image of the set S under I. In particular, I(X) C Y is called the range of I. In other words, the subset {/(x) E Y : x E X} C Y is called the range of I. If Y 1 C Y, then the inverse image of Y 1 under I, denoted by 1- 1 (y 1 ), is the subset of X defined by 1- 1 (y 1 ) = {x EX: I(x) E Y 1 }. If Y 1 = {y} C Y, then we write 1-1(y) instead of 1-1({y}). Note that 1- 1 (Y 1 ) is a well defined set irrespective of whether I has any inverse or not. The following result is trivial. 1.9. Proposition. Let X and Y be any two sets, I : X  Y be a given function, A C X and B C Y. Then I(A) S; B iff A S; 1- 1 (B). There are several similar properties of images and inverse images (preim- ages) but these will be discussed in Chapter 2 in a more general setting. However we shall recall some more elementary background material as we proceed. For two mappings I : A  Band g : B  C, we can define the composite mapping g 0 I : A  C by (g 0 I)(x) = g(/(x)). The mapping I : A  B is said to map A onto B iff the codomain and the range set are equal, Le. I(A) = B. Therefore, to prove that I is onto, one must start with an arbitrary b E B and then find at least one a E A such that I(a) = b. The mapping is said to be 1-1 (one-to-one) iff it maps distinct elements into distinct elements, Le. I (a1) :j:. I (a2) for all a1, a2 E A with a1 :j:. a2. More formally, I is one-to-one iff for a1, a2 E A, l(a1) = l(a2) => a1 = a2. A mapping which is both one-to-one and onto is called bijective. 3 The map I : A  B is said to have an inverse if there exists a function 9 : B  A such that g(/(a)) = a for all a E A and I(g(b)) = b for all b E B. Here 9 is called the inverse of I. We have a simple and useful result which we state without proof. 1.10. Proposition. Let A and B be two sets and I : A  B. Then I has a unique inverse 9 iff I is bijective. 3The term 'one-to-one', 'onto', and 'one-to-one correspondence' are sometimes re- ferred as 'injective', 'surjective', and 'bijective' mappings, respectively. 
12 Chapter 1: Analysis and Linear Algebra It is important to observe that the inverse image of any subset of B exists even if f : A  B is neither one-to-one nor onto. For example, let f : Z -+ Z be given by f(n) = Inl, where Z denotes the set of all integers. Then, f is neither one-to-one (because f( -n) = f(n) for each n) nor onto (because, there exists no n E Z with f (n) = -1) so that f is not bijective. By the last proposition, it follows that f has no inverse. On the other hand, inverse images certainly exist (eg., f-1 ({I, 2}) = { -1, -2,2, I}). The following proposition is useful. 1.11. Proposition. Let X and Y be any two sets, f : X  Y be a given function, A C X and B C Y. Then (i) A C f-1(f(A)), with equality if f is one-to-one. (ii) f(f-1(B)) C B, with equality if f is onto. Proof. (i): If a E A, then f(a) E f(A) so that a belongs to the set {x : f(x) E f(A)} and this implies that a E f-1(f(A)). Thus, A C f-1(f(A)). Next we prove the reverse inclusion under the assumption that f is one-to-one. For this, we let x E f-1(f(A)). Then f(x) E f(A) so that f(x) = f(a) for some a E A. But, since f is one-to-one, then x = a and therefore x E A, as desired. (ii): Let y E f(f-1(B)). Then y = f(x) for some x E f-1(B) which means that f(x) E B. Thus, y E B; Le. f(f-1(B)) C B. To prove the reverse inclusion under the assumption that f is onto, let us take an arbitrary element b E B. As B C Y, by ontoness of f, we have b = f(x) for some x E X. But then f(x) = b E B which gives x E f-1(B), Le. f(x) E f(f-1(B)), and hence b E f(f-1(B)). . 1.12. Remark. Consider the function f : IR  IR, x I-t x 2 . If A = [0,1] and B = [-1,1], then we have f(A) = A, f-1(f(A)) = [-1,1] ct A and f-1(B) = B, B ct f(f-1(B)) = A. These observations verify the validity of the strict inclusion in Proposition 1.11. . 1.13. Examples. (1) Consider the mapping f : A  B, a I-t a 2 , where A and B are subsets of IF. Then f(a1) = f(a2) ==> (a1 +a2)(a1 -'a2) = 0 ==> a1 = a2 if a1 + a2 :F o. Therefore, we have 
1.2. Functions and Countability 13 (i) Let A = R and B = IRt , the set of all nonnegative real numbers. Since there exist a1, a2 E A such that a1 + a2 = 0, f is not one-to-one in this case. Similarly, if A = B = Z, then f is not one-to-one because of similar reasoning. (ii) Let A = B = B+, the set of all positive real numbers. Then, for each a1, a2 E A, we have a1 + a2 :j:. 0 and therefore, f is one-to-one in this case. Similarly, we see that if A = B = N,. the set of natural numbers, then f is one-to-one. (iii) If A = B = JR, then f is not onto because the set of all real numbers is not the image of IR under our mapping. Also, if A = B = N then f is not onto. However, if A = IR and B = IRt then f is onto. In fact, when A = B = R+ , f is bijective. (iv) If A = B = C, then f is not one-to-one but onto. (2) The mapping f : Z  N U {OJ = No by x I-t Ixl is onto. (3) The mapping f : Z  Z described by x I-t x + 1 is onto whereas f : N  N defined by x I-t x + 1 is not onto because there is no element a E N with the property that f (a) = a + 1 = 1. On the other hand f : No  N, x I-t x + 1, is onto. (4) The mapping f : Z  Z by x I-t 2x is not onto. For, let b E Z. Then we have to solve the equation f(a) = b = 2a, Le. a = b/2. But b/2 is not necessarily an integer when b E Z. However, if B denotes the set of even integers then f : Z  B by x I-t 2x is onto. (5) The mapping f : IR  [-1,1], x I-t sin x, is onto whereas f : R  JR, x I-t sin x, is not onto. . A sequence Z1, Z2, . .. of points in IF (where 1F denotes either the field C of complex numbers or the field IR of real numbers) is really a mapping f : N  IF. If f is a sequence, we write, in keeping with the tradition, Zn instead of f(n), so that the points f(n) = Zn are called the (n-th) terms of the sequence. The other common notation to denote the sequence is by either {zn} or {Zn}n1' or {Zn}  l' for the sequence f, where Zn = f(n) E IF. It is purely a notational reason to let the sequence to begin with Z1. For Z E IF, the sequence given by Zn = Z for all n E N is called the constant sequence with value z. Let 9 : N  N be an increasing sequence of natural numbers. Then the sequence {Zg(n)} is called a subsequence of the sequence {zn}. We often write g(k) = nk so that 1 < n1 < n2 < ..., and thus, the sequence {Wk} defined by Wk = Zn", 
14 Chapter 1: Analysis and Linear Algebra is the subsequence of {zn}. Intuitively, a subsequence is obtained by 'throw- ing away' some terms of the original sequence. subsequences 'are used ex- tensively in analysis. For example, the concepts such as compactness can be handled nicely with the help of subsequences. Often we talk of an indexed family of objects: Notation like {Za}aEA is commonly used for indexed families, where A is the indexing set. The point here is that for each a E A, there is an object Zao For instance, the infinite sequence {Zn}  1 is a family indexed by N, the set of natural numbers. Further, the concept of the sequences of points defined on an arbitrary set is similar. A set S is said to be countable if it is either finite or there exists a one-to-one correspondence between N and the set S. A set is said to be uncountable if it is not countable. For example, N, Z and Q are all countable sets. On the other hand, we know from real analysis that IR is uncountable and hence, the set of all rational numbers is uncountable. In particular, any interval which contains more than one point is uncountable. Indeed, the fact that there are uncountably many real numbers in (0, 1) follows from constructing, for example, a set of all infinite sequences of O's and l's which is uncountable. Therefore, we have a natural question to look at: How about other familiar sets that are uncountable? We note that, according to the definition, the counting convention is via bijections, and the set of real numbers actually have a lot more numbers than the set of rational numbers. In general, given a set X, does there exist a method of constructing another set from X that will contain more elements than X? If X is countable (finite or infinite), then the answer is trivial, because if X is finite then one can simply obtain a new set just by adding one more element that does not belong to X. If X is count ably infinite, then the new set obtained by adding a finite number of elements or even countably infinite number of elements to X would again be countable. Hence, we have to think of some other method. Indeed, a method of getting a bigger and bigger set follows from the definition of power set: "The power set of a given set X is the set of all subsets of X, denoted by P(X)". Thus, the notion of cardinality of a set X comes in to play a role. If a set X is finite, then the number of elements of X is defined to be the cardinality of X, denoted by IX I or card X. Thus, two finite sets A and B have same size, Le. card A = card B, if they contain the same number of elements. An important question is how to carry the notion of equal size over to infinite sets such as Nand Z? Now, we have the definition. Given two arbitrary sets A and B (finite or infinite), cardA = cardB iff there exists a bijection between them. In particular, the notion of equal size is an equivalence relation, and we then associate a number called cardinal number to every class of equal sized sets. At this point, it is important to note that it is often difficult to find the cardinal number as it requires th3:t the function is both one-to-one and onto. But it is usually easy to find one-to-one functions than onto functions. Now, we state without proof the 
1.3. Review of Differentiability in Real Line 15 following theorem due to Cantor-Bernstein. 1.14. Theorem. (Cantor-Bernstein) Let A and B be two sets. If there exists a one-to-one function f : A  B and another one-to-one function 9 : B  A, then card A = card B. This theorem can be used to show, for example, that card (IR x IR) = card III Moreover, the fact that Q is countable can also be obtained by showing that Q and Z x Z have the same cardinality (prove this!). 1.3 Review of Differentiability in Real Line In this section, we show how the idea of the derivative of a function at a point in IR, as a linear approximation, is fundamental to the understanding of the derivative in the real case. We are assuming the concepts of limit, continuity, and uniform continuity for functions defined on subsets of III One of the important concepts in real and complex analysis is the notion of neighbourhoods. Given a point a E IR and a positive number €, the open interval (a-€,a+€)={xEIR: a-€<x<a+€} is called a neighbourhood of the point a. Likewise, if.a = al + ia2 is a given complex number and r is a positive number, then the open disc (a;r) := {ZEC: Iz-al <r} : = { (x, y) E ]R2 : (x - a) 2 + (y - b) 2 < r 2 } is called a neighbourhood of a E C. If G C IR, then G is called open iff for every a E G there exists an € > 0 such that (a - €, a + €) C G. Similarly, D C C is said to be open iff for every a ED, there exists a r > 0 such that (a; r) C D. We shall also name the complements of open sets. They will be called closed sets but we shall discuss them in detail later. Let G be an open subset of IR containing a point a. We say that a function f : G  IR is differentiable at a if the limit (1.15) I . f(x) - f(a) 1m x-+a X - a exists and is finite. Then we say that f has a derivative at a and this limit is denoted by f'(a). IT f has a finite derivative for every point of G, then f is differentiable on G and, in that case, f'(x) is a function of x and the following notation may be used instead of f'(x): D f(x), :x f(x),  , y', f(1) (x). 
16 Chapter 1: Analysis and Linear Algebra y x  </J(x) = f(a) + f'(a)(x - a) o a x Figure 1.2: Description of derivative concept in real line The idea of the derivative of f at a given point is connected with the notion of a tangent to the graph of y = f(x) at this point. To clarify this geometric significance n terms of the quotient (f(x) - f(a))/(x - a), we may rewrite (1.15) as (1.16) f(x) = f(a) + f'(a)(x - a) + 7J(x)(x - a), where lim x -+ a 1J(x) = O. This equation is, in fact, the basis for the differen- tiability theory of functions of several variables. If we introduce a new map L : IR  IR defined by Lx = f'(a)x, then (1.16) becomes (1.1 7) Urn If(x) - f(a) - L(x - a)1 . 0, x-+a Ix - al where the expression involving absolute sign are convenient when proving results about differentiable functions. Equation (1.17) may be interpreted as saying that f(a) + L(x - a) is a good approximation to f(x) at a. By the definition of limit, (1.17) is equivalent to say that for each € > 0 there is a 6 > 0 such that ( If (x) - f(a) - f'(a)(x - a) I < €Ix - al whenever Ix - al < 6. Now, the usual geometric interpretation of the derivative at a point may be stated as follows: Let </J(x) = f(a) + f'(a)(x - a) =: f(a) + L(x - a). By the definition of </J(x), the graph of y = </J(x) is the set of all pairs (x, y), x E G, which is the equation of the line in the plane passing through the point (a, f'(a)) with slope f'(a), see Figure 1.2. If a function is differentiable, it may not have any corners. This observation makes it easy to decide whether a function is differentiable if we know the graph of the function. There is another way of defining the derivative at a point which says that, if f : G  IR is differentiable at a, then there exists a function </J whose graph 
1.3. Review of Differentiability in Real Line 17 is a straight line and this function provides the best linear approximation to f at a. In fact, (1.15) means that f'(a) can be approximated by f(a + h) - f(a) h for sufficiently small h, which is same as to say that y = f(a + h) - f(a) can be approximated by f'(a)h. If we write h = dx, we find that y  f'(a)dx where the symbol  denotes the "approximately equal to". From this, one gets the differential dy at a point a by dy = f'(a)dx, d or dx f(x) = f'(a). x=a Are f1 (x) = x 2 , f2(X) = -IX and f3(X) = Ixl differentiable at O? One can draw the graphs of these functions to analyze the geometric significance of the quotient (fj(x) - fj(a))/(x - a) for j = 1,2,3. If f is defined by { X for x > 0 f(x) = _2X2 for x < 0 then we have f'(x) = 21xl which is not differentiable at O. By writing f(x) - f(a) = ( f(X) - f(a) ) (x - a), x-a it is seen immediately that every function which is differentiable at a point is continuous thereat. The converse is not true. For example, the function f(x) = Ixl is continuous on IR but is not differentiable at the origin. We summarize the discussion of this section and reformulate the defini- tion of the differentiability as follows: 1.18. Theorem. Let G be an open subset of IR and a E G. We say that a function f : G  ]R is differentiable at a iff there exists a linear map L : IR -+ ]R such that f(x) = f(a) + L(x - a) + R(x), where the remainder function R(x) satisfies the condition lim IR(x)1 = O. x-+a Ix - al It is this theorem which provides a method of generalizing the concept of derivative of f defined from a subset of n-dimensional Euclidean space 
18 Chapter 1: Analysis and Linear Algebra an into IRm in which case the corresponding L in the last theorem becomes an m x n-matrix, which we denote by D f(a) and is called the derivative of f at a. The generalization of this idea is studied in advanced calculus as well as in advanced analysis courses. Since we require the concept of norm on n-dimensional Euclidean space IRn to give the precise definition of the differentiability in the higher dimension, we do not wish to include this definition at this stage. However, the following result is important to remember and we shall make a general statement later in Chapter 2 (see Corollary 2.83). 1.19. Proposition. Every continuous function on the closed interval [a, b] is uniformly continuous therein. 1.20. Continuously differentiable functions on IR. Recall that if f is differentiable at every point on I = (a, b) c IR, then f'(x) not only exists as a function on I but also that f is continuous on I. Hence, it makes sense to ask two natural questions that are considered to be important: Is f' continuous on I? Is f' differentiable on I? and so on. In general, the answer is no for both the questions as we see from the examples below. A function f : I = ( a, b)  IR is said to be continuously differentiable on I, or of class C 1 on I, if f' exists on I and the function f' : I  IR is continuous. The class of all one time continuously differentiable on I is denoted by C 1 (I). If f' is differentiable on I, then f is twice differentiable on I. If the second derivative f" is continuous, then f is of class C 2 on I. More generally, one can define k-times continuously differentiable func- tions on I for any positive integer k, and we denote the class of all such functions by Ck(I), see Exercise 5.162. Thus, COO (I) denotes the class of all infinitely differentiable functions on the open interval I. 1.21. Example. The function f defined by f(x) = { 0 x n for x < 0 for x > 0 belongs to cn-1(IR), but does not belong to Cn(IR). See Figure 1.3 for the case n = 2, where g(x) = { :2 for x E [-1,0) for x E [0,1]. It is easy to see that 9 and g' are both continuous, but g"(x) is discontinuous at the origin and hence, 9  C 2 [-1, 1]. Indeed, , (0) 1 . x2 - 0 0 9 =lm 0 =, x-+o x- and g'(x) = { 0 2x for x E [-1, 0) for x E [0, 1] 
1.4. Concept of Derivative in Complex Plane 19 y (0,1)--------- (1,1) y = f(x) I I I I I I I (-1,0) o (1,0) x Figure 1.3: 9 E C 1 [-I, 1] but not in C 2 [-I, 1] whereas I . g' (x) - 0 _ I " 2x - 2 1m -lm--, x-+o+ X - 0 x-+O+ X which shows that g' is not differentiable at the origin. Another simple example is given by { xn1x 1 f(x) = o for x :j:. 0 for x = O. It can be easily seen that this function belongs to en (IR), but does not belong to e n + 1 (IR) since fen) is not continuous at the origin. Another simple example of function f such that fen) exists at the origin but the function fen) is not continuous at the origin may be given by f(x) = x2n sin (l/x) for x :j:. 0 and f(O) = o. . 1.4 Concept of Derivative in Complex Plane The definitions of limit, continuity, differentiability and uniform continuity are some what analogous to those in Real Analysis. In this section, we briefly discuss about the concept of differentiability in the complex plane. Suppose that a complex-valued function f is defined on D C C and Zo is either in D or on the boundary of D. We say that the function f has a limit l as z  Zo and write lim f(z) = l or f(z)  l as z  Zo Z-+ZO iff for any given € > 0, there exists a 6 = 6(€, zo) > 0 such that If(z) -ll < € whenever zED and 0 < Iz - zol < 6. i.e. iff for each € > 0 there exists a 6 > 0 such that f(z) E (l; €) whenever z E [(zo; 6) \{zo}] n D. 
20 Chapter 1: Analysis and Linear Algebra First, it should be noted that the function need not be defined at Zo in order to have a limit at ZOo Secondly, it is the punctured disc (zo; 6) \{zo} which is involved in D, Le. Zo need not be in D. Thirdly, even if the condition that Zo E D holds, we may have j(zo) :j:. t. In real variable theory we do not have the freedom which a complex variable has, for, if Zo = Xo E ]R, a neighbouring point z = x  Xo has only two possible ways either on left or right. In the complex case z can approach Zo in any manner in the complex plane. As in Real Analysis, if a limit exists then it is easy to see that it is unique. A function j : D  C is continuous at Zo E D iff limz-+zo j(z) exists and equals the function value j(zo). We say that j is continuous on D or j : D  C is continuous when j is continuous at all points of D. Note that j is continuous at Zo iff the following three conditions hold: . j(zo) is defined . lim j(z) exists z -+ Zo . lim j(z) should be equal to j(zo). z -+ Zo In terms of our earlier notation, the definition of continuity is that for a given € > 0, there exists a 6 > 0 such that Ij(z) - j(zo)1 < € whenever zED and Iz - zol < 6, or equivalently, j(z) E (j(zo); €) whenever z E (zo; 6) n D. To some extent the rules for differentiation of a function of complex variable are similar to those of differentiation of a function of real variable. Since C is merely }R2 with the additional structure of addition and mul- tiplication of complex numbers, we can immediately transfer most of the concepts of}R2 into those for the complex field C. In fact, we have already done so when we analyzed the concept of distance (modulus). We say that a complex function j defined on an open set D is differen- tiable at an interior point Zo of D if the limit (1.22) I . j(z) - j(zo) 1m z-+zo Z - Zo exists. The value of the limit, denoted by j'(zo), is called the derivative of j at Z00 The quantity j'(zo) is generally a complex number. Sometimes it is advantageous to write z = Zo + h, h, a complex number, so that (1.22) is equivalently written as j ' ( ) _ I . j(zo + h) - j(zo) ZO-lffi h . h-+O 
1.4. Concept of Derivative in Complex Plane 21 Again we note that the limit exists irrespective of the path along which h  O. In terms of '€ - 6' notation, the limit in (1.22) exists iff given any € > 0, there exists a 6 = 6 ( €, zo) > 0 such that j(z) - j(zo) _ j'(zo) < € whenever 0 < Iz - zol < 8. Z - Zo For a given Zo, one can also consider the function 1J : D  C defined by { I(z)-/(zo) - f'(zo) 1J(z) = z-zo o for z :j:. Zo for z = Zo. Then (1.22) is equivalent to limz-+zo 1J(z) = 0 so that 1J is continuous at Zo. This observation shows that f (z) is differentiable at Zo iff there exists a number f'(zo) and a function 1J, continuous at Zo, satisfying the condition 1J(zo) = 0 such that (1.23) f(z) = f(zo) + (z - zo)f'(zo) + (z - zO)1J(z). Note that, as in the real case, the explicit expression in the form (1.23) has the advantage of containing no limit since this is replaced by the continuity of 1J(z). The function f is said to be differentiable on (in) the open set D if it is differentiable at every point of D. A function f is said to be analytic at a point a ED, where D is some open set in C, if there exists an open disc (a; 6) in D such that f is differentiable at all points of (a; 6). Thus, f is analytic in an open set D is equivalent to say that f is differentiable on D. A function f : [a, b]  C is said to be continuously differentiable on [a, b], or a map of class C1 (denoted by C[a, b]) if the function f(t) = u(t) +iV(t)1 t E [a, b], is continuously differentiable on [a, b], Le. u' (t) and v' (t) exist. for each t in [a, b] and are continuous functions of t on [a, b]. Note that f(t) is differentiable on [a, b] means that f'(t) exists on (a, b), and I . f(a + h) - f(a) lIll , h-+O+ h I . f(b + h) - f(b) 1m h-+O- h both exist. We denote these limits by f'(a+) and f'(b-:-), respectively. The space of all scalar-valued and continuously differentiable functions f on [a, b]" is usually denoted by Ci[a, b]. If the map f is from [a, b] into IR (instead of IIlapping into C) then it is denoted by C 1 [a, b] := C[a, b]. 
22 Chapter 1: Analysis and Linear Algebra 1.5 Concept of Riemann and Riemann Stieltjes Integrals One of the useful concepts is integration and is introduced in a calculus course especially for 'finding the area under a curve'. Is it possible to think of 'integration' in the form of summation? The summation interpretation of integration will make many of its properties easier to remember and also to get more information without much hardship. First, we recall how to construct the Riemann-Stieltjes integrals (in particular, the Riemann integrals). We will briefly discuss the Lebesgue integrals later, in Section 3.6. Let a < b. A partition P of the closed interval [a, b] is a finite set of points {XO, Xl, . . . , xn} satisfying a = Xo < Xl < X2 < ... < Xn-l < X n = b. We denote the set of all partitions of [a, b] by II[a, b]. If P = {Xk}k=O E II[a, b], then Xk = Xk - Xk-l defines the length of the k-th subinterval [Xk-l, Xk]. In this case, we define the norm or mesh of the partition by IPI=max{Xk: k=l,...,n}. If PI and P 2 are any two partitions of [a, b], then we say that the partition P 2 is a refinement of the partition PI, written PI C P 2 , if P 2 contains all the points from PI and some additional points, again sorted in order of magnitude as defined for any partition P. Since each of the subintervals formed by the partition P 2 is contained in a subinterval which arises from the partition PI, it follows that IP 2 1 < IPll whenever PI C P 2 . Let f: [a,b]  IR be a bounded function. Let P = {XO,Xl,...,Xn} be a given partition of [a, b]. For each k, 1 < k < n, we let M k = sup{f(x) : X E [Xk-l,Xk]} and mk = inf{f(x) : X E [Xk-l,Xk]}. For a nondecreasing function a on [a, b], define ak := a(xk) - a(xk-l). We define the 'Upper and lower Riemann-Stieltjes sums for f, defined on [a, b], with respect to a by n n U(P,f,a) = LMkak and L(P,f,a) = Lmkak' k=l k=l 
1.5. Concept of Riemann and Riemann Stieltjes Integrals 23 y . I i o Xo Xl X2 X3 X4 X5 X6.. .Xn-lX n X Figure 1.4: Lower Riemann sum y .. . '..",' :.: ." ..... fj . '">, ": ".. I Xo Xl X2 X3 X4 X5 X6.. .Xn-IX n X o Figure 1.5: Upper Riemann sum respectively. Since mk < Mk and ak is nonnegative, the lower sum is always less than or equal to the upper sum: L{P,f,a) < U{P,f,a) for each partition P, see Figures 1.4 and 1.5. The following lemma shows that the upper sum is decreasing with respect to a refinement of the partition, while the lower sum is increasing with respect to a refinement of the partition: 1.24. Lemma. H P l < P 2 , then we have L{Pl,f,a) < L{P 2 ,f,a) < U{P 2 ,f,a) < U{Pl,f,a). Proof. Let P l = {a = Xo < Xl < X2 < ... < Xi-l < Xi < ... < Xn-l < X n = b} be a partition of [a, b]. Obtain P 2 from P l by adjoining a point t in between the subinterval [Xi-l, Xi] for some fixed i: P 2 = {a = Xo < Xl < X2 < ... < Xi-l < t < Xi < ... < Xn-l < X n = b}. Now, n U{Pl,f,a) = LMkak k=l 
24 Chapter 1: Analysis and Linear Algebra where as n U(P 2 , 1,0:) = E MkO:k + M(o:(t) - O:(Xi-l)) + M'(O:(Xi) - o:(t)) k=l, k;f:.i where M = sup{/(x) : x E [Xi-I, t]} and M' = sup{/(x) : x E [t, Xi]}. If we set M i = max{M, M'}, then we have MiO:i - Mi(o:(t) - O:(Xi-l)) + Mi(O:(Xi) - o:(t)) > M(o:(t) - O:(Xi-l)) + M'(O:(Xi) - o:(t)) which shows that U ( P 2 , I, 0:) < U (PI, I, 0: ). Letting m = inf{/(x) : x E [Xi-I, t]}, m' = inf{/(x) : x E [t, Xi]}, and proceeding exactly in the same way with mi = min {m, m'} in place of M i , we see that the corresponding inequality for the lower sum follows similarly. It follows that if P 2 contains one point more than PI, then the inequalities of the Lemma hold. The general case follows by the method of induction. _ As a consequence of the last result, for any partitions PI and P 2 of [a, b], we have L(Pl, 1,0:) < L(PI U P 2 , 1,0:) < U(P 1 U P 2 , 1,0:) < U(P 2 , I, 0:). One can show that the following limits exist: --=b ! 1 do::= lim U(P, 1,0:) = inf{U(P, 1,0:) : P is a partition of [a, b]} a IPI-+O and (b Ida:= lirn L(P, I, a) = sup{L(P, I, a) : P is a partition of [a, b]}.  IPI-+O These are respectively called the upper and lower Riemann-Stieltjes inte- grals of 1 with respect to 0: over [a, b]. In particular, for any partition P, this observation yields that b --=b L(P, I, a) < i lda(x) < !/da(x) < U(P,I,a). We remark that the upper and lower sums depend on the particular choice of the partition while the upper and lower integrals are independent of the 
1.5. Concept of Riemann and Riemann Stieltjes Integrals 25 partitions. Hence, a natural question is: will the two quantities, namely the upper and lower integrals, ever coincide? The bounded function f defined on the closed interval [a, b] is said to be Riemanri-Stieltjes integrable on [a, b] if the upper and lower Riemann-Stieltjes integrals are equal. In this case, we denote their common value simply by lab f do: or lab f(x) do:(x). We call this integral as the Riemann-Stieltjes integral of f with respect to a on [a, b]. We let ROl[a, b] denote the set of all Riemann-Stieltjes integrable functions with respect to a on [a, b]. If a(x) = x, then the Riemann-Stieltjes integral reduces to the Riemann integral of f over [a, b]. In this case, the upper and lower Riemann-Stieltjes sums will be called the upper and the lower Riemann sums, respectively. The set of all Riemann integrable functions on [a, b] will be denoted by R[a, b]. In particular, we raise te following fundamental questions: . Is every monotone function on [a, b] Riemann integrable? . Is every continuous function on [a, b] Riemann integrable? . Is every bounded function which has a finite number of discontinuities in [a, b] Riemann integrable? . Is every bounded function which has an infinite number of disconti- nuities in [a, b] Riemann integrable? . Is every monotone function which has an infinite number of disconti- nuities in [a, b] Riemann integrable? Before we answer these questions, it would be interesting to develop some theorems which will easily lead to examples of Riemann integrable functions and partly answer some of the above questions. We begin with the following criterion for Riemann integrability: 1.25. Theorem. Let f : [a, b]  IR be bounded. Then f E R[a, b] iff for every € > 0 there exists a partition P of [a, b] such that U(P, f) - L(P, f) < €. Proof. This is an easy and standard result that follows from the defi- nition. We leave the proof as an exercise. - 1.26. Proposition. Every monotone function on [a, b] is Riemann integrable on [a, b]. Proof. It suffices to consider the case when f is monotone increasing on [a, b]. A similar argument works if f is monotone decreasing. Divide [a, b] into n-equal intervals and consider the partition a = Xo < Xl < X2 < ... < Xn-l < X n = b 
26 Chapter 1: Analysis and Linear Algebra with Xk = a + k(b - a)/n so that Xk - Xk-1 = (b - a)/n, that is [ (b - a) 2(b - a) ] P = a, a + n ' a + n ' . · · , b · As Xk-1 < Xk and f is increasing, we have for each k E {I, 2, . . . , n}, M k = sur{f(x) : x E [Xk-1, Xk]} = f(Xk) and mk = inf {f (x) : x E [x k -1 , X k]} = f (x 1e-1 ). Thus, n U(P, f) - L(P, f) - L(Mk - mk)(XIe - Xk-1) k=1 n b-a - L...J (f(Xk) - f(Xk-1)) n k=1 _ b - a (f(b) - f(a)) . n Now, the right hand side of the last equality approaches 0 as n  00 and so given € > 0, we can find n with b-a (f(b) - f(a)) < € n which proves the existence of a partition P with U(P, f) - L(P, f) < €. Thus, by Theorem 1.25, it follows that f is Riemann integrable. - 1.27. Proposition. If f: [a, b]  IR is continuous, then f is Riemann integrable. Proof. Let € > 0 be given. Since f is continuous on the compact set [a, b], it is uniformly continuous on [a, b] (see Corollary 2.83). Therefore, there is a 6 > 0 such that if Ix - yl < 6 and x, y E [a, b], then € If ( x) - f (y ) I < b _ a . Choose n such that ba < 6 and consider the partition [ (b-a) 2(b-a) ] P = a, a + n ' a + n ' . · · , b · Here a + k(b - a)/n = Xk so that Xk - Xk-1 = (b - a)/n. Now if x,y E [Xk-1,Xk], then Ix - yl < 6 and so If(x) - f(y)1 < €/(b- a) holds. Thus, o < M k - mk = sup {f(x)} - inf {f(x)} < b € xE[xle-t,xle) xE[xle-t,xle) - a 
1.6. Vector Spaces 27 and using this inequality we get n U(P, f) - L(P, f) = E(M k - mk)(xk - Xk-l) k=l <  ( ba ) C:a ) =€ which shows that f is Riemann integrable, by Theorem 1.25. . 1.28. Example. Consider the Dirichlet function f : IR  IR defined by { I ifxEQ f(x)= 0 ifxEIR\Q and the unit interval [0, 1]. If 0 < Xk-l < Xk < 1, then M k = sup f(x) = 1 and mk = inf f(x) = o. XE[XIe-bXIe] XE[XIe-bXIe] This shows that any partition P, L(P, f) = 0 and U(P, f) = 1. Thus, f is not Riemann integrable on [0,1]. Now, consider { l/q g(x) = 0 if x = p/q E Q (in lowest form) if x E R\Q and the unit interval [0, 1]. This function is also called the Dirichlet func- tion. Note that 9 is neither monotone nor continuous (show!) on [0, 1], but it is Riemann integrable on [0, 1]. We leave this as an exercise. As noted in the beginning of this section, to integrate such functions, the concept of Lebesgue integral is required. . 1.6 Vector Spaces An abstract mathematical system that embodies a generalization of familiar concept of vector is a vector space. We define first what a vector space is. In general, the vector spaces we shall encounter will be defined only for one of the two fields: the field IR of the real numbers or the field C of the complex numbers. When need arises, we shall specify whether we consider a complex vector space or a real vector space. 1.29. Definition. A vector space over a field IF of scalars, denoted by (V, IF) or simply by V, is a nonempty set V of objects called vectors equipped with two operations called addition and scalar multiplication described as follows: 
28 Chapter 1: Analysis and Linear Algebra (1) For u,v E V, we have u + v E V [Closed under addition] (2) For A E IF and u E V, we have AU E V. [Closed under scalar multiplication] These two operations must satisfy the following conditions for all u, v, w E V and all scalars A, J.t E 1F: (AI) u + v = v + u [Commutative with respect to addition] (A2) (u+v)+w=u+(v+w) [Associative with respect to addition] (A3) There is a vector in V, denoted by () (or ()v or 0 or Ov), called zero vector, such that u + () = u for all u E V. [Zero element] (A4) For ech u E V, there is a vector, denoted usually by -u, in V called additive inverse such that u + ( -u) = () for all u E V. (-u is called additive inverse of u E V). (SI) (AJ.t). u = A · (J.tu) [Associative with respect to scalar multiplication] (S2) A' (u + v) = A · u + A · v [Distributive with respect to addition] (S3) (A + J.t) · u = A · u + J.t · u [Additive inverse] [Distributive with respect to scalar multiplication] (S4) 1. u = u for all u E V . [Identity for scalar multiplication] Note that it does not matter how the operations of addition u+v and the scalar multiplication AU are defined. All we require is that these operation must satisfy all the above axioms. We shall first present a set of important examples of vector spaces. 1.30. Examples of vector spaces. Two simple examples are as fol- lows. (1) The field IF itself is a vector space over IF with respect to the usual addition and scalar multiplication. (2) The set Mmxn (IF) of all m x n matrices over the field IF forms a vector space with respect to the usual matrix addition and scalar multiplication. . 1.31. Examples of sets which are not vector spaces. We have 
1.6. Vector Spaces 29 (1) The set Mnxn (]F) of all n x n matrices A over the field IF with the determinant of A being zero is not a vector space because it is not closed with respect to the matrix addition. (2) If S = {A E Mnxn(JF) : detA:j:. OJ, then S is not closed with respect to the matrix addition. (3) The set of all solutions X of a nonhomogeneous system of equations described by a matrix system AX = B, where B :j:. 0, does not form a vector space. . 1.32. Space r (}Rn or CR). The space en is the higher dimensional analog of C and this space is called the n-dimensional (complex) space. Thus, the space en of all n-tuples 4 of complex numbers is defined by the Cartesian product of n-copies of C: en = {z = (Zl,Z2",.,Zn): Zk E C, k = 1,2,...,n}. The elements of en are called points or vectors, and the rules for addition and scalar multiplication, in strict analogy with the corresponding opera- tions in C, are defined in a natural way: Z+W =(Zl+Wl,Z2+W2,...,Zn+wn), AZ = (AZl, AZ2, . . . , AZ n ), where Z = (Zl, Z2,"', zn), W = (Wl, W2,..., w n ) belong to en and A E C. Thus, Z + W and AZ belong to en. Recall that Z = W iff Zj = Wj, for all j = 1,2,. . ., n. It is easy to verify that all the axioms of the vector space are satisfied. Thus, (en, C) forms a vector space. If each Zj (j = 1,2, . . . , n) is real then in this case the space is called n-dimensional real space, denoted by }Rn: . ]Rn = {x = (x 1 , X2, . . . , X n) : x k E IR, k = 1, 2, . . . , n } . Similarly, (IRn, IR) is a vector space. Unless otherwise stated explicitly we shall assume the standard operations, see Figure 1.6. According to the convenience, we can consider the element in r (en or }Rn) either as column vector: Zl Z= Zn (thought of as a n x 1 column matrix) or as a row vector: Z = (Zl,Z2,... ,zn). 4The n-tuples are regarded as vectors and are also considered as points or elements orO. 
30 Chapter 1: Analysis and Linear Algebra Y  N  x + Y = (Xl + Yl, X2 + Y2) ... ..-4  '-'"    I  I  I I I I I I o x Figure 1.6: Addition of vectors in }R2 The idea of interchanging row vector and column vector notation will be helpful while using the Matrix theory. 1.33. Space of continuous functions CF[a, b]. Denote by C(X)5 the set of all continuous complex valued functions on a compact set X (usually X will be a compact Hausdorff space). This is a simple example of a function space, Le. the space whose elements are themselves functions defined on a space. In particular, we are mainly concerned in the case X = [a, b]:6 CF[a, b] = {I : [a, b]  1F1 I is continuous from [a, b] into F} where 1F may be C or IR, and [a, b], a < b, is the bounded closed interval in III When 1F = C, the members of CF[a, b] may be regarded as a parametric representation of continuous curves in C. We remind the reader that every continuous function from [a, b] into}R has a maximum and a minimum. For I, 9 E CF[a, b] and A E C, the addition I + 9 and the scalar multiplication AI are defined in a natural way by the rules: (I + g)(t) = I(t) + g(t) (A/)(t) = A/(t) t E [a, b]. It is not hard to show that the vector space axioms are satisfied for the space CF[a, b]. 5The letter C suggests "continuous", and the definition of compactness and Hausdorff space will be discussed later, as we will not really use these concepts until we define metric spaces. 6Most of the authors use the notation C[a, b], instead of CF[a, b], but often with a word stating that whether they are dealing with complex valued or real valued functions. By continuous we mean the unbroken graph, or else one could use "E - 6" definition of continuous functions. 
1.6. Vector Spaces 31 1.34. Subspaces. Let V be a vector space (over]F). A subset 8 of V is said to be a subspace of V if 8 is itself a vector space when the addition and scalar multiplication of V are used. Some straightforward examples of subsets which are not subspaces are (1) The subset 8 = {(a, a 2 ) E IR2 : a E IR} is not a subspace of IR 2 because it is neither closed with respect to the scalar multiplication (eg. 2(1, 1)  8) nor with respect to the addition. (2) The subset S = {(a, b) E IR2 : b > O} is not a subspace of IR 2 , because (0,-1) = -1(0,1)  8. (3) The subset 8 = {(a, b, c) E IR3 : lal = Ibl = Icl} is not a subspace of IR3 because (-1,1,1) + (1,1,1) = (0,2, 2)  8. (4) Each of the subsets 8 1 = {(a,O) E IR2 : a E IR} and 8 2 = {(O,b) E ]R2 : b E IR} is a subspace of IR2 whereas the union 8 = 8 1 U 8 2 does not form a subspace of jR2, because 8 is not closed with respect to vector addition (Note that el = (1,0), e2 = (0,1) are in 8 but el + e2 = (1, 1) is not in 8). . We have the following simple characterization of a subspace whose proof is routine and hence we leave it as an exercise. 1.35. Proposition. A nonempty subset 8 of a vector space V over the field IF is a subspace of V iff AU + J.tV E 8 whenever u, v E S, A, J.t E F. We note that a subset 8, which does not contain the zero vector 8, of a vector space V cannot be a subspace. Now we present simple examples of subspaces. (1) The set of all polynomials defined on IR, denoted by P(IR) is a subspace of the vector space V of the set of all functions defined from IR into ]R over the field III (2) The set 8 = {p E P(IR) : p(O) = O} is a subspace of P(IR). . On the other hand, if we define 8 1 - {p E P(IR) : p(O) = I} 8 2 - {p E P(IR) : p(x) < O} 8 3 - {p E P(IR) : p(x) > O} then we see that none of 8 1 , 8 2 and 8 3 forms a subspace of P(IR). 1.36. Linearly independent sets and bases. Let V be a vector space over a field IF. A linear combination of a set of vectors {VI, V2, . . . , v m } in V is an element of V and is of the form m L CkVk, where Ck'S are in F. k=1 
32 Chapter 1: Analysis and Linear Algebra A set of vectors {Vl, V2, . . . , v n } in a vector space V over a field ]F is called linearly independent if there exist no scalars Cl, C2, . . . , C n E ]F such that n ECkVk = 0 and k=l n E I C kl 2 > O. k=l If there exist scalars Cl, C2, . . . , C n E 1F which satisfy the above condition, then we say that the set of vectors {Vl, V2, . . . , v n } is linearly dependent. In the later case, anyone of the Vk'S, with Ck :j:. 0, will be a linear combination of the others: Vk = t ( Cj ) Vj. i=l Ck i1e A set of vectors {Vl, V2, . . . , v m } in a vector space V is called a spanning set for V iff every vector v E V can be written as a linear combination of Vl, V2, . . . , V m . We denote the so spanned set simply by span { Vl , V2, . . . , v m } and therefore, span {Vl, V2,. · · , v m } = { f CkVk: Ck E F, k = 1,2,..., m } . k=l We now come to the important concept of basis of a vector space. A set B = {Vl, V2, . . . , v n } of vectors in a vector space V forms a basis for V iff (i) B is linearly independent (ii) V = span { Vl , V2, . . . , v n }, Le. every vector in V is a linear combina- tion of the elements of B. If {Vl, V2,.. . . , v n } is a linearly independent set, then we have n n n E akVk = E bk V k ==> E(ak - bk)Vk = 0 k=l k=l k=l ==> ak - b k = 0 for k = 1,2, . . . , n, and in view of this, we note that (i) and (ii) together is equivalent to the statement that every vector in V is uniquely expressible as a linear combination of the vectors Vl, V2, . . . , V n . The vector space V is said be finite dimensional if there exists a finite set of vectors that spans V, Le. the number of basis elements is finite. Otherwise, we say that V is an infinite dimensional vector space. If a vector space V has a basis consisting of n vectors, then we say that V has a dimension n and we write dim V = n. The subs pace containing only the zero vector, namely {OJ (zero space) of V, is said to have finite dimension. Naturally, we define dim {OJ = O. We use ek to denote the n-tuple (Xl, X2, . . . , x n ) where Xk = 1 and x j = 0 for 1 < j :j:. k < n, Le. ek is the element in }Rn with 1 in the k-th 
1.6. Vector Spaces 33 place and zero elsewhere. Then En = {el, e2, . . . , en} becomes a basis for ]in and is called the standard basis for }Rn. Let B = {VI, V2, . . . , v n } be a basis of V. Then for every vector V E V, we have the unique representation V = ClVl + C2 V 2 +... + CnV n in scalar unknowns Cl, C2, . . . , Cn. The column/row vector Cl [V]B := C n or in consistence with our earlier discussion [V] B : = (Cl, C2, . · · , c n ) , is called the coordinate vector of V with respect to the basis B. Conversely, given [V]B = (Cl, C2,. .. , cn), we can recover the vector by writing v = EZ=l CkVk. If V = }Rn and if En = {el, e2,. . . , en}, then for v E }Rn, we have [V]En = v. 1.37. Proposition. The dimension of en over C is n. Proof. We need to show that there exists a set of n linearly indepen- dent vectors in en over C but every set of n + 1 or more vectors in en is linearly dependent. Clearly, {el, e2, . . . , en} is linearly independent since each (Zl, Z2, . . . , zn) E en can be written as E j 1 zjej and this is a zero vector iff all the Zj'S are zero. On the other hand, suppose that we are given (n + 1 )-vectors VI, V2, . . . , Vn+l, where Vj = (alj, a2j, . . . , anj), j = 1, 2, . . . , n + 1. Form the matrix system AZ = 0, where A is a matrix of order n x (n + 1) with Vj as the j-th column of A and Z = (Zl, Z2,. . . Zn, Zn+l)(n+l) xl, or with Zl Z= Zn+l i.e. the homogeneous system of equations are determined by (1.38 ) n+l L ZjVj = O. j=l 
34 Chapter 1: Analysis and Linear Algebra Since this is a system of n equations with n+ 1 unknowns Z1, . . . , Zn+1, there exists a nontrivial solution C = (C1, C2, . . . , C n + 1) (n+ 1) x 1 where at least one of the Cj'S is nonzero, satisfying (1.38), that is n+1 AC = 0 = E CjVj = O. j=1 Thus, V1, V2, . . . , V n +1 are linearly dependent over C. . 1.39. Remark. Is the dimension of en over IR 2n? . 1.40. Proposition. Let {V1, V2, . . . , v m } span a finite dimensional vector space V. If S = {W1, W2, . . . , w n } is any set of vectors in V, then S is linearly dependent whenever n > m. Alternatively, if S is linearly independent then n < m. Proof. Let n > m. By the definition of spanning set, we can write m Wj = E aijVi, for each j = 1, 2, . . . , n, i=1 so that tCjWj = ( ta1jCj ) V1 +... + ( tamjCj ) vm. 3=1 3=1 3=1 Consider the system of equations n E aij Cj = 0, for each i = 1, 2, . . . , m, j=1 or equivalently, the matrix equation AC = 0, where A is an m x n matrix with n > m. Note that we have more unknowns than the number of equa- tions in this linear system of equations and therefore, we have a nontrivial solution c = (ci, . . . , c)n x 1. Thus, for this nontrivial solution c, we have . n EcjWj = 0 j=1 so that the set S is linearly dependent. . 1.41. Corollary. Let V be a finite dimensional vector space with dim V = n. Then we have the following statements: (1) Every set of linearly independent vectors ofn elements is a basis for V. In particular, the number of elements in any two bases of V is same. 
1.7. Linear Transformations between Vector Spaces 35 (2) Any spanning set for V must contain at least n elements. Proof. (1) Let S = {V1, V2, . . . , v n } be a linearly independent set of vec- tors in V. We show that this set spans the space V. Let v E V be arbitrary. Then, the set {V1, V2, . . . , V n , v} is linearly dependent (since dim V = n) and therefore there exist scalars c, C1, . . . , c n , not all of them being zero, such that n cv + ECjVj = O. j=1 As S is a linearly independent set, we have c :j:. 0 so that n V = E(-Cj/C)Vj. ;=1 Thus, v E span (S) and hence S is a basis for V. The next part follows from Proposition 1.40. (2) Let the set B = {V1, V2, . . . , v m } span V. If B is linearly independent, then, by definition, this would be a basis for V and therefore, by part (1), we have m = n. If B is linearly dependent, then at least one of the vectors, say Vk, is a linear combination of the remaining vectors in B. Thus, Vk can be removed from B so that the resulting subset of B forms a spanning set for V. This process may be continued (if necessary) until we obtain a linearly independent spanning set containing less than m vecto)"s and consequently, by part (1), we have m > n. _ The space CF[a, b], where IF = C or JR, is an infinite dimensional vector space. In fact, the subset {t n }nO C CF[a, b] is linearly independent, since n f(t) = E akt k = 0 => ao = a1 = . · . = an = 0 for each n > 0, k=O so that the space CF[a, b] cannot be finite dimensional. 1.7 Linear Transformations between Vector Spaces Let V, W be two vector spaces, both over the same field IF. Now, we shall briefly look at the mappings from V into Wand such mappings are also called operators or transformations. Our particular interest is when the operator maintains a special "structure" in the following sense: An operator T : V  W is said to be additive if it preserves the addition: T(V1 + V2) = T(V1) + T(V2), for V1, V2 E V, and is said to be homogeneous if it preserves the scalar multiplication: T(av) = aT (v), for a E IF, v E V. 
36 Chapter 1: Analysis and Linear Algebra Note that we also follow the convention of writing T(v) := Tv whenever this notation is convenient. The map T : V  W is called a linear trans- formation/operator if it is both additive and homogeneous. Some authors call a linear transformation T from V into itself as a linear operator on V. But, we do not follow this convention. A transformation that is not linear is the one which does not satisfies either the additivity condition or the homogeneity condition or both. A linear transformation T : V  W is called isomorphism if it is bijective. In such cases, we say that V and W are isomorphic. We first observe that every linear transformation T : V  W satisfies TOv = Ow, where Ov and Ow are the zero elements of V and W, respec- tively. A simple example of a transformation which is additive but not homogeneous is given by T : C  C, z I-t z , with IF = C. An example of a transformation which is homogeneous but not additive is given by T : }R2 \ { (0, y) : y E }R}  IR, X (Xl, X2) I-t -. Xl Clearly, a linear transformation preserves the structure of the linear com- bination of the vectors and, for all Vl, . . . , V n in V and all scalars Cl, . . . , C n in F, we have T ( tCkVk ) = tCkTvk' k=l k=l In particular, each linear transformation T maps a line segment connecting Vl and V2 into the line segment between TVl and TV2: T(AVl + (1 - A)V2) = ATvl + (1 - A)Tv2, A E [0, 1]. Finally, we remark that if we are given a linear transformation T from a finite dimensional vector space V into another vector space, then we can determine the transformation in the following sense: if {Vl,.", v n } is a basis for V, then for any vector v E V we have v = E 7 1 CjVj for some scalars Cj (j = 1,. . . , n) in IF so that Tv = E 7 1 cjTvj. 1.42. Transformations which are linear, and not linear. First, we list a few simple examples of linear transformations: (1) T:}R2  }R2, X = (Xl, X2) I-t (Xl + X2, Xl - X2). (2) T:}R2  }R2, X = (Xl,X2) I-t (Xl +X2,Xl). . (3) T:}R2  }R2 , X = (x 1 , X2) I-t (Xl - X2, Xl)' (4) T: IRn  }Rm, X I-t Ax, where A is an m x n matrix. (5) T:}R3 -+ IR 2 , X = (Xl, X2, X3) I-t (Xl, X2). (6) T: Mnxn (}R)  Mnxn (IR), A I-t A - At. (7) T: Mmxn (IR)  Mnxm (IR), A I-t At, or A I-t _At. 
1.7. Linear Transformations between Vector Spaces 37 (8) T: Mnxn (JR)  Mnxn (IR), A I-t AB - BA, where B is a fixed n x n matrix. (9) T: Mnxn (IR)  IR, A = (aij) I-t Elin au. (10) T: CR[a, b]  IR, f I-t J: f(t) dt. Next, we list below some simple examples of transformations which are not linear: (1) T: Mnxn (JR)  JR, A I-t det A. (2) T: IR2  IR 2 , X = (Xl, X2) I-t (Xl + k l , X2 + k 2 ), where at least one of k l , k 2 is a fixed nonzero real number. (3) T: IR  IR, X I-t x2 . (4) T: IR 2  JR2 , X = (Xl, X2) I-t (lxll, 2Xl - X2). (5) T: IR 2  JR2 , X = (Xl, X2) I-t (Xl + X2, Xl + 1). (6) T: IR2  ]R2, X = (Xl,X2) I-t (XlX2,Xl). (7) T:]R2  ]R2, X = (Xl, X2) I-t (Xl + 1, XlX2). Let L(V, W) denote the space of all linear mappings from a vector space V into another vector space W. -In fact, if Tl : V  Wand T 2 : V  W are two linear transformations and A is a scalar in IF then the sum Tl + T 2 and the scalar product ATl are defined by { (Tl + T 2 )v = Tlv + T 2 v (AT1)V =A(TIV) , forallvEV,AEF. Then, for linear operators T l , T 2 E L(V, W) and for c E IF, v, v' E V, we have (T l + T)(cv + v') - Tl(CV + v') + T 2 (cv + v') - [CTl v + Tl v'] + [cT 2 v + T 2 v'] - C(Tl + T 2 )v + (T l + T 2 )v', and similar identity holds for ATl' Thus, Tl + T 2 and ATl are also linear transformations so that the set L(V, W) is easily seen to be a vector space under the pointwise operations of addition and scalar multiplication defined above. The zero vector in L(V, W) is the zero transformation 8 : V  W, v I-t Ow, where Ow represents the zero vector in tv; and 8 or 0 are also used in place of Ow. The additive inverse of T E L(V, W) is - T defined by - T = (-l)T. IT W = V, then we often use the notation L(V) instead of L(V, V). The identity transformation on V is defined by Iv : V -+ V, v I-t v, for all v E V. 
38 Chapter 1: Analysis and Linear Algebra Whenever there is no confusion, we often write I in place of Iv. Let T E L(V, W), where V and Ware finite dimensional vector spaces. Then the set NT of vectors v E V for which Tv = 0 is called the null space of T, and the set RT of vectors W E W such that Tv = W for some v E V, is called the range space of T. From this definition, it is clear that "T is onto iff RT = W". Sometimes, we refer the spaces NT and RT as the kernel of T and the image of T, respectively. In this terminology, the usual notation for NT and RT will be KerT and ImT, respectively. As a consequence of these definitions we can derive certain properties of the spaces NT and RT. 1.43. Theorem. Let T : V -t W be a linear map from the vector space V into the vector space W. Then we have (i) NT and RT are subspaces of V and W, respectively. (ii) T is one-to-one iff NT = {OJ. Proof. (i) Let T : V  W be a linear map. Consider NT = {v E V : Tv = O} and R T = {w E W : Tv = w for some v E V}. Then TO = T(O.v) = O.Tv = 0 and therefore, NT is nonempty, since o E NT. If Vl,V2 E NT, then for a,{3 E F, we have aVl + {3v2 E NT, because T(avl + {3v2) = aTvl + {3 Tv 2 = 0..0 + ,8.0 = O. Again, since TO = 0, RT is nonempty. If Wl , W2 E RT, then there exist Vl, V2 in V such that TVl = Wl and TV2 = W2. Then, for a, {3 E F, there exists aVl + {3v2 in V such that T(avl + {3v2) = aTvl + {3 Tv 2 = aWl + {3w2 so that aWl + {3w2 E RT. Thus, NT and R T are the subspaces of V and W, respectively. (ii) Assume that T is one-to-one. If v E NT, then TO = 0 = Tv which, because of one-to-oneness of T, gives v = o. Thus NT = {OJ. Conversely, if NT = {OJ then, because of the linearity of T, we have TVl = TV2 ==> T(Vl - V2) = 0 ==> Vl - V2 E NT ==> Vl - V2 = 0 and therefore, T is one-to-one. . Note that Theorem 1.43(i) applies only when the map is linear. For example, if T : IR  IR is defined by x I-t x 2 , then the range RT is IRt which is not a subspace of IR. We define the rank of T, denoted by rank (T), to be the dimension of the subspace RT, and the nullity ofT, null (T), to be the dimension of the subspace NT. Clearly, N] = {OJ, R[ = V, No = V and Ro = {OJ. 
1.7. Linear Transformations between Vector Spaces 39 Further, no confusion should arise from the fact that we shall not distinguish either zero/identity transformation or zero/identity element in V and that in W. One of the fundamental results in linear algebra is the following theorem which is known as Rank-Nullity theorem. This theorem gives a complete characterization of one-to-one and ontoness in terms of the null space and the range space of T. 1.44. Theorem. Let T : V  W be linear, where V is a finite dimensional vector space. Then we have dim NT + dim RT = dim V. In particular, we have ( a) If dim V > dim W, then T cannot be one-to-one. (b) If dim V < dim W, then T cannot be onto. (c) If dim V = dim W and dim NT > 0, then T cannot be onto. Proof. Case (i): IT dim V = n and NT = V, then RT = {OJ so that the theorem holds trivially. Case (ii): Assume that dim NT = k > 0 and dim V = k + m. Let B = {V1, V2, . . . , Vk} be a basis for NT. Then we can expand the set B so that the extended set { V1 , . · · , Vk, Vk+1 , · · · , Vk+m} forms a basis for V. Thus, to complete the proof, it suffices to show that the set {TVk+1, . . . , TVk+m} forms a basis for the range space RT. To prove this, we choose an arbitrary point wERT. Then there exists v E V such that Tv = w, where v = E ;+  CjVj for some scalars Cj. Since T is linear and TVj = 0 for j = 1,2,. . . , k, it follows that k+m k+m W = Tv = L cjTvj = L cjTvj. j=1 j=k+1 Thus, {TVk+1,. . . , TVk+m} spans the range space R T . Next, we show that this set of vectors is linearly independent. Suppose that k+m L cjTvj = O. j=k+1 
40 Chapter 1: Analysis and Linear Algebra Since T is linear, this can be written as ( k+m ) T E CjVj j=k+1 =0 so that k+m u:= E CjVj E NT, j=k+1 and since {V1, V2,. . . , Vk} is a basis set for NT, we must have k U = ECjVj j=l for some scalars Cj, 1 < j < k. From the last two equations, it follows that k k+m E CjVj + E (-Cj)Vj = O. j=l j=k+ 1 As {V1, . · . , Vk, Vk+1, . . . , Vk+m} is linearly independent, it follows that Cj = o for j = 1,2,..., k + m which shows that the set {TVk+1,..., TVk+m} is linearly independent. Case (iii): IT dim NT = 0, Le. NT = {OJ, and if {V1, V2,. . . , Vk+m} is a basis for V, then it follows that the set {TV1, TV2, . . . , TVk+m} forms a basis for the range space RT, and once again the theorem follows. . The following examples help us to understand the ideas behind Theo- rems 1.43 and 1.44. (1) Any linear map T : ]R4  IR3 cannot be one-to-one. (2) Any linear map T : M 2X2 (IR)  IRs cannot be onto. (3) The linear map T : IR 2  IR2, (Xl, X2) I-t (-Xl + 2X2, 0) is neither one-to-one nor onto, since NT = {(X1,X2) : Xl = 2X2, X2 E JR.} so that dim NT = 1. (4) The linear map T : P2(IR)  P 2 (1R), p(x) = ax2 + bx + C I-t ax2 + (b - 2c)x + a - b + 2c is neither one-to-one nor onto, since NT - {p(x): a = 0, b = 2c, C E IR} - {p(x): p(x) = c(2x + 1), C E IR} so that dim NT = 1. (5) The linear map T : P3(IR)  P 3 (IR), p(x) I-t p'(x) is neither one-to- one nor onto. 
1.7. Linear Transformations between Vector Spaces 41 (6) The linear map T : Pn{Ii)  JRn+1, E=o akx k I-t (ao, a1, . . . , an), is one-to-one and onto. If dim V = dim W, then, from Theorem 1.44, it follows that T is one-to-one <==} T is onto. Here P n (F) denotes the set of all polynomials of degree less than or equal to n in z (if IF = C) (and respectively in x if 1F = JR) with complex coefficients over the field C (with real coefficients if 1F = JR). The linearity of T in Theorems 1.43 and 1.44 is essential, for it is a simple exercise to construct examples of functions from JR into JR that are not one-to-one but are onto, and vice versa. Now we have the following 1.45. Corollary. Let V and W be two finite dimensional vector spaces over the same field IF. Then we have (i) If T : V  W is bijective, then dim V = dim W (ii) If dim V = dim W, then there exists a bijective linear transformation from V into W. Proof. If T is one-to-one and onto, then dim NT = 0 and RT = W so that by Theorem 1.44, we have dim W = dim V. To prove the second part, we assume that dim W = dim V. Let B = {V1, V2, . . . , v n } and B' = { W1, W2, . . . , w n } be any two bases of V and W, respectively. Then, we see that there exists a unique linear transformation T : V  W such that TVj = Wj for j = 1,2,. . . , n (how?). If W E W is an arbitrary vector, then we have the unique representation n W = L CjWj for some scalars Cj. j=1 Now, for v = E j 1 CjVj E V we have n n Tv = LCjTvj = LCjWj = W j=1 j=1 which implies that T is onto and therefore, dim RT = dim W = dim V. Again, by Theorem 1.44, we have dim NT = 0, i.e. T is one-to-one. - From Corollary 1.45, we observe that if dim W = dim V, then T : V  W is one-to-one iff T is onto. 1.46. Example. Consider the linear map T : Mnxn (lR)  Mnxn (IR), A I-t A - At. Now, to find the null space NT, we need to solve T A = A - At = 0 
42 Chapter 1: Analysis and Linear Algebra for A. If A = (aij)nxn, then A - At = (aij - aji) = (bij), so that 0 b I2 bIn -b 12 0 b 2n A - At = , where b ij = aij - aji, -bIn -b 2n 0 which shows that A - At = 0 yields the condition aij = aji for each i and j. Thus, NT is the set of all n x n symmetric matrices and therefore n 2 -n dimNT=n+ 2 n(n + 1) 2 Since RT = {B E Mnxn(IR) : A - At = B, for A E Mnxn(IR)}, we obtain 0 b I2 bIn -b 21 0 b 2n RT= b ij E JR . -b nl -b n2 0 and therefore, we find that R T is the set of all skew-symmetric matrices. Thus, n 2 -n dimRT = 2 n(n - 1) 2 Note that T is neither one-to-one nor onto. . Recall that the dimension of IF (where IF is either C or JR) over itself is one. Linear maps from a vector space V into IF, T : V  IF, play a special role in the theory of vector spaces, and therefore they have a special name "linear functionals" , see Chapter 5. The definite integrals of continuous function is one of the most important examples of linear functionals in mathematics. 1.8 Inequalities Geometric and integral inequalities have a prominent place particularly in real, complex and functional analysis. Among the tools used in establishing these inequalities, convex and concave functions are especially important. In this section, it is shown that these tools yield several important inequal- ities. There are some standard reference books on inequalities, for example [HLP, Mi]. Most of the inequalities in this section are valid in more general 
1.8. Inequalities 43 Figure 1.7: Convex and non convex domains in the complex plane setting. However, we present only those which are relevant to the scope and the topics of this book. 1.47. Definition. A nonempty subset S of a vector space V is said to be convex ifAxl + {I - A)X2 E S whenever Xl, X2 E S and 0 < A < 1. Geometrically, this means that given two arbitrary points in a convex set, the line segment joining them is also in the set. We note that the line passing through Xl and X2 is the set {AXI + {I - A)X2 : A E JR} and therefore, the restriction {AXI +{1-A)x2 : A E [0, I]} is the line segment [Xl, X2]. Clearly, a singleton set, the interior of circles and ellipses in JR2, the solid ellipsoids and cubes in JR3 are convex. Indeed, if V is a vector space then every linear subspace in V is trivially convex, see Proposition 1.35. Let P (POl resp., where a E JR) denote the set of all analytic functions p on  = {z: Izl < I} such that p{O) = 1 and Rep{z) > 0 (Reeiap{z) > 0 resp.) in. Then P is a convex set whereas P a is a nonconvex set, see Figure 1.7. A real-valued function f on a convex set S is said to be convex on S if the inequality f{AXI + (1- A)X2) < Af{XI) + (1- A)f{X2) holds for all Xl,X2 E S, 0 < A < 1. Hthe above inequality is reversed, then f is said to be concave. Alternatively, f is concave iff - f is convex. Thus, the convexity (concavity) of the real valued function f : [a, b]  JR means that the chords joining the two points on the graph of f always lie above (below) the graph of f, see Figure 1.8. A real-valued function f on S is called logarithmically convex (concave), or simply log-convex (log-concave) if f is positive and log f is convex (concave). The function is strictly convex (concave) if the functional inequality above is strict. 
44 Chapter 1: Analysis and Linear Algebra y y y= f(x) y = f(x) Xl o X o X Figure 1.8: Convex and concave curve in the real variable case 1.48. Example. Consider f : IR  IR defined by f(x) = z2. Then with J.t = 1 - A and A, J.t > 0, we have Af(x) + J.tf(y) - f(AX + J.ty) = AJ.t(X - y)2 > 0 which proves the convexity of the square function. . It is an important observation that the intersection of arbitrary collec- tion of convex sets is trivially a convex set (see also Exercise 1.78). The following fundamentalf;;ult is easy to derive by the method of induc- tion. 1.49. Proposition. Let S be a convex set in V, Xk E S and Ak > 0 for k = 1,2, . . . , n such that E=l Ak = 1. Then E  1 AkXk belongs to S. IT S c V, then the intersection of all the convex sets containing the given subset S is called the convex hull of S and is denoted by co(S), see Figure 1.9. In fact, it can be seen that co(S) := { t AkXk : Xl,..., X n E S, A1'...' An > 0, k=l t Ak = I } · k=l The closed convex hull of S is the intersection of all the closed convex sets which contains the given set S as we see in the foll.owing proposition. 1.50. Proposition. The set co(S) is the smallest convex set con- taining the given set S. Proof. By definition S c co(S). To show that co(S) is convex, we let X = E  1 AkXk and y = E  1 AX, where Ak, A are nonnegative real 
1.8. Inequalities 45 ., , , , , , , I I t. Figure 1.9: Description for convex hull of a set numbers such that E=l Ak = E ;; 1 A = 1. Then, it can be easily seen that m+n AX + (1 - A)Y = E J.tkX-k k=l where { X k Xk = x'. , for k = 1, 2, . . . , n for k = n + j, j = 1, 2, . . . , m and { AAk for k = 1, 2, . . . , n J.t k = , ' , '. £ k .. 1 2 1\1\ lor = n + J, J = , ,..., m. Note that J.tk > 0 and E  n /.-Lk = 1. The convexity of co(S) follows. . 1.51. Examples. (i) Let V = {x = (Xl,...,Xn) E}Rn: E  lX = I}. Then n co(S) = {x E}Rn: Ex < I}. k=l (ii) Let V = C and S be the circle {z: Izi = R}, R > o. Then co(S) is the closed disc {z: Izl < R}. (iii) Let V = C and S = {Zl, Z2}. Then co( S) = [Zl, Z2], the line segment connecting Zl and Z2. . Next we list some of the basic properties of convex and concave functions from which several classical examples of these functions defined on (a, b) may be obtained. For a detailed discussion about this topic we refer to [Roc]. Now, we provide a list of basic properties for the case of convex functions and similar results follow for the case of concave functions. 
46 Chapter 1: Analysis and Linear Algebra 1.52. Proposition. Let I be a real valued function defined on [a, b]. Let g: [c,dj  IR where the range of I is contained in [c,dj. Then we have the following statements: (I) Let I be differentiable on (a, b). Then I is convex on (a, b) iff I' is increasing on (a, b). (2) Let! be twice differentiable. Then I is convex on (a, b) iff the second derivative I" is nonnegative throughout (a, b). (3) If I and 9 are convex and 9 is increasing, then the composite function 9 0 I is convex on (a, b). Proof. First, we give a proof of (1). The remaining cases can be verified easily. Let x' E (x, y), where x, y E (a, b). Then I is convex on (a, b) iff (x',/(x')) lies on or below the line segment joining (x,/{x)) and (y,/(y)) in IR 2 . Thus, I is convex iff f(x') - f(x) < f(y) - f(x') for a < x < x' < y < b. x' - x y - x' If we apply Mean value theorem 7 from calculus to the last inequality, we see that I is convex on (a, b) iff I' (x) < I' (y) and the proof follows. (2) The reverse implication in (2) is really easy. If I"(x) > 0, then I' is increasing. Thus, by Mean value theorem, we have I(AX + (1 - A)Y) - I(x) < [/'(AX + (1- A)y)][(1 - A)(Y - x)] < ((1- A)/A)[/(y) - I{AX + (1- A)Y)] which, after a simplification, is just what we require to prove. . For instance, using Proposition 1.52 we find that the following functions are convex: (i) I (x) = x P , P < 0, p > 1, x E (0, 00 ) , (ii) I{x) = eX, x E IR, (Hi) I{x) = -log x, x E (0,00), (iv) I(x) = x log x, x E (0,00). . 1.53. Proposition. (Jensen's inequality) Let I be convex in (a, b) and let {Xl, X2,. . . , x n } be a set of points in (a, b). Then we have f (  AkXk) <  Akf(Xk) 7Mean value theorem asserts that if I is continuous on [a, b] and differentiable on (a, b) then, for some c E (a, b), one has I' (c) = (f(b) - f(a»/(b - a). This observations says that there exists a point c such that the slope of tangent line at c equals the slope of the secant line from (a,/(a» to (b, f(b». 
1.8. Inequalities 47 for Ak > 0, E  1 Ak = 1. Proof. From the definition of convexity of ! on (a, b), we see that f is convex on (a, b) iff (1.54 ) ! (A1 X 1 + A2 X 2) < A1!(X1) + A2!(X2) for A1, A2 > 0 and A2 = 1 - A1. Now we show the Jensen's inequality by the method of induction. Assume that it is true for n - 1, Le. f (  AXk) <  Af(Xk) for A > 0, E - : Ak = 1. Assume E Z- : Ak = 1 - An. Then there exists at least one p, p E {I, 2, . . . , n}, such that Ap is strictly less than 1, and without loss of generality, we can assume that An < 1 and therefore, we can write n ( n-1 Ak ) L AkXk = AnXn + (1 - An) L 1 _ A Xk =: AnXn + (1 - An)Y' k=l k=l n If we let A = Ak / (1 - An), then n-1 '" A' = A1 + A2 + · · · + A n -1 = 1 - An = 1 L...J k I-A I-A. k=l n n Thus, y E (a, b) and, therefore, we have (  AkXk) < Anf(xn) + (1 - An)f(y) n-1 A < Anf(xn) + (1 - An) L 1 _ \ f(Xk) k=l n n - L Ak!(Xk) k=l and the Jensen's inequality follows. . 1.55. Remark. Let 0 < Xl < X < X2 with \ _ X2 - X d \ _ X - Xl 1\1 - an 1\2 - . X2 - Xl X2 - Xl Then, the convexity condition (1.54) for !(x) = -logx. becomes (see also (1.67) ) (1.56) XlX-'\l < A1X1 + (1- A1)X2 
48 Chapter 1: Analysis and Linear Algebra so that, by symmetry, the inequality (1.56) holds for all positive Xl, X2 and o < Al < 1. . A direct and simple application of Jensen's inequality applied to -log X gives the following result (for an equivalent version of this result we refer to Exercise 1.79). 1.57. Proposition. Let Xl, X2,. . . , X n , Al, A2,. . . , An be nonnega- tive real numbers with E Z- l Ak = 1. Then n n II X1c < E Akxk. k=l k=l The right hand side and the left hand side of this inequality are respec- tively called weighted arithmetic mean and weighted geometric mean of the numbers Xl, X2, . . · , X n with respect to the weights Al, A2, · · · , An. Let f be analytic in  = {z : Izl < I} such that f'(z)  0 in . Identifying ]R2 with C, we say that f is convex in  if f() is a convex set in C. (Note that the unit disc itself is a convex domain in C). Analytically, convex functions f on the unit disc  are characterized by the condition [Du, Porn] ( Zf"(Z) ) Re 1 + j'(z) > 0, z E . Consider the following functions z, z/(1 - z), z/(1 + z), -log(1 - z) and the odd function 1 ( 1 + Z ) 2 log 1 - z · Using the above analytic characterization, it can be easily seen that these functions are convex in the unit disc . Note that the function z/(1 - z2) is not convex in  (see Figures 1.10-1.12). 1.58. Proposition. (Triangle Inequality) For a, b E C, we have (1.59) la + bl < lal + Ibl, and equality holds in this inequality iff a = tb, t > O. We also have the inequality (1.60 ) la + bl < lal Ibl 1 + la + hi - 1 + lal + 1 + Ibl' 
1.8. Inequalities 49 y v 1 u> -- 2 ,.... -':.. ".. , '-. iJ..., . ,..;. # '::'-"...,. ..". .:-, . .:- ,-, , \ 1 -1:0 21 I t I: L___ u ,- 0 \." \:.. \.:: . ,- .- .,.....:--- '-' X :-- , ) i ..>-' Figure 1.10: Image of 6. under j(z) = z, z/(1 - z) v '>.: J 1 1 1 1 .1 o 1 1 _ 2 1 1 J 1 1 ___.J u u<! 2 Figure 1.11: Image of 6. under j(z) = z/(l + z) v + i1r 4 .: ........?"'......  .......-r:........-: .- _":_:.'"-..:'..."''' -"!'"!'  - - _.- _.- -...-:--....- _.-  - -. -- - - _.- -- - - ""!"':..--_... o i u  "_'.;_ _ _"-Mo.-_.-._... _ _.,.. ..-.w-.................j;._ _ .;.w M/-'._"Mr _.:....,-.;_ ._*.;4JM. _ Mill _ MIl _ ".;.,.".;:..::"_....... _ _-......:......._-_..._ _ _....--..ii..-._ - - .....:... i1r 4 Figure 1.12: Image of 6. under j(z) = (1/2) log«l - z)/(1 + z» 
50 Chapter 1: Analysis and Linear Algebra Proof. For the proof of both (1.59) and (1.60), we first note that there is nothing to prove if either a = 0 or b = 0 or a = -b. Therefore, we can assume that a -# 0, b :j:. 0 and a + b :j:. O. Now, the definition of modulus gives that la + bl 2 - (a + b)( a + b) - lal 2 + 2Re (ab) + Ibl 2 < lal 2 + 21abl + Ih1 2 , since Re z < IRe zl < Izl, _ lal 2 + 21allbl + Ibl 2 _ (Ial + Ib!)2, so that la + bl < lal + Ibl and the equality holds in this inequality iff labl = Re (ab), i.e. ab is purely real and nonnegative; which means that a = sib = tb, for some s, t > O. Now (1.59) follows. Assuming a :j:. 0, b :j:. 0 and a + b :j:. 0, we see that the inequality (1.60) is then equivalent to the inequality 1 <1- 1 _ 1 1 + la + bl - 1 + lal 1 + Ibl which by multiplication gives (1.61 ) la + hl(1 -lab!) < lal + Ibl + 2la b l. This inequality is trivially true if 1 < labl. Thus, it suffices to show the last inequality only for 0 < labl < 1. However, if 0 < labl < 1, we have a stronger inequality la + bl(1 -lab!) < (Ial + Ib!)(1 -lab!) < lal + Ihl < lal + Ibl + 2labl, thanks to the triangle inequality (1.59). This reasoning verifies (1.61) and hence the assertion. _ 1.62. An alternate proof of the inequality (1.60). Applying the triangle inequality la + bl < lal + Ihl, we directly see that (with la + bl  0) la+bl 1 + la + bJ ( 1 ) -1 - 1 + la + bl ( 1 ) -1 < 1 + lal + Ibl lal + Ibl 1 + lal + Ibl lal Ibl - 1 + lal + Ibl + 1 + lal + Ibl lal Ibl < 1 + lal + 1 + IhI' 
1.8. Inequalities 51 The inequality (1.60) also follows from a general result that appears in [AVV, Remark 2.116(e)]. To give another proof of (1.60) we consider the function t 1 f:(-l,oo)IR, tt-t l+t =l- l+t ' which is clearly increasing for t E (-1,00). Further, since la + bl < lal + Ibl, we deduce that la + bl tal + Ibl tal Ibl 1 + la + bl = f(la + bl) < f(lal + Ibl) = 1 + lal + Ibl < 1 + lal + 1 + IbI" There is a general procedure to obtain new metrics from old metrics which yields this triangle inequality as a special case. For this procedure, see Remark 2.41. . 1.63. Proposition. For A E (0, 1], we define (i) I(t) = 1 - A + At - t'x (ii) g(t) = 1 + t'x - (1 + t)'x. Then the functions I and 9 are nonnegative for all t > o. Proof. (i) Clearly I' (t) = A(1 - t'x-l) so that I'(t) < 0 for all t E (0,1), and f'(t) > 0 for all t E (1,00). Thus, for all t > 0, we have I(t) > f(l) = 0 with equality iff t = 1. (ii) It is easy to see that (see also Lemma 2.29 and Remark 2.31) g' (t) = A(1 + t)'x-l [(tj (1 + t))'x-l - 1] > 0 for t > 0 so that g(t) > g(O) = 0 for t > O. . IT we substitute t = ajb, and A = p in the function 9 defined in Propo- sition 1.63, then for 0 < p < 1 and a, b > 0 we have the inequality (a + b)P < a P + if and there is nothing here to prove if a = 0 or b = O. Using this inequality we find that n n n n L IZk + wkl P < L(lzkl + IWkl)P < L IZkl P + L IWkl P k=1 k=1 k=1 k=1 for complex numbers ZI,Z2,... ,Zn, Wl,W2,... ,W n . In fact, since I(x) = x P , P > 1, is convex on [0,00), we see that I(tx + (1- t)y) < tf(x) + (1 - t)f(y) 
52 Chapter 1: Analysis and Linear Algebra and in particular if x = aft and Y = b/(1 - t), where a, b > 0, we deduce that (1.64) (a + b)P < tl-Pa P + (1 - t)l-PlJP for every t in (0,1), and the equality occurs in the last inequality iff t = a/(a + b). Thus, we have (1.65) inf [tl-Pa P + (1 - t)l-plJP] = (a + b)P; tE(O,I) . or equivalently, inf [tl-Pa + (1 - t)l-Pb] = (a l / P + bl/P)P. tE(O,I) 1.66. Remark. For an alternate proof of the nonnegativeness of the function f in Proposition 1.63, we assume A E (0, 1) and consider the function f : IR+  IR+ , t t--+ t l - A . Then f(t) > 0 for t E IR+ and f satisfies the conditions of Mean value theorem in [a, b], b > a > O. Therefore, there is a point c in the open interval (a, b) such that f(b) - f(a) = (b - a)f'(c). Since f'(c) = (1- A)C- A < (1 - A)a- A , the above equation gives bl- - al- < (b - a)(1- A)a-\ i.e. G f < G ) A + 1- A, and the above inequality becomes equality if a = b. Simplification of the last inequality gives the arithmetic-geometric mean inequality (briefly, AM- GM inequality) aAb l - A < Aa + (1 - A)b, a, b E IR+ , with equality iff a = b. Further, this observation shows that f defined in Proposition 1.63 is positive for all t > 0, and for t :j:. 1 whenever we let t = a/b. Here, we also include a direct proof of AG mean inequality. The func- tion f : IR+  IR defined by f(x) = log x is concave downwards (since f'(x) = l/x which is decreasing and f"(x) = -1/x2 < 0). Therefore, by Proposition 1.52, if (xo, Yo) and (Xl, YI) are on this graph, then a general point (x, y) lying on the chord joining (xo, Yo) and (Xl, YI) will be below it, Le. y < log x. In other words, for A E [0,1], we have Alogx + (1 - A) logy < 10g(Ax + (1 - A)Y), x, Y E IR+; 
1.9. Exercises 53 Taking exponential on both sides and using the fact that exp(x) is increas- ing, it follows that (see also (1.56)) exp[Alogx + (1 - A) logy] < AX + (1 - A)Y. Since exp[A log x + (1 - A) log y] = exp[A log x] . exp[(1 - A) log y] = x A yl-A, this inequality implies that (1.67) x A yl-A < AX + (1 - A)y, X, Y E IR+ , with equality iff x = y. If we choose x = t A - 1 a and y = tAb, t > 0, in (1.67), then we find that x A yl-A = a A b 1 - A and, therefore, (1.67) is equivalent to (1.68 ) a A b 1 - A < At A - 1 a + (1 - A)tAb, for every t > 0 and this inequality becomes equality iff t = a/b. Note that t = 1 is (1.67) (see also (1.56)). . 1.9 Exercises 1.69. Determine whether the following statement are true or false. Justify your answer. (a) There exists a function f : IR  IR which is continuous at infinitely many points but between every two points there is a point of discon- tinuity and vice versa. (b) The function f(x) = (1 + l/x)X is strictly increasing and lim f(x) = e. x-+oo (c) The function f : IR  {O, I} defined by f(x) = {  if x is rational if x is irrational is not one-to-one but onto. (d) The function f : IR  IR, x I-t x 3 is one-to-one and onto. (e) The sine function f : IR  IR, x I-t sin x, is not one-to-one, but the restriction 9 : [-1r/2, 1r/2]  IR, x I-t sinx, is one-to-one and onto. (f) Sum of two convex functions in (a, b) is convex in (a, b). (g) If a sequence {l/Jn (x)} of convex functions in (a, b) converges to l/J( x) for each x E (a, b), then l/J is convex in (a, b). 
54 Chapter 1: Analysis and Linear Algebra (h) The function 1 is convex on (a, b) iff 1 is continuous and 1 ( Xl + X 2 ) < I(X1) + I(X2) ( ) 2 - 2 ' Xl, X2 E a, b · (i) If a continuous function 1 : (0, 1)  IR is convex on (0,1), then 1(1/2) < fo1 I(t) dt. (j) For 0 < p < 1, the inequality 1- xP < (1- x)P holds for all X E (0, 1). (k) For x, a E (0,1), we have the inequality x Q (I-x)l-Q < a Q (I-a)l-Q. Note: In the special case, a = l/p and x = qa P /(qaP+pb q ), we obtain the alternate proof of Young's inequality in p. 69 (see Proposition 1.63 ) a P b q ab < - + -, a, b > 0, 1 < p, q < 00, p-1 + q-l = 1. p q (I) The set Q forms a vector space provided that the scalar filed is Q itself. (m) The set of all solutions of the differential equation tFy dx 2 + 4y = 0 forms a vector space over III (n) The set of all complex sequences {zn} forms a vector space over C under the coordinatewise operations: Z + W = {Zk + Wk}kl, AZ = {AZk}kl, for A E C, where Z = {Zk}kl and W = {Wk}kl are two sequences in C and A E C. ( 0) Consider the two nonstandard operations of addition and scalar mul- tiplication (denoted by EB and 0 respectively) on JR2 over the field IR defined by (X1,X2) EB (Y1,Y2) = (Xl + Y1 + a,X2 + Y2 + b), A 0 (X1,X2) = (AX1 + Aa - a, AX2 + Ab - b) for A E JR, where a, b are two fixed real numbers such that at least one of a, b is a nonzero real number. Then the whole plane ]R2 becomes a vector space over the field III (Note that if a = b = 0, then the above definition describes the Euclidean space). (p) The set of all real valued even (odd respectively) functions in x E ]R forms a subs pace of the set of all real valued functions in x over the field IR. 
1.9. Exercises 55 (q) The set of all real sequences {x n } forms a vector space over the real IR but not over C under the coordinatewise operations as above. (r) For x = (Xl, X2), Y = (Y1, Y2) E IR 2 and a E IR, if we define the addition and the scalar multiplication by x+y = (X1Y1,X2Y2) a · x = (ax, ax), then IR2 is not a vector pace over IR. (s) If Pn(z) is the vector space of complex valued polynomials, then the transformation T : Pn(z)  Pn(z), p(z) t--+ p'(z), is linear. (t) The set {(I, 0), (i, O)} in C2 is linearly independent over the field IR but not over the field C. (u) Each set of n + 1 or more vectors in the vector space IRn over IR is necessarily linearly dependent. ( v) IT S = {V1, V2, . . . , v n } is a linearly independent set in a vector space V, then so does the set Sl = {V2 - a2V1,V3 - a3V1,..' ,V n - a n V 1}, where a2, a3, . . . , an are some scalars. . (w) If T E L(V) is not one-to-one, then there exists a nonzero S E L(V) such that T S = O. (x) If T E L(V) is not onto, then there exists a nonzero S E L(V) such that ST = O. (y) span {x + 1,x + 2,x 2 -I} = span {I +x 2 ,x 2 - x, 3 - 2x} = P2(X) and P2(IR) = span {I + x + x 2 , 1 + x, 2 - 3x}. (z) If a and b are fixed nonzero real numbers, then the mapping T : P2(X)  P3(X), p(x) t--+ ap'(x) + b fox p(t) dt, is one-to-one but not onto. 1.70. If T : en  en is such that R T = en, then show by a direct method that NT = {OJ. 1. 71. Construct an example of your own for a vector space V to have the following properties: (1) Linear map T : V  V which is neither one-to-one nor onto (2) Linear map T : V -+ V which is not one-to-one but onto (3) Linear map T : V  V which is one-to-one but not onto (4) Linear map T : V  V which is one-to-one as well as onto. 1.72. Determine the truth of the following statements with justification (see Examples 1.13): (1) f: IR  IR, x t--+ x 2 , is neither one-to-one nor onto. 
56 Chapter 1: Analysis and Linear Algebra (2) f: IR  IR+ , x I-t x 2 , is not one-to-one but onto. (3) I: IR+  IR, x I-t z2, is one-to-one but not onto. ( 4) I: IR+  IR+ , x I-t z2, is one-to-one and onto. 1.73. Construct an example of your own for a function 1 on ]R to have the following properties: (1) 1 is continuous at the irrational points but discontinuous at the ra- tional points (2) 1 is differentiable, but the derivative is not continuous (3) 1 is n-times differentiable, but not (n + I)-times (4) 1 is everywhere continuous and nowhere differentiable on III 1.74. Let V be the set of all polynomials in x E IR with real coefficients, and let T, S be two mappings in L(V) defined by T(P(x)) = p'(x) and S(P(x)) = 1 111 p(t) dt. Prove that both T and S are linear on V, dim NT > 0 and T S = Iv but ST:/;Iv. Note: The space V is not a finite dimensional vector space. 1.75. Is the map T: Mnxn(IR)  Mnxn(IR), A I-t -(A - At), linear? H so find the null space and the range space of T and their respective dimensions. 1.76. Show that the product Ig of two positive real valued functions I, 9 defined on an open interval (a, b) is decreasing (increasing, convex re- spectively) whenever I, 9 are both decreasing (increasing, convex respec- tively) . 1.77. Give a geometric proof of the Young's inequality (see also p. 69): Let y = I(x) be a real-valued continuous, unbounded, and strictly increas- ing function on [0,00) with 1 (0) = O. If x = g(y) is the inverse of I, then for a, b > 0, ab < 1 B f(x) dx + 1 b g(y) dy. Equality holds iff 1 (a) = b. 1.78. Show that the intersection of an arbitrary family of convex,sets in IRn is convex. 
1.9. Exercises 57 1.79. If Xl,X2,... ,X n > 0 and Pl,P2,... ,Pn > 1 with EZ=l p;l = 1, then show that n n Pk II Xk < L  k=l k=l Pk I with equality iff xii = X2 = · · · = xn . 
Chapter 2 Concepts in Metric Spaces The plane has both algebraic(vector space) and geometric(distance) prop- erties. In the earlier chapter we have discussed some of the algebraic and geometric properties of the complex field or plane as the case may be. In this chapter, first we discuss the notion of metric spaces with several ex- amples of metric spaces and study some of the topological properties of metric spaces as these spaces are important in both theoretical and applied fields like functional analysis, numerical analysis, physics, economics and engineering. 2.1 Metric Spaces: Definitions and Examples We will realize that some of the results and their proofs on metric spaces are mainly the translation of the proof from the Euclidean setting to the new setting of metric spaces. For example, as we have studied in the first course on real and complex analysis, the functionS x x X  JR, (z, w) I-t Iz - wi, X = IR or C has the following properties: (i) Iz - wi = 0 <==> z = w, Le. the distance between the two points is zero iff the two points coincide; (ii) Iz - wi = Iw - zl, i.e. the distance from the point z to the point w is same as the distance from the point w to the point z; (ill) Iz-(I < Iz-wl+lw-(I, where z, w, ( E X, Le. in Euclidean geometry it says that the length of one side of a triangle (with vertices z, w, () cannot exceed the sum of lengths of the other two sides. 8The symbol 'x' denotes the Cartesian product of two sets. 
60 Chapter 2: Concepts in Metric Spaces Based on these three properties, we study more general spaces and functions defined on them. In particular, these observations motivate us to introduce the following: 2.1. Definition. Let X be a nonempty set and let d(.,.) be a map- ping/function from X x X to IR, d : X x X  IR, satisfying the following conditions for all x, y and z in X: (Ml) d(x, y) = 0 {::::::} x = y [Identity] (M2) d(x,y) = d(y,x) [Symmetry] (M3) d(x,y) < d(x,z) +d(z,y). [Triangle inequality] Then d is called a metric or a distance function 9 on X. The set X together with a metric, denoted by (X, d), is called a metric space. The conditions (Ml)-(M3) are usually called the metric axioms. By a pseudo-metric d on X we mean that the function d that satisfies the axioms (Ml)-(M3) except that d(x, y) = 0 does not imply x = y. In this case, the space (X, d) will be called pseudo-metric space. As we shall see in several examples below, it is always possible to have different metrics defined on the same set X. Thus, it should be noted that when one refers to metric spaces, it is always the pair (X, d). However, we are usually concerned only with one metric at a time, and so we often talk of "the metric space X" especially when the specific metric d on the underlying set X is clearly indicated. It is customary to refer the members of a metric space as 'points'. By setting y = x in (M3) gives o = d(x, x) < 2 d(x, z), Le. d(x, z) > 0 for each x, z E X which shows that d is always nonnegative. In other words,. the distance between two points is nonnegative and the distance between a point and itself vanishes. Conversely, the only point at distance. 0 from x is x itself. From now onwards, we can consider the distance function as a mapping from X x X into IRt. By the triangle inequality (M3), we note that every metric space (X, d) satisfies the inequality (2.2) Id(x, y) - d(x, z)1 < d(y, z) for each x, y, z E X. Most often to verify whether a given function satisfies all the axioms of a metric, it is nontrivial to check the triangle inequality as we do for instance in Examples 2.5-2.38, because the other two axioms can be easily verified. We say that the metric space (X, d) is a bounded metric space if there exists M > 0 such that d(x,y) < M for all x,y E X. Otherwise (X,d) is said to be unbounded. 9The real number d( x, y) is called the distance between the two points x E X and y E x. 
2.1. Metric Spaces: Definitions and Examples 61 2.3. Definition. (Diameter and distance between sets) Let (X, d) be a metric space and A, B be two nonempty subsets of X. The diameter 10 of the subset A is defined as d(A) := sup{d(x, y) : x, YEA}. The number d(A,B), called the distance between A and B, is defined by d(A, B) := inf{ d(x, y) : x E A, y E B}. Here 'sup' and 'inf' denote respectively the usual 'least upper bound' and 'greatest lower bound' of a set of real numbers. Clearly, d(A, B) = d(B, A) because d(x, y) = d(y, x). If A and B are singleton sets {a} and {b} respectively, then d(A, B) reduces to d(a, b) which is the distance between the two points a and b. If there exist points a E A and b E B such that d(A, B) = d(a, b), then we say that the distance from A to B is attained. If d(A) < 00, then we say that A is bounded. Otherwise it is unbounded and we say that it has infinite diameter. In the metric space (IR, d), if A = IR+ and B = IR- then d(A, B) = O. This example shows that d(A, B) = 0 does not necessarily imply that A and B have points in common. If Xo E X is fixed, then the distance from Xo to A is defined by d(xo, A) := inf{d(xo, y) : yEA}, see Figure 2.1. For the empty set 0, the usual convention is d(x,0) = 00, d(A,0) = d(0, A) = 00, d(0) = -00. We note that the diameter of A is the distance between the two most distant points of A, if such points exist. But, for example, if (IR, d) is the Euclidean metric space and A = (0,2] C IR then d(A) = 2 even though no two points of A have distance exactly 2. But there do exist points x, y E (0,2] with distance as, near as we please to 2 and further we also see that there exist no two points x, y E (0,2] such that the distance is greater than 2. 2.4. Metric subspace. Let (X, d) be a metric space, and Y C X be nonempty. We can define a distance function dy : Y x Y -+ IRt by dy(a, b) = d(a, b) for a, bEY. In other words, dy is the restriction of the metric d : X x X  IRt to Y x Y. Then it is trivial to see that dy is also a metric on Y. This metric is called the relative metric induced on Y by the metric d on X. We call (Y, dy) a metric subspace of (X, d). As usual, we refer Y as simply a subspace of X rather than (Y, dy) as a subspace of (X, d). lOSome authors use the notation diam A to denote the diameter of A. 
62 Chapter 2: Concepts in Metric Spaces ,... - ....... ", ./ A / '\ / \ I \ I xo I \ ./ \ I ./ "\ \ / / ./ / 't , / " ", ........ ."" - -- Figure 2.1: Description for the distance from a point to a set 2.5. Euclidean metric on C and IR. If X = C or IR, define d : , X X X  IRt by the usual Euclidean distance d(z,w) = Iz - wi, z,w E X. Then (X, d) is a metric space. We note that (X, d) is unbounded, while the space (X, p), where p(z, w) = min{ d(z, w), 2}, is seen to be a bounded metric space. 2.6. Discrete metric. The discrete metric (or trivial metric) on any nonempty set X is defined by d(x,y) = { ifx=y if x :F y x, Y E X. Note that we do not assume X to be finite. Clearly eacl\ of the following inequalities hold: O=d(x,x) < d(x,z)+d(z,x)=2d(x,z) withy=x, d(x,y) < d(x,y)+d(y,y)=d(x,y) withz=y, d(x,y) < d(x,x)+d(x,y)=d(x,y) withz=x. Further, for different values of x, y, z, the above inequalities give the tri- angle inequality (M3). Thus, every nonempty set X can be made into a 
2.1. Metric Spaces: Definitions and Examples 63 bounded metric space in a trivial way as above. Even though the discrete metric space is not too interesting  it stands, most often it is helpful in constructing counterexamples. The set of all rationals Q is a subspace of IR with respect to the Euclidean metric. But, if we treat Q as a discrete metric space then Q would not be a subspace of (IR, d), where d is the Euclidean metric. This is because the metric used on Q is different from Euclidean metric. If (R, d) is the discrete metric space and A = {4, 100} C IR, then d(A) = sup{d(4, 100),d(4, 4), d(100, 100)} = sup{l,O,O} = 1. \J 2. 'T. Metric on the extended set of N. For X  N := N U {oo} and f(x) = 1/x, we define d(x, y) = If(x) - f(y)1 f(x) f(y) o if x, yEN if x E N, y = 00 if x = 00, yEN if x = 00, y = 00. OJ Then (N, d) is a metric space which is clearly bounded. \J 2.8. Chordal metric on C. We recall a basic result from complex analysis, see [Po]. Stereographic projection determines a one-to-one corre- spondence between the unit sphere S of radius 1/2 with center at (0,0, 1/2) in 1R3 minus the north pole N = (0,0, 1), and the complex plane via the correspondence  + i1] z+-+ 1-( where ({, 1], () E S\{(O, 0, I)}, z E C with Rez Imz Izl 2  = (1 + I z I 2 ) ' 17 = (1 + IzI 2 ) ' (= 1 + Izl 2 · \J IT we define C = C U {60}, then we have the one-to-one correspondence \J between Sand C by mapping the point at infinity with the north pole (0,0, 1) and the points in the complex plane C with that of the points on the sphere S ininus the north pole (0,0,1), respectively. This idea of \J Riemann allows us to define a new function on C: for z, wEe we define (see ExampleS 2.8 and 2.96) Iz-wl (2.9) x(z, w) = V(l + Izl2)(l + Iw1 2 ) and x(z,oo) = lim X(z,w) = VI , w-+oo (1 + Iz12) 
64 Chapter 2: Concepts in Metric Spaces see [Ah, Po]. Clearly, X(z, w) < 1. Indeed, X(Z,w) < 1 <==} 1+lzI 2 1IwI 2 +2Re(z w ) > 0 <==} 11 + z w l 2 > O. Geometrically X( z, w) represents the Euclidean distance of the preimages of Z and w under the stereographic projection of the sphere of radius 1/2 with center at (0,0, 1/2) onto the complex plane, thought of as the xy-plane, from the point (0,0, 1). Observe that X (  , :, ) = X(z,z'), z,z' E t = c u {co}. A higher dimensional analog of this notion will be explained in detail later in this chapter, see pages 65 and 116. 2.10. Proposition. For a, b, c E C, we have the triangle inequality (2.11) x(a, c) < x(a, b) + X(b, c). Proof. Let a, b, c E C and let X be defined by (2.9). Clearly la - bl > 0 which is equivalent to 11 + a bl 2 < (1 + laI 2 )(1 + IbI 2 ). If we apply this inequality to the identity (a - c)(l + bb) = (a - b)(l + be) + (b - c)(l + ab), then we find that la - cl(l + Ib1 2 ) < la - bill + bcl + Ib - cl11 + abl < la - bl v 1 + Ibl 2 V 1 + Icl 2 + Ib - cl V 1 + lal 2 v 1 + Ibl 2 from which we immediat ely get t h e inequa lity (2.11), if we divide the last inequality by (1 + Ib1 2 ) V I + lal 2 V I + Ic1 2 . . Now, for X = C, let X(z,w) be defined by (2.9). Clearly (i) X(Zl, Z2) > 0; (ii) X(Zl, Z2) = 0 <==} Zl = Z2; (iii) X(Zl, Z2) = X(Z2, Zl); (iv) X(Zl, Z3) < X(Zl, Z2)+X(Z2, Z3); (In fact, the triangle inequality follows from Proposition 2.10.) (v) X(O, Zl) < X(O, Z2) provided IZll < IZ21 < 00; (vi) X(Zl,Z2) < IZl - z21 = d(Zl,Z2). 
2.1. Metric Spaces: Definitions and Examples 65 v In particular, we call X defined on C, the extended complex plane C U v {oo}, the chordal metric on C and X(z, w) is called the Chordal distance of z from w, see [Ah, Po]. This allows us to treat the point at 00 like any other point. 2.12. Convergence of sequences. Let us briefly discuss the familiar concepts of convergence of sequences and series in IR or C. In the real or complex number system, when we have a sequence {zn} in C or IR con- verging to a limit z, we write IZn - zl  0 as n  00. Here IZn - zl is the Euclidean distance between Zn and z. Now, we generalize this notion to the analysis of metric spaces. A sequence {xn} of elements of a metric space X is said to be convergent if there exists a point x E X such that d( X n , x)  0 as n  00, and the point x (such point should be unique as we see below) is called the limit point of the sequence {x n }. Common notation expressing that the sequence {xn} converges to x are: x = lim X n , X n  x. n-+oo If a confusion arises, we shall usually say 'xn  x in the metric d of X' rather than just 'xn  x'. In terms of €-N notation, we say X n  x if for a given € > 0 there exists a positive integer N = N(€) such that d(xn, x) < € for all n > N. Geometrically, this means that X n E Bd(X; €) for all n > N, where Bd(X; 6) := {y EX: d(y, x) < 6}. Thus, an equivalent definitions of convergence of a sequence may be formu- lated as 2.13. Proposition. Let {xn} be a sequence in a metric space (X, d). Then X n  x in (X, d) iff d(xn, x)  0 in the Euclidean space IR. In addition to the above Proposition, the following will be useful 2.14. Proposition. If €n are nonnegative real numbers such that €n  0 as n  00 in the Euclidean space IR and d(xn,x) < €n for all (sufliciently large) n, then X n  x in (X,d). Proof. By definition, €n  0 in IR iff for arbitrary € > 0, there exists a positive integer N = N(€) such that €n < € for all n > N. But then, d(xn,x) < €n < € for all (sufficiently large) n, which means that X n  x in (X,d). . 
66 Chapter 2: Concepts in Metric Spaces 2.15. Uniqueness of the limit. H {x n} is a sequence in (X, d) such that d(xn,x)  0 and d(xn,Y)  0 for some x and Y in X, then from (M3) we see that d(x,y) < d(x,x n ) + d(xn,y)  0 so that d(x, y) = 0, which gives x = y. Thus, the limit is unique. Here is an important result about subsequences and is left as an exercise for the readers. 2.16. Proposition. If {xn} is a sequence in a metric space (X, d) that converges to x in (X, d), then each subsequence {x n ,.} also converges to x in (X, d). One way of proving that a sequence fail to converge is to find two con- vergent subsequences but with different limits. For example, the sequence { III' 1 I } 1, 1 - 2' 3 ' 1 - 4 '.." 2k _ 1 ' 1 - 2k "'. in IR (with usual metric) does not converge, since it contains two subse- quences {xn} with X n = 1/(2n -1), and {x} with x = l-l/(2n), which converge to 0 and I, respectively. 2.17. Convergence of series. Let {zn} be a sequence of complex numbers. Then, the sequence {Sn}nl of partial sums of {zn} is defined by n Sn = LZk. k=l H Sn  Z as n  00, then the original sequence is said to constitute the terms of the convergent series. In other words, if Sn  Z then we say that the series E  1 Zk converges to Z or has a sum Z; that is, 00 Z = lim Sn =  Zn n-+oo  n=l where the limit Z is called the sum of the series, and can be easily seen to be unique. H the sequence {sn} does not converge, then we say that the series is divergent. By the triangle inequality ISnl < IZll + I Z 21 + · · · + IZnl := Un. If {Un} converges, Le. if E  llzkl < 00, then the series E  1 Zk is said to be absolutely convergent. Absolute con'vergence implies convergence, but not the converse as the example E  1 (-I)n /n points out.' 
2.2. Holder and Minkowski Inequalities 67 We shall present some further important examples of metric spaces that arise naturally. 2.18. The metric space of all sequences of complex numbers. Let X be the set of all infinite sequences of complex numbers not necessarily convergent, not even bounded. Let {k n } be an arbitrary fixed sequence of positive real numbers such that the sum E  1 k n converges (For instance, k n = 2- n , 3- n or I/n! etc.). For z = {Zn}n>l and w = {W n }n>l in X, - - define d by (  IZn -wnl d z, w) = L..- k n 1 I _ I' n=l + Zn W n First, we note that the series defining d(z, w) converges, since k n IZn - W n I < k n 1+l z n- w nl and that E  1 k n converges. Secondly, the triangle inequality (M3) IS immediate from the inequality (1.60). Further, since 00 d(z,w) < E k n < 00, n=l the space (X, d) is bounded. This metric is called Frechet metric for X. 2.19. Product metric spaces. Given a set of metric spaces (Xk, dk), k = 1, 2, . . . , n, we define the Cartesian product (or simply product) X of the metric spaces Xk, k = 1,2,. . . , n, by X = {(X1,X2,...,X n ): Xk E Xk, k = 1,2,...,n}. If we define the function d oo by the formula doo(x,y) = max dk(Xk,Yk) lkn then (X, d oo ) becomes a metric space and this space is usually denoted by Xl X X 2 x... X Xn. Can you define some other metric d on X x X? 2.2 Holder and Minkowski Inequalities In this section, we discuss several important inequalities and their conse- quences. 2.20. Definition. If p > 1, then we say that a number q is said to be conjugate index of p if p-1 + q-1 = 1, q = 00, q = 1, for 1 < p < 00 for p = 1 for p = 00. 
68 Chapter 2: Concepts in Metric Spaces q !.+!.=1 _ __ __ _po __ __ (2, 2) p q I I I I I I I I I I I I ----------------------------------------------- I I I I I I I I o p Figure 2.2: The graph of p-l + q-l = 1 for 1 < p, q < 00 We note that, for 1 < p, q < 00, the first condition in the above combi- nation can be put in anyone of the form p q (p - 1) (q - 1) = 1, q = l ' P = 1 or p + q = pq, p- q- and this simple observation is often useful while dealing with the triangle inequality in the discussion on metric spaces. Clearly 2 is the only number which has its own conjugate; 1 and 00 are considered to be conjugate index, see Figure 2.2. ' We next look at three specific lemmas that are especially important and interesting. 2.21. Lemma. (Holder's inequalities) Let 1 < p, q < 00 and p-l + q-l = 1. Suppose that Zl, Z2, . . ., Wl, W2, . . . , are complex numbers. Then we have the following statements: n ( n ) l/p ( n ) l/q (i)  IZkWkl <  IZkl P  IWklq [Finite sums] with equality iff all Zk are 0, or there exists a constant M such that IWklq = MlzklP for all k (p = q = 2 is known as Cauchy-Schwarz inequality, see also Corollary 6.42. 00 ( 00 ) l/P ( 00 ) l/q (ii)  IZkWkl <  IZklP  IWklq [Infinite sums] (Hi) L lu(s)v(s)1 ds < (L lu(s)IP dS) lip (L Iv(sW dS) l/q [Integrals] Here {} is a bounded closed interval in IR (although the inequality 
2.2. Holder and Minkowski Inequalities 69 continues to hold in a general setting) and u(s) and v(s) are Lebesgue measurable functions defined on n. Proof. (i) Let a, b > 0 and consider f (t) = 1 - A + At - t A . Then, from Proposition 1.63 with A = 1 I p and t = a P Ib q , it follows immediately that a P b q (2.22) ab < - + - - P q with equality iff a P = b q , see also Remark 1.66. Inequality (2.22) is known as Young's inequality and for a simple extension of (2.22), we refer to Exercise 1.79. Now we consider ( n ) l/p 0: =  IZklP , ( n ) l/q /3 =  IWkl q · If a{3 = 0, then either a = 0 or {3 = 0 or both. In either case the result is trivial as both sides of (i) equal to zero. If a{3 > 0, we let IZkl b = IWkl (k ) a=, /3 =1,2,...,n so that the inequality (2.22) becomes IZkllwkl < IZklP + IWklq {3 /3 (k = 1,2,...,n) a - paP q q with equality iff {3qlzkl P = aPlwk Iq, for each k = 1,2, . . . , n. From the last inequality we see that n { 1 n 1 n } ( 1 1 ) E IZkWkl < 0:/3 o:P E IZkl P + {3 q E IWkl q = a{3 - + - = a{3 k=l p k=l q k=l P q with equality iff {3qlzkl P = aPlwklq, i.e IWklq = MlzklP, for each k = 1,2,...,n, where M = {3qla P . The conclusion of part (i) of the lemma follows at once. (ii) Follows from (i). Without loss of generality we assume that 00 00 E IZkl P < 00 and E IWkl q < 00. k=l k=l Note that by (i) every partial sum EZ=l IZkWkl is bounded so that the series E  1 IZkWkl converges and the conclusion follows from (i) if we let n  00. ,(iii) Again, without loss of generality, we assume that In lul P ds < 00 and In Ivl q ds < 00. Since p-l + q-l = 1, the inequality (1.68) (with a P for a, bP for b and A = lip) may be rewritten as (2.23) ab < 1 r 1 / q a P +  t 1 / P b Q -p q 
70 Chapter 2: Concepts in Metric Spaces so that (2.24 ) inf[! rl/qa P + ! tl/Pb q ] = ab. t>O p q Therefore, if we choose a = lu(s)1 and b = Iv(s)1 in (2.23), then we get lu(s)v(s)1 < !rl/qlu(s)IP + !tl/Plv(sW p q which, by integrating both sides, yields k1u(s)v(s)1 ds <  r l/q (k'u(s)IP dS) +  t l/p (k'V(SW dS) · Now, taking infimum over all t > 0 and using (2.24) with a P = k1u(s)IP ds, b q = L Iv(sW ds, we find that k1u(s)v(s)' ds < (k'U(s)IP ds riP (k'v(SW ds r lq which proves (iii). . 2.25. Remark. The inequality (2.22) follows from Exercise 1.79 if we assume n = 2, al = a, a2 = b, Pl = P, P2 = q in Exercise 1.79. . 2.26. Lemma. (Minkowski's inequalities) Let 1 < p < 00. Sup- pose that Zl, Z2, . . ., Wl, W2, . . . , are complex numbers. Then we have the following statements: ( n ) IIp ( n ) IIp ( n ) IIp (i)  IZk ::I: wkl P <  IZkl P +  IWklP [Finite sums] with equality iff either all Zk are 0, or IWkl = Mlzkl for all k and for some M > 0 or else P = 1 and, for each k, either Zk = 0 or Wk = MkZk for some M k > O. ( 00 ) IIp ( 00 ) IIp ( 00 ) IIp (ii)  IZk ::I: wkl P <  IZkl P +  IWkl P [Inf ini te sums] (Hi) (L I(u::l: v)(s)IP ds riP < (L lu(s)IP ds riP + (L Iv(s)IP ds riP [Integrals] 
2.2. Holder and Minkowski Inequalities 71 Hereu and v are Lebesgue measurable functions denned on n. lEO < p < 1, then the inequality signs in (i), (ii) and (iii) are reversed. Proof. (i) Applying Holder's inequality (Lemma 2.21(i)) for p > 1, we have n ( n ) IIp ( n ) l/q  IZkl{(lzkl + IWkDP-1} <  IZklP  (IZkl + IWkD(p-l)q ( n ) IIp ( n ) l/q =  IZkl P  (IZkl + IWkDP , since (p - l)q = p. From the equality part of Lemma 2.21(i), we note that we have the equality in the above inequality if {(IZkl + IWkDP-l}q = (IZkl + IWkDP = MPlzkl P for some M, Le. there exists M l = M - 1 > 0 such that IWkl = Mllzkl. Similar inequality holds for Wk: n ( n ) IIp ( n ) l/q  IWklHlzkl + IWkD P - 1 } <  IWkl P  (IZkl + IWkD P · Therefore, writing (IZkl + IWkD P = IZkl(lzkl + IWkDp-l + IWkl(lzkl + IWkD p - l and then applying the last two inequalities we obtain that ( n ) l-l/q ( n ) IIp ( n ) IIp  (IZkl + IWkDP <  IZkl P +  IWkl P · Since 1 - l/q = l/p, the desired inequality is immediate from the last inequality if we use the triangle inequality Iz + wi < Izi + Iwi. Thus we complete the proof of (i). (ii) Without loss of generality, we assume that E  1 IZklP < 00 and E  llwkl P < 00. By (i), we see that ( n ) IIp ( 00 ) IIp ( 00 ) IIp  IZk ::I: wkl P <  IZkl P +  IWklP for every n, and since the two series on the right converge we obtain that the series on the left converges. Therefore (ii) follows. 
72 Chapter 2: Concepts in Metric Spaces (Hi) Again, without loss of generality, we assume that In lu(s)IP ds < 00 and In Iv(s)IP ds < 00. From (1.64), we see that for all t E (0,1), L lu + viP ds < L (Iul + Ivl)P ds < t l - p L lul P ds + (1 - t)l-p L Ivl P ds. Taking infimum over all t E (0,1) and then using (1.65), we get L I(u + v)(s)lP ds < [ (L lu(s)IP ds rIP + (L Iv(s)IP ds r/Pr which proves (Hi). . 2.27. Corollary. Let 1 < p, q < 00 and p-l + q-l = 1. Then all the zeros of the polynomial P(z) = EZ=o akzk (ak E C, an :j:. 0) lie in the closed disc  R = {z E C: Izi < R}, where { ( a a n k P ) q/P } l/q . R := 1 + I: k=O Proof. For Izl > 1, we have n-l k Izqln - 1 Izqln t; Iz I q = Izlq - 1 < Izlq - 1 · Using this and the Holder inequality (see Lemma 2.21(1)) we find that n-l ( n-l ) l/p ( n-l ) l/q ( n-l ) l/p  lakzkl <  lakl P  Izklq <  lakl P (IZlqln1)l/q ' Now for Izl > R > 1, the last inequality gives n-l IP(z)1 > lanznl - E lakllzl k , by the triangle inequality (1.59), k=O > lanllzl n [1 - la1 (  laklP) IIp CIZlq  1)l/q ) ] in which the square bracketed term in the last inequality is nonnegative provided { a a n k P ) q/P } l/q . Izi > R = 1 + (  k=O 
2.2. Holder and Minkowski Inequalities 73 Thus all the zeros of P(z) lie in, the closed disc  R = {z E C: Izi < R}. . The next lemma is straightforward consequence of the triangle inequal- ity (1.59). 2.28. Lemma. (i) If ZI, Z2,..., WI, W2,..., (1, (2,... are complex numbers, then 11 . sup IZk - wkl < sup IZk - (kl + sup I(k - wkl, k1 k1 k1 and max IZk - wkl < max IZk - (kl + max I(k - wkl. l:Sk:Sn l:Sk:Sn l:Sk:Sn (ii) If f{t), g{t), h{t) are either arbitrary continuous complex valued func- tions on a closed interval [a, b], or arbitrary bounded functions on [a, b], then sup If(t) - g(t)1 < sup If(t) - h(t)1 + sup Ih(t) - g(t)l. tE[a,b] tE[a,b] tE[a,b] (Note that this inequality continues to hold if we replace the closed interval by an arbitrary compact set X.) (iii) If ZI, Z2 are any two complex numbers, then for a fixed real number p, 1 < p < 00, we have I Z I :f: z21 P < 2Pmax{lzIIP, IZ2IP} < 2 P (l z II P + IZ2IP). Proof. (i) We prove the first part and the second part follows by similar procedure. For each k > 1, we have IZk - wkl < IZk - (kl + I(k - wkl, by the triangle inequality, < sup IZk - (kl + sup I(k - wkl k1 k1 and since this holds for every k, the conclusion follows. (ii) For the proof of the first part, as in Case (i), we have If{t) - g(t)1 < If(t) - h(t)1 + Ih(t) - g(t)1 < sup If{x) - h(x)1 + sup Ih(x) - g(x)1 xE[a,b] xE[a,b] which shows that f(t) - g(t) is bounded on [a, b]. Therefore, taking the supremum on the left we obtain the required inequality. The second part follows similarly. 11 The least upper bound of a nonempty set S of real numbers is denoted by sup S. 
74 Chapter 2: Concepts in Metric Spaces (iii) Without loss of generality we can assume that IZll < IZ21. Then, the triangle inequality (1.59) gives I Z I ::I: z21 < I Z ll + I Z 21 < 21 z 21 and the required inequality follows by raising to pth power. - A more refined form of Lemma 2.28(iii) is contained in the following (2.30) 2.29. Lemma. For a, b > 0, we have { 2p-l(aP + bP) (a + b)P < - a P + 1JP ifl < p<oo if 0 < p < 1. Proof. Without loss of generality, we can assume 0 < a < b so that if we divide the inequality (2.30) by a P , we find that (2.30) is equivalent to show that (1 + x)P { 2P-l g(x) := < 1 + x P - 1 ifl < p<oo if 0 < p < 1 for x > 1. Since , _ p(l + x)p-l (1 - x P - 1 ) 9 (x) - (1 + X p )2 < 0, for x > 1, the function 9 is decreasing for x E (1, 00) if p > 1, and it is increasing if o < p < 1. Thus g(x) < g(l) = 2 P - 1 if p > 1 and since g(x)  1 as x  00, g(x) < 1 for 0 < p < 1. This reasoning completes the proof. _ 2.31. Remark. From the proof of Lemma 2.29, we also see that if o < p < 1 then (a + b)P > 2P-l(a P + 11') for a, b > o. Indeed, for an alternate proof of Lemma 2.29 for p > 1, we consider the function f(x) = Ixl P for x E III Then { pl x l p-l X > 0 } f ' ( x ) = ' - = X l x l p - 2 > 1 I IP -l 0 P , p- -p x , x < (meaning 0 when x = 0). Clearly, f'(x) is strictly increasing on III In particular f is convex on (0,00) and therefore, for x, y E (0,00) we must have f((x + y)/2) < (f(x) + f(y))/2 which gives the desired inequality. Letting y = 1 in the last inequality we deduce that ((1 + x)/2)P < (1 + x P )/2, Le. (1 + x)P < 2 P - 1 (1 + x P ), p > 1 and the conclusion follows. . 
2.3. Metric Spaces IP(n), IP and C[a, b] 75 2.3 Metric spaces lP(n), lP and C[a, b] 2.32. The metric spaces lP(n) and lP. If X =  (IF = C or IR) and if 1 < p < 00 is fixed, then (X, d p ) is a metric space with the metric d p defined by (2.33 ) dp(z, w) = ( n ) l/p  IZk - wkl P max IZk - wkl lkn ifl < p<oo if p = 00, wherez = (Zl,Z2,...,Zn), W = (W1,W2,...,W n ) E JF'1. The space X defined with this metric is usually denoted by lP(n) and is actually the space r with the metric d p (.,.) defined by (2.33). We note that the triangle inequality for 1 < p < 00 follows from Minkowski inequality (see Lemma 2.26(i)) and for p = 00, this is immediate from Lemma 2.28(i). Even though Minkowski inequality fails when p < 1, lP(n) is still a vector space for 0 < p < 1 because z + w, AZ E lP(n) whenever z, w E lP(n) and A is a constant. We remark that the natural metric produced by the case p = 00 is known as maximum metric on X. When p = 2 and X = IRn , the corresponding metric ( n ) 1/2 d 2 (x,y) =  IXk - Ykl 2 is called Euclidean metric on IRn. Here x = (Xl, X2, · · . , x n ) and Y = (Y1, Y2, . . . , Yn) E IR n . It is important to find relationship between the metrics dp(z, w) on x=r: doo(z,w) < dp(z,w) < n 1 / p d oo (z,w) for p > 1. First, we shall prove the following inequalities: doo(z,w) < d 2 (z,w) < y'ndoo(z,w) doo(z,w) < d 1 (z,w) < ndoo(z,w) d2(Z,W) < d 1 (z,w) < y'nd 2 (z,w) for n > 1 and for z, w E IP(n). First we note that n max IZk - wkl <  IZk - wkl := d 1 (z,w) < n max IZk - wkl l<k<n - - l<k<n - - k=l - - which is equivalent to doo(z,w) < d 1 (z,w) < ndoo(z,w) 
76 Chapter 2: Concepts in Metric Spaces and other two inequalities may be proved similarly. In fact, since d p (z , w) > I Zk - W k I for each k = 1, 2, . . . , n, we have dp(z,w) > max IZk - wkl = doo(z,w). - lkn On the other hand, for p > 1, we have n [dp(z,w)]P = E IZk - wkl P < n[ max IZk - wkl]P = n[doo(z,w)]P l<k<n . k=l - - so that dp(z,w) < n 1 / p d oo (z,w). Thus, we have (2.34 ) doo(z,w) < dp(z,w) < n 1 / p d oo (z,w) for p > 1. Now, since lim p -+ oo n 1 / p = 1, passing the limit p  00 in the last inequality, we see that doo(z,w) = lim dp(z,w), z,w E r, p-+oo and because of this reasoning, we use the index 00 for denoting the space defined in (2.33) by loo(n). In fact the last equality (2.34) also follows from a simple observation that for a > b > 0 and p > 1 a < (a P + lJP)l/P = a[1 + (b/a)P]l/P < a2 1 / p  a as p  00. In particular, we have ( n ) l/p max IZkl = lim  IZklP , 1 <k<n p-+oo L....J - - k=l Z E lP(n) and we remark that the relationship between different metrics plays a very important role in numerical analysis. The spaces lP and 1 00 , for infinite sequences are analogous to lP(n) and loo(n), respectively. In fact, this is a natural analog of the forego- ing Example 1.32 for a set of oo-tuples {Zk}k>l that are pth summable: E  11 z kl P < 00 (IT p = 2 this condition is called square summable). Let Z = {Zk}k>l and w = {Wk}k>l be two sequences in IF. For 1 < p < 00, we define the sequence space lP (pronounced "little ell p") by12 {z:  Iz"IP < 00 } if 1 < p < 00, { z: sup IZkl < oo } if p = 00. lk<oo lP = 121f we consider real sequences we get real space lP, otherwise the complex space lP, 1 < p < 00. However, we do use the same notation irrespective of whether we deal with complex space or real space. 
2.3. Metric Spaces lP(n), lP and C[a, b] 77 (1 2 is called the space of square summable sequences whereas II is called the space of absolutely summable sequences.) In the sequence spaces oper- ations for addition and scalar multiplication are defined, in a natural way, coordinatewise (as in Examples 1.32): Z + W = {Zk + Wk}kl, AZ = {AZk}k>l, for A E IF. By Lemma 2.28(iii), we note that 00 00 L IZk :f: wkl P < 2 P L(lzkl P + IWkI P ), for 1 < p < 00 k=l k=l and therefore for 1 < p < 00 we have Z, W E lP => Z + W E lP. For p = 00, by Lemma 2.28(i), Z, W E ' 00 => Z + W E ' 00 and AZ E ' 00 for A E 1F. Thus lP, 1 < p < 00, becomes a vector space with respect to the above rules for addition and scalar multiplication. Obviously, one cannot find a finite number of elements in lP which can span the space lP, and therefore lP is an infinite dimensional vector space for each 1 < p < 00. We shall meet this example on and often. Let X = lP, 1 < p < 00, where lP is defined as above. For Z = {Zk}k>l and W = {Wk}k>l in X, - - define d p (z , w) by dp(z,w) = ( 00 ) IIp  IZA: - wkl P sup IZk - wkl lk<oo if1 < p<00 if p = 00. Then (X, d) is an unbounded metric space because d(nz, nw) = nd(z, w) for every n > o. We observe that the triangle inequality (M3) for 1 < p < 00 and p = 00 follows from Lemmas 2.28(i) and 2.28(iii), respectively. 2.35. Remark. For X = rand 0 < p < 1, let d p be defined by (2.33). That is ( n ) IIp dp(z,w) =  IZk - wkl P 
78 Chapter 2: Concepts in Metric Spaces where Z = (Zl, Z2,..., zn), W = (WI, W2,..., w n ) E r. Then d p does not define a metric on X. In fact, for z=(I,I,O,...,O)j (=(0,1,0,...,0) and w={O,O,O,...,O) we have dp(z, w) = 21/p, dp(z, () = 1 = d p ((, w) so that 2 1 / P =d p (z,w) >dp(z,()+dp{(,w) =2 which shows that the triangle inequality is not satisfied. . 2.36. Remark. Since this definition of d p for 0 < p < 1 does not satisfy the triangle inequality for the space IP, it follows that this definition does not define a metric on IP for 0 < p < 1. However, in this case, if we define a map d : lP x IP  IRt by a new formula 00 d(z,w) = E IZk - wkl P , 0 < p < 1, k=l then (IP, d) becomes a metric space (show!). . 2.37. The metric space of bounded functions. For a nonempty set X, the space B(X) of bounded, but not necessarily continuous, scalar- valued functions on X is defined by B(X) = {f : there exists K > 0 such that If(t)1 < K for t EX}. This is another example of function spaces. First we note that B{X) is a real (or complex) vector space under pointwise addition and multiplication by a scalar: (f + g).(t) = f(t) + g(t) and ()..f)(t) = )..f{t), for t E X. IT f,g E B{X), we can introduce a metric on the space B{X) by the formula 13 doo{f, g) = sup If(t) - g(t)l, tEX for each pair f, 9 E B{X). Lemma 2.28(ii) is precisely the triangle inequal- ity (M3) for B{X) and the verification for the other axioms of the metric space is trivial. The space B(X) is then a metric space. 2.38. The metric space of continuous functions. The concept of continuous functions defined on an arbitrary metric space will be introduced 13We use the term "max" only when the supremum is actually attained. This is in fact the case in Example 2.38 but not necessarily in Example 2.37. 
2.3. Metric Spaces IP(n), lP and C[a, b] 79 s doc (f, g) s = g(t) s = f(t) / o a b t Figure 2.3: Greatest width d oo (I, g) later in Section 2.5. Let [a, b] c IR be a (nonempty) closed and bounded interval. Now we introduce a metric on the function space Cc[a, b]. Let the distance between two members of f, 9 E Cc[a, b] be maximal distance between their values. Note that in this definition If(t) - g(t)1 is continuous on the closed interval [a, b] and as a consequence of this If(t) - g(t)1 attains its maximum value (and hence it is finite for every pair of functions f,g), which we take it as doo(f,g)). Thus, we have doo(f,g) = sup If(t) - g(t)l, f,g E Cc[a,b], tE[a,b] see Figure 2.3. Now (Ml) and (M2) trivially holds while the triangle inequality follows from Lemma 2.28(ii). In this way we make the space Cc[a, b] into a metric space and the metric is called supremum/maximum metric on Cc[a, b]. Note that a continuous real/complex valued function on an arbitrary metric space X may fail to be bounded as the example f(t) = t on IR points out. For f, 9 E C[a, b] := CR[a, b], we define another metric (a special case of LP-space which we shall discuss separately) dd!,g) = l b I!(t) - g(t)1 dt. This is explained in Figure 2.4 by the area between the two curves. Also, Figure 2.3 illustrates the meaning of the distance function doo(f,g) when f, 9 E C[a, b]. The vertical line drawn at the place where the width is largest apart, the length of this arrow is doo(f,g) and this gives the distance between f and 9 in C[a, b]. 
80 Chapter 2: Concepts in Metric Spaces s d 1 (f,g) s = g(t) s = f(t) I  o a b t Figure 2.4: d 1 (I, g) = area of the shaded portion More generally, we let C(X, Y) to denote the set of all continuous func- tion 1 from" a metric space (X, d) into a metric space (Y, p) such that the quantity sUPzEX p(/(x), f(a)) is finite, where a E X is a fixed point. Then as above, it may be easily shown that C(X, Y) is a metric space with the metric given by the formula doo(f, g) = sup p(f(x), g(x)). xEX The finiteness condition on d oo (f (x), f ( a )) ensures that d oo (I, g) is finite. Now, if Y := IF where IF = C or IR, then we write C(X) := C(X, y).14 In the special case X = [a, b], we use the notation Cc[a, b] if IF = C, and C[a, b] if IF = III 2.39. Hyperbolic metric. For IZII < 1 and IZ21 < 1, define 1 ( 1 + 4>(Zl, Z2) ) Zl - Z2 d(Zl, Z2) = 2 log 1 t/J( ) ' t/J(Zl, Z2) = 1 - · - Zl,Z2 -ZlZ2 Then the function d defines a metric and is called hyperbolic metric on the unit disc . This metric is important in the study of Hyperbolic geometry. 2.40. Example. Suppose that we are given a metric space (X, d), bounded or not. Then we can always find metrics d* and d' on X defined by d*(x, ) = d(x, y) Y l+d(x,y) and d'(x,y) = min{l,d(x,y)} 14This notation is particularly used when X is a compact space (note that every continuous function on a compact set is bounded). One could also use the same notation C{X) = C{X,IP) to denote the space of bounded continuous functions from the metric" space X to IF with the sup metric. 
2.3. Metric Spaces IP(n), IP and C[a, b] 81 such that (X, d*) and (X, d') are bounded. To show that d* is a metric we first verify the triangle inequality (M3) for d*: d*(x, y) 1- 1 1 + d(x,y) < 1- 1 1 + d(x, z) + d(z, y)' d(x, z) + d(z, y) 1 + d(x, z) + d(z, y) < d(x,z) d(z,y) l+d(x,z) + l+d(z,y)' - d*(x, z) + d*(z, y). by (M3) by (1.60) Alternately, to verify the triangle inequality (M3), we only need to observe that the function f(x) = x/(1 + x) is an increasing function for x > o. The other two axioms follow from the fact that d is a metric. Similarly d' can be shown to be a metric. Note that all distances, d* as well as in d', are less than 1. . 2.41. Remark. We have already met several metric spaces. Now, we indicate how certain modifications of a metric space yield other metric spaces. Let h : [0, 00)  [0, 00) be a strictly increasing function such that h(O) = o and h(t)/t is decreasing. Then it is easy to see that h is subadditive, Le. h(x + y) < h(x) + h(y) for all x, y E (0,00). Furthermore, if d is a metric on a set X, then so is the composition hod, see [AVV, Remark 7.42 (1)]. Moreover, if h(t)/t is strictly decreasing, then the metric hod satisfies a strict triangle inequality for any three distinct points in X. In particular, if d is a metric and a E (0, 1], then also dO: is a metric and setting h(t) = t/(1 + t) we get a quick proof of the triangle inequality for d* in 2.40. Another example in .this respect is that if (X, d) is a metric space for which there exist distinct points x, y, z E X such that d(x, z) = d(x, y) + d(y, z), then it can be easily seen that d(x, z)P = (d(x, y) + d(y, z))P > d(x, y)P + d(y, z)P. Therefore, d P cannot be a metric on X for any p > 1. . By the completeness property of JR, it follows that if A e x, B c X then the set S={d(x,y): xEA, YEB}cIRt 
82 Chapter 2: Concepts in Metric Spaces possess the infimum as well as the supremum (if S is bounded). This observation helps us to present several other concepts in connection with the metric spaces. 2.4 Basic Topology We have already discussed the notion of neighbourhoods in ]R and C. More precisely, if a E IR then a 6-neighbourhood of a in IR means that an open interval {x E IR: Ix-al < 6} where 6 > O. On the other hand, if a E C then an open disc' {z E C: Iz - al < 6} for some 6 > 0 is a 6-neighbourhood of a. In each of the cases the metric involved is the Euclidean: d(x, y) = Ix - yl with x, y E IR in the real case whereas d(z, w) = Iz - wi, z, w E C in the complex case, see Examples 1.32 with n = 1. We have a natural generalization of this notion for a general metric space (X, d): For a EX, 6 > 0, an open ball with center a with radius 6 > 0 is the set B(a;6) := {x EX: d(x,a) < 6}. In the ordinary plane, one says 'disc' rather than 'ball'. The standard convention is that the word 'ball' is used for the solid region. When it is necessary we use the notation Bd(a; 6) to refer to the ball with respect to the metric d. For simplicity, we may write B(6) for B(O; 6) or Bd(6) for Bd(O; 6). Let X be a metric space and let A eX. A point a E A is said to be an interior point of A, if there exists a 6 > 0 such that B(a; 6) C A, and the set A is called an open set iff every point of A is an interior point. The set of all interior points of A is called interior of A and is denoted by int A or AD. Consistent with the old notation the open ball B(a; 6) is often called a 6-neighbourhood of a. Note that it is trivial that X itself a open set. The empty set 0 is an open subset of X (since there exists no point at which the required conclusion could fail to satisfy). The closed ball with center a with radius 6 > 0 is the set {x EX:: d(x, a) < 6} . and it may be appropriate to denote the closed ball by the symbol B[a; 6]. The sphere with center a with radius 6 > 0 is S(a;6):= {x EX: d(x,a) = 6}. 2.42. Remark. It is a tempting fallacy to assume that in a metric space (X, d), the closure of the open ball B(a; 6) is {x EX: d(x,a) < 6}. 
2.4. Basic Topology 83 However, we can easily verify that if x is in the closure of B(a; 6) then d(x, a) < 6 holds. In the discrete metric space (X, d) the closure of B(a; 1) is {a} while {x EX: d(x,a) < 1}=X. In normed spaces, the set {x EX: d(x,a) < 6} is denoted by B(a;6), the closure of the open ball B(a; 6), where d(x, a) = IIx - all, see Example 3.10. Similarly, S(a; 6) can be denoted by 8B(a; 6) in normed spaces. . Note that if X = JR, then S(a; 6) := {a - 6, a + 6}. For example, the line segment {(x,Y)EIR 2 : -1<x<l, y=l} is not an open set in IR 2 since every 6-ball contains points with y  1. But {(x, y) E IR 2 : -1 < x, y < I} and {(x, y) E IR 2 : (x - 1)2 + y2 :F I} are both examples of open sets in }R2. In situation when there exist several metrics under consideration on the same space X, then the symbol such as B (a; 6) may be modified by indicating the metric involved in X by a subscript, as Bd(a; 6). We shall soon see that the concept of "open set" defined above will be used often. As a simple justification of the terminology that we have introduced, let us prove that every open ball is an open set. 2.43. Proposition. An open ball B(a; R) = {x EX: d(x, a) < R} in a metric space X is an open set. Proof. We show that around each point ( such that d( ( , a) < R there is a small 6-ball around ( which is entirely lying in the R-ball around a. Let ( be an arbitrary point of B(a; R). Then d((, a) < R, so that 6 = R - d((, a) > O. Now, B((; 6) = {x EX: d(x, () < 6} and therefore, if x is such that x E B((; 6) then by triangle inequality we have d(x, a) < d(x, () + d((, a) < 6 + d((, a) = R which shows that if x E B((;6), then one has x E B(a;R), Le. B((;6) C B(a; R). As ( E B(a; R) being arbitrary, we obtain that B(a; R) is open. - We note that int (A) could be empty. For example, let X = C, d is the usual metric for C and A={zEC: Izl=I}, or {z=x+i: XEIR}. 
84 Chapter 2: Concepts in Metric Spaces ." .",.-..-. ..1:-' ..... ....... ....--.... - '"-' .... , /' " , , " , , / , / , l , I \ I \ I \ f , I \ t ] ( r \ , , ,: . I C -a =:;J \ a , a-d a a+d \ j \ I \ I \ I , \ I " , / '-. , , '-., " , .... , ." ",-- _ .....;J. .; ... - .... - --- Figure 2.5: Description of open ball in Euclidean spaces (n = 1,2,3) Then int (A) = 0. We note that int (A) is an open set and if A is open, then int (A) = A. In particular, int (int (A)) = int A. In fact, int (A) is the largest open subset of A. IT A, B are two subsets of a metric space (X, d), then it can be easily shown that A c B => int (A) C int (B). Indeed, a E int (A) implies that there is a ball B(a; d) in A and hence in B, since A C B. Thus, a becomes an interior point of B. As a converse of the Proposition 2.43, we can characterize open sets in a different way by means of open balls. 2.44. Proposition. Let G be a subset of a metric space (X,d). Then G is open ill G is a union of a family of open balls. Proof. Let G be open. IT G = 0, there is nothing to prove. IT G  0, then for each a E G there exists a 6a > 0 such that B(a; 6a) C G. Since this inclusion holds for each a E G, it follows that G C U{B(a;d a ) : a E G} c G. (The use of notation 6a is to remind the reader that it depends on the point a E G.) Conversely, let G = U{ B : B E .r}, where .r is a family of open balls. IT the family .r is empty, then G = 0 and so G is open. IT .r is nonempty, then for each point a E G there exists a ball B in .r such that a E B. Thus, by Proposition 2.43, B is open; so there exists ad> 0 such that B(a; 6) C BeG. Therefore, G is open. - 2.45. Definition. A collection B = {G i : i E A} of non empty open sets in a metric space (X, d) is called an open base for (X, d) if every open set in X is a union of sets from B. 
2.4. Basic Topology (1,-1) l.-:'- -:<.-.-.'.,- -".-- ... 1 1 f ... t. t. 1 I t 0 . . . . 1/ t- t (: :: t L_. _.'....., _'_ .......... _ '_ (-1,-1) 85 (1,1) :""" - - _. - .......,... - - 1 I t 1 I 1 '1 .J I 1 1 1 I 1 I 1 1 1 .................. - - ,_,...._1 (1, -1) Figure 2.6: Open unit ball with respect to the d oo metric Proposition 2.44 shows that the collection of open balls form an open base. We now illustrate by examples to show how open balls may take on different interpretation for different choices of the metric d. 2.46. Examples. For x = (Xl, X2) and Y = (YI, Y2) in ]R2, we con- sider the maximum metric doo on ]R2 (see 2.32): doo(x, y) = max{lxl - YII, I X 2 - Y21}. Then, in this metric space (]R2, doo), the unit ball is given by B(O; 1) = {x E]R : doo(x,O) = max{lxII, IX21} < I} = {x E ]R2 : IXll < 1 and )x21 < I} = {x E JR2 : Xl, X2 E (-1, 1) } so that B( 1) is simply the interior of the square with vertices .(:1:1, :1:1), see Figure 2.6. In general, B(a; 6) is the interior of the square of side 26 with center at a = (al, a2), and with sides parallel to the coordinates axes. Similarly, the ball B (a; <5) in the 3-dimensional space (IR 3 , d oo ), where doo(x,y) = max IXk - Ykl, for X,Y E]R3, tk3 is the set of points inside the cube with center at a = (at, a2, a3), sides of length 26, and with the faces parallel to the coordinates planes. In (]R2, d t ), where d l (x, y) = IXI - YII + IX2 - Y21, the unit ball is given by B(O; 1) = {x = (Xl,X2) E ]R2 : dl(x,O) = IXII + IX21 < I} 
86 Chapter 2: Concepts in Metric Spaces (0,1);, ;' , "' , ., , "' , , , , "'. ,., ;'  ;' ;' , ;' (-1,0)", 0 , , ".y " , \. , , , / ,. ./ ./ ./ ./ / (1,0) (0, -1) Figure 2.7: Open unit ball with respect to the dl metric and therefore, B(O; 1) is the interior of the square (diamond shape with respect to the metric d 1 ) bounded by the four lines X1  X2 = 1, which is the interior of the diamond shaped square with vertices (1, 0) and (0, 1). See Figure 2.7 for the cases p = 1 and 00. In general, the ball B(a;) in (]R2, d 1 ) is a square with center a = (a1, a2) E ]R2, diagonals parallel to the coordinate axes and with the length of the diagonals equal to 2. In (1R2,d 2 ), where d(x,y) = (E=llxk -Y k I 2 f/2, the ball B(aj6) is the interior of the circle of radius  with center a = (a1, a2) E ]R2. The unit ball in this case is the disc of radius 1. What we have discussed in this example is the three special cases: p = 1, 2, 00 with n = 2 for IP(n). One can see that if p is close to 1, the' unit ball of lP(2) is nearer to the unit ball of 1 1 (2). Similarly, if p is close to 00, the unit ball of lP(2) is nearer to the unit ball of 1 00 (2). Indeed, if we select p > 1, then the unit ball in (1R2, d,) becomes a figure with curved sides. More precisely, since the unit ball in (]R2 , d p ) is B(O; 1) = {x = (Xl, X2) E ]R2 : (IX1IP + IX2IP)1/p < I}, we observe that as p increases from 1 to 00, these balls are all convex, and these move steadily from a diamond through a circle to a square, see Figure 2.8 (with dp(x,O) = IIxllp). We may also note that in ]R3, they move from an octahedron through a sphere to a cube. Finally, for x = (Xl, X2) and Y = (Y1, Y2) in ]R2 , we define p on ]R2 by (Xl - Y1)2 (X2 - Y2)2 a 2 + .b 2 p(X, Y) = 
2.4. Basic Topology 87 X2 ..-lIxlioo = 1 IIxll2 = 1 IIxll oo < 4- -1 0 1 Xl Figure 2.8: Description of balls with respect to the d p metric (0, b) --- - -,, -,- " ...." ..... ',..,. ,:.-., \ I I \ ( -a, O)(,,:. 0 . .,,<::, .. .......::......."""-..--..:- .))( a, 0) jI ....#.::.) ..... (()": -=b) Figure 2.9: Ellipse where a and b are two fixed positive real numbers. Then (]R2, p) is a metric space and the ball B(O; <5) is simply the interior of the ellipse with its cen- ter at the origin (0,0) and its semi-major and semi-minor axes are equal to a and b respectively, see Figure 2.9. We shall soon see that the met- rics d oo , d 1 , d 2 and p defined above are all equivalent metrics on ]R2 (see Definition 2.67). . 2.47. Example. The open ball with center at xo(t) and radius <5 with.respect to te supremum metric doo(f,g) = SUPtE[a,b] If(t) - g(t)1 on the space C[a, b] is given by B(xo; <5) := {x E C[a, b] : doo(x, xo) < 6}. This means that Ix(t) - xo(t)1 < 6 for each t E [a, b] and therefore, the 
88 Chapter 2: Concepts in Metric Spaces xo(t) + 8 --.... ;,'" ....,  , / , I , " .... ;' / / I / / /  / ;' " " , , " .... '----'  ;' o a b Figure 2.10: Balls in (C[a, b], d oo ) ball around Xo consists of all functions x(t) such that the graph of x(t) lies within a band about xo(t) of width 8 either side. The region in which the graph of x(t) must lie is shown in Figure 2.10. . We remark that if Y e x, then the open balls, in X and in Y are not necessarily same, because for a E Y we have By(a; 8) = Bx(a; 8) n Y where the subscript is used to indicate the metric space involved in the balls: By(a;8) := {x E Y: d(x,a) < 8}, Bx(a;8):= {x EX: d(x,a) < 8}. Therefore, it is important to distinguish between the statements "A is open in X" and "A is open in Y". In fact, there exist examples of subspaces Y of 1R such that a open subset of Y need not be open in III Now, we state and prove the following key properties for open sets. 2.48. Proposition. Let (X, d) be a metric space. Then we have (i) The subsets 0 and X of X are open sets (ii) An arbitrary union of open subsets of X is an open set (Hi) The intersection of a finite number of open subsets of X is an open set. Proof. (i): The interior of the empty set 0 is obviously empty so that o = int 0, which shows that 0 is open. The whole space X is also open subset of itself, since by the definition of an open ball B(a; 8) in X it is true that B(a; 8) C X for each a E X so that every point of X is an interior point. 
2.4. Basic Topology 89 (ii): Let {G i : i E A} be a family of open subsets of X. Let x E UiEA G i . Then there exist some i E A such that x E G i . Since G i is open, there exist a 6 > 0 such that B(x;6) C G i C U G i iEA and therefore, U G i is open. iEA (iii): Let {Gi : 1 < i < n} be a finite collection of open sets and x E n G i . Then x E G i for each i = 1,2,..., n. Since G i is open for 1 <i<n each i there exist a 6 i > 0 such that B ( x' 6. ) C G. , t _ t and therefore, with 6 = minlin 6 i , we have B(x; 6) c n B(x; 6 i ) C n G i lin lin which shows that n G i is open. lin . In Proposition 2.48, we state that only finite intersections of open sets are open. Is an intersection of infinite number of open sets is again open? We provide examples to show that the answer is "no". For example, con- sider (X, d) where X = [-1,1] C IR and d is the usual metric given by d(x, y) = Ix - yl. We know that open intervals in IR are open sets (relative to the usual metric), but {OJ is not open. Thus, for the open intervals, Dn=(-I/n,}/n), nEN, . we see that 00 n Dn = {OJ. n=l Indeed for each n E N, 0 E Dn so that 00 o E n Dn. n=l To show that n  1 Dn contains no element other than 0, we assume that a E n  1 Dn and a :j:. o. Then lal > 0 so that there is a kEN such that I/k < lal which shows that a  Dk = (-I/k, I/k). This is a contradiction to our assumption that a E n  1 Dn. Hence, 0 is the one and only member of n : 1 Dn showing that n  1 Dn = {OJ. On the other hand, we have 00 n En = (-1/2,1/2), n=l E _ ( _ n n ) n- n+I'n+I 
90 Chapter 2: Concepts in Metric Spaces and 00 n D = [0, 1], D = (-l/n, 1 + 1/n). n=l 2.49. Definition. (Limit point and Closed set) A point a in a metric space X is a limit point of A C X iff, for every € > 0, the open ball B(a; €) contains a point of A other than a itself (if a happens to be the point of A) Le. if for every € > 0 (B(a; €) \{a}) n A :F 0. The point a mayor may not be in the set A. A set A C X is said to be closed if it contains all its limit points. Alternatively, we have the following equivalent characterization for closed sets. 2.50. Proposition. In a metric space X, A C X is closed iff its complement X \A is open. Proof. Follows from the following sequence of equivalence: A is closed <==} A contains all its limit points <==} A C contains no limit points of A <==} for every a E A c, there exists 6 > 0 such that (B(a;6)\{a})nA=0, Le. B(a;6)nA=0 <==} for every a E AC, there exists 6 > 0 s.t B(a; 6) C AC <==} A C is open. This completes the proof. . 2.51. Corollary. A closed ball {x EX: d(x,a) < R} is a closed set. We have an equivalent form of Proposition 2.48 for closed sets: 2.52. Proposition. Let (X, d) be a metric space. Then we have (i) The subsets 0 and X of X are closed sets (ii) An arbitrary intersection of closed subsets of X is a closed set (iii) The union of a finite number of closed subsets of X is a closed set. 
2.4. Basic Topology 91 Proof. (i) Since Xc = 0 and 0 c = X, the proof of this part follows from Propositions 2.50 and 2.48(i). For the proof of the second and the third part, it suffices to note the identities: Ca Gi) C = i Gi, (  Gi) C = i G and apply Proposition 2.50. . 2.53. Proposition. Let Y be a nonempty subset of a metric space (X,d) and a E X. Then a E Y iff there exists a sequence {Yn} in Y such that Yn  a. Proof. Let a E Y . Then for each n E N, there exists Yn in YnB(a; 1/n). Since 1 d(Yn, a) < N whenever n > N, the sequence {Yn} converges to a. Conversely, let {Yn} be a sequence in Y converging to a. Then, for € > 0, there exists a positive integer N such that d(Yn, a) < € whenever n > N. Since {Yn : n > N} is infinite, it contains a point of Y different from a. Hence, a is a limit point of Y. . For example, we have (i) In a metric space X, every singleton set {x} is closed. To see this, it suffices to observe that a sequence in a singleton set is necessarily a constant sequence and hence, converges to its constant value. Further, since finite union of closed sets is closed, it follows that, in a metric space, every finite subset is closed. (ii) The intervals [a, b], [a, 00) and (-00, b] are closed subsets of IR. (iii) The set Z of all integers is a closed subset of IR, since 00 ZC= U (n,n+1) n=-oo is open. (iv) The set Q of rational numbers is neither a open nor a closed subset R . 
92 Chapter 2: Concepts in Metric Spaces o 1 o ! 3  3 1 Figure 2.11: Cantor set 2.54. Example. Consider the Euclidean metric space (]R, d), where d(x, y) = Ix - yl and A = IR+. Then 0  A. Since B(O; €) = (-€, €) contains points of A, we conclude that 0 is a limit point of A. If a E IR+ be a positive number, then B(a; €) = (a - €, a + €) which contains points other than a itself. Thus, every a > 0 is a limit point of A. However, if a is negative real number then it cannot be a limit point of A. . 2.55. Cantor set. Let us look at the construction Cantor 15 ternary set E on the interval [0, 1]. In this construction, we recursively remove the open middle-third of each of the remaining closed intervals, see Figure 2.11. Step 0: Given the unit interval [0, 1], divide it into three parts of equal lengths: [0, 1/3], (1/3,2/3), [2/3,1]. Step 1: Remove the open middle-third (1/3,2/3) to get E 1 = [0, 1/3] U [2/3, 1]. Step 2: Continue Step 1 to delete the central open middle-thirds, namely, (1/9,2/9) and (7/9,8/9) to get E 2 = [0, 1/9] U [2/9,3/9] U [6/9, 7/9] U [8/9, 1]. Step n: At the n-th step we have En which consists of 2 n closed intervals of length 3- n . 15G.Cantor (1845-1918) is known for his contribution to set theory. 
2.4. Basic Topology 93 The set E of all points from [0, 1] which remain after infinite steps of the process is called the Cantor set. Thus, the Cantor set E is defined as the infinite intersections E = n  1 En which is a subset of III The set E is closed by Proposition 2.52. 2.56. Topological spaces. A topology 7 on a nonempty set X is a family of subsets of X (called open sets) satisfying the conditions (i)-(iii) of Proposition 2.48. A set X together with a topology 7, Le. the pair (X,7), is called a topological space. One says that 7 defines a topology of the set X. For example, from Proposition 2.48, every metric space is automatically a topological space. A topological space need not be a metric space in general. However, there are very important topologies which are not of this kind. One such topology is the so called Zariski topology which plays a key role in the subject of algebraic geometry. Since this is beyond the scope of this book, we do not get in to explain examples of Zariski topology. It is interesting to point out that a metric space has special properties which some topological spaces do not posses. Recall that a subset G of a metric space (X, d) is called open if for every a E G there exists a 6 > 0 such that a ball B(a; 6) is completely contained in G. A topological space X is called H ausdorfJ if, for each pair of distinct points Xl, X2 in X, there exist two open sets G l and G 2 such that Xl E G l , X2 E G 2 and G l n G 2 = 0. In other words, every two distinct points can be separated by disjoint open sets. It is obvious that every metric space is a Hausdorff space. Motivated by Proposition 2.48 we make the following definition. 2.57. Definition. (Topology of  metric space) Given a metric space (X, d), the set O(d) of all open sets of X is called the topology of the metric space (X, d). Many subsets of X are neither open nor closed. On the other hand, it is possible (see Theorem 2.64) for a set to be both open and closed: for example, X and 0 are both open and closed. However, for most common spaces, the whole space and the 0 are the only sets that are both open and closed. In any case, since open subsets of a topology are just the complements of the closed sets, a topology can be equally defined in terms of closed sets. We have the following definition in terms of closed sets: "A topology 7 on a set X is a family of subsets of X (called closed sets) satisfying the conditions (i)-(iii) of Proposition 2.52. 
94 Chapter 2: Concepts in Metric Spaces 2.5 Continuity and Equivalent Metrics The study of continuity for real (complex) valued functions relies on the presence of Euclidean distance on IR (C). Therefore, it is natural to gen- eralize the idea of the continuity of a function on metric spaces and our formal definition of continuity in the context of metric spaces is a straight- forward translation of the real or complex definition with the Euclidean metric replaced by the general metric function. Let (X, d) and (Y, p) be metric spaces and let f be a mapping of X into Y. We say that f is continuous at Xo E X iff, for every € > 0, there exists a 8 = 8(€,xo) > 0 such that (2.58 ) p(f(x), f(xo)) < € whenever x E X and d(x, xo) < 8. The condition (2.58) is equivalently written as f(x) E Bp(f(xo); €) whenever x E X n Bd(XO; 8), or simply by f(X n Bd(XO; 8)) = f(Bd(XO; 8)) C Bp(f(xo); f). Thus, f : X  Y is continuous at a E X iff for every open ball B' with center at f (a) there is an open ball B with center a such that f (B) c B'. The function f : X  Y is said to be continuous 16 on X iff it satis- fies this condition at every point of X. This is a good old classical "€-8" definition of continuity characterization. In fact, "€-8" parts of the elemen- tary analysis makes extensive use of the topological ideas and techniques. Clearly, if X and Y are the subset of IR, and both d and p are the usual metric for IR, then this definition is exactly the usual definition of the conti- nuity of f. It is trivial that every function on a discrete metric is continuous, see Theorem 2.64. Another trivial example is that every constant function from a metric space into another metric space is continuous. It is easy to rephrase the above definition of continuity in terms of open sets and limit points which is useful in characterizing continuous functions. 2.59. Proposition. Let (X, d) and (Y, p) be two metric spaces. (1) The function f : X  Y is continuous at Xo E X iff f(xo) = lim f(x n ) n-+oo for every sequence {xn} such that X n E X for n = 1,2, . .. and X n  Xo as n  00; that is every continuous function in metric spaces preserve the convergence. This equivalence is called the sequential characterization of continuity. 16Many authors call a function simply a map. 
2.5. Continuity and Equivalent Metrics 95 (2) The function f : X  Y is continuous on X iff for every open set G c Y, the inverse-image of G, denoted by f-l(G) = {x EX: f(x) E G}, is open in X (This part remains valid if the open set G in Y is replaced by a closed set in Y. This property is called the inverse image characterization of continuity). Proof. (1) (=> ): Consider a sequence {xn} in (X, d) such that X n  Xo and assume that f is continuous at Xo. Continuity of f at Xo implies that, . for a given € > 0 there exists a 6 such that p(f(x), f(xo)) < € whenever d(x, xo) < 6. For this 6, since X n  Xo, there exists an N such that for all n > N, d(xn,xo) < 6. Thus p(f(xn), f(xo)) < € for all n > N, Le. the sequence {f(xn)} in (Y, p) converges to f(xo). ( {::: ): Suppose that f is not continuous at Xo. Then there exists an € > 0 such that for every 6 > 0 and x E X, d(x,xo) < 6 but p(f(x), f(xo)) > €. For each n E N, we choose 6 = 1/n. We find that there exist X n in X with 1 d(xn,xo) < - and p(f(xn),f(xo)) > €. n But now, the sequence {xn} converges to xo in X while the sequence {f(xn)} does not converge to f(xo) in Y which is a contradiction. Thus, we complete the proof of (1). (2) (=»: We prove the second part using the standard metric space arguments. First, let f be continuous on X and let G.be an open set in Y. If f-l (G) = 0, then it is open. Otherwise, let Xo be an arbitrary point of f-l(G) so that f(xo) E G. As G is open containing f(xo), there exists an € > 0 such that By(f(xo); €) C G. Continuity of f at Xo then implies that there exists a 6 > 0 such that f(Bx(xo;6)) C By(f(xo);€) c G Le. Bx(xo;6) c f-l(G). Hence, Xo is an interior point of f-l (G). Since Xo is arbitrary, it follows that f-l(G) is open in X. ({:::): Suppose that for each open set G in Y, f-l(G) is open in X. Let Xo E X and € > 0 be given. Then A = By(f(xo); €) is an open ball which is an open set in Y, and so, by assumption, f-l (A) is open in X. Since 
96 Chapter 2: Concepts in Metric Spaces f-l(A) is open in X and Xo E f-l(A), there exists an open ball Bx(xo;c5) such that Bx(xo; c5) C f-l (A). Hence f(Bx(xo; c5)) c A = By(f(xo); f). Continuity of f at Xo is thus established. Since Xo is an arbitrary point of X, this proves that f is continuous on X. Part (2) continues to hold if the open set in Y is replaced by a closed set in Y. This equivalence follows from the fact that f-l(G) is open for every open set G iff f-l (F) is closed in X for every closed F in Y, because f-l(Y\G) = X\f- 1 (G). . Another useful characterization for continuous function may be stated without proof. 2.60. Proposition. For two metric spac es (X , d) and (Y, p), the function f : X -t Y is continuous in X iff f( A ) c f(A) for all A c x. Note that Proposition 2.59 does not say whether f(G) is open whenever G is open in X. In fact, the map f : IR -t JR, x I-t x 2 , clarify this fact. For example, if G = (-1, 1) then f (G) = [0, 1) which is not open in IR. Similarly, the image of a closed set is not necessarily closed for continuous functions. Further, this proposition can be used to verify whether a function is continuous, and is particularly easy when the domain G is either a open set or a closed set as we do for several examples below. From Proposition 2.59, we can also conclude that, in the study of continuity of functions in the metric spaces, it is the family of open sets in each space which is important rather than the actual metric itself. 2.61. Remark. According to Proposition 2.59, to check the discon- tinuity of the real function 1 : (IR 2 , d) -t (IR, d) at (0,0) defined by { XY f(x,y}= ;2+ y 2 if (x, y)  (0,0) ifx=y=O it is enough to consider the sequence X n = (  ,  ) E ]R2. Here d is the usual metric on IR 2 and JR, respectively. Note that X n -t (0,0) as n -t 00, and f(x n ) = 1/2 for all n > 1 so that f(x n ) -t 1/2  1(0,0). Therefore, f is not continuous at (0,0). . In the following example, we demonstrate the continuity concept in various equivalent ways. 
2.5v Continuity and Equivalent Metrics 97 2.62. Example. The characteristic function Xa on a set G is defined by Xa{x) = {  ifxEG if x  G. If G is open in ]R, then for each a E G there exists a ball B(a; 6) in G such that xa(B(a;6)) C Xa(G) = {I} C B(Xa(a);€) where € > 0 is any positive number. Thus, every characteristic function defined on an open set is continuous therein. . 2-.63. Example. Consider the Dirichlet function f : IR  IR defined by f{x) = {  ifxEQ if x E IR \ Q. This is an example of a function defined neither by a equation nor drawn by a curve. Is this function continuous? Is this function Riemann integrable? (see Example 1.28). However, we give three different proofs to show that this function is discontinuous at every point of III Let Xo be an arbitrary fixed point of III Then for each 6 > 0, the open interval (xo - 6, Xo + 6) contains infinitely many both rational and irrational numbers, see [Ap] ("The open interval (a, b) for a < b is uncountable"). In particular, there exist points x E (xo - 6, Xo + 6) such that If(x) - f(xo)1 = 1 for every 6 with Ix - xol < 6. This observation shows that f : ]R  IR is nowhere continuous. Secondly, let {x n } be a sequence of irrational numbers such that X n  x, where x E Q. Then f(x n )  0, f(x) = 1. Therefore, f(x n ) does not con- verge to f(x) and, by Proposition 2.59(1), f is not continuous at the rational points x. Similarly, by considering a sequence of rational numbers that con- verges to a irrational number, we can show that f cannot be continuous at the irrational points. Thirdly, we prove that f is not continuous by finding an open set whose inverse image is not open. To apply this criterion, we choose an open interval G = (3/4,2) in III Now f-l(G) = {x E IR : f(x) E G = (3/4, 2)} = Q which is not open in III Thus, by Proposition 2.59(2), f is not continuous. Similarly, one can easily verify the discontinuity of certain functions by finding either an open set whose inverse image is not open, or a closed set 
98 Chapter 2: Concepts in Metric Spaces whose inverse image is not closed. For example, applying this criterion, we see that 1 : IR  IR defined by f(x) = { : ifx=O if x :F 0 is not continuous. Indeed, 1- 1 (1/2,5/2) = (1/2,5/2) U {OJ and 0 does not belong to the interior of (1/2,5/2) U {OJ. . In the discrete metric space (X, d), we have (i) Sd(XO; 1) = {x EX: d(xo,x) = I} = X\{xo} (ii) Sd(xo;6) = {x EX: d(xo,x) = 6} = 0 for 0 <.6:F 1 (iii) Bd[XO; 6] = Bd(XO; 6) = Bd[XO; 1] = X for 6 > 1 (iv) Bd[XO; 6] = Bd(XO; 6) = Bd(XO; 1) = {xo} for 0 < 6 < 1 (v) Bd[XO; 1] = X and Bd(XO; 1) = {xo}. 2.64. Theorem. Every subset of discrete metric space (X, d) is open as well as closed. Every function 1 from a discrete metric space (X, d) into a metric space (Y, p) is continuous. Proof. Let a E X be arbitrary. Then we see that the ball Bd(a; 6) in the discrete metric space (X, d) is described by Bd(ajc5) = { } if 6>1 if 6 < 1. Clearly, each singleton set {a} in the discrete metric space (X, d) is open, because {a} = B (a; 1) C {a}. Further, each G c X is a union of singletons and is therefore open. Also, for each G eX, the complement GC is open and hence, G is closed. Now, given € > 0 there exists a 6, e.g. 6 = 1 (with Bd(a; 1) = {a}), such that I(Bd(a; 1)) = {/(a)} C Bp(/(a); f). We remark that if we consider X = IR with discrete metric d and Y = IR with usual metric p, then Bd(O; 1) = IR and {OJ = Bd(O; 1) C Bp(O; 1) = (-1, 1). . 2.65. Example. Consider 1 : (IR, d)  (IR, p), x I-t x, where d is the discrete metric whereas p is the usual metric on III By Theorem 2.64, 1 is continuous. In fact, every subset of the discrete metric space (IR, d) is open. Hence, for each open set Gin (IR, p), 1-1(G) is open in (IR, d) and therefore, 1 is continuous. On the other hand, even though S = {x} is open in (IR, d), 
2.5. Continuity and Equivalent Metrics 99 1 (S) = {x} is not open in (IR, p). Also, this example shows that even if 1 is a continuous mapping from a metric space X into another metric space Y and if S is open in X, then it is not necessary that I(S) is open in Y. . 2.66. Example. Consider the usual metric space (IR, d) and define 1 : (IR, d)  (IR, d), x I-t 3. Then for every open set G in IR, we have f-l(G) = {: if3EG otherwise and, since 1-1(G) is open, 1 is continuous on III However, since 1((0,1») = {3}, {3} is not an open subset of (IR, d). Thus, 1 need not take an open set onto an open set. . 2.67. Definition. (Equivalent metrics) Two metrics d and p on the same set X are said to be equivalent if the two topologies induced by the two metrics are the same: O(d) = O(p), Le. if every open set in (X,d) is open in (X, p) and vice versa. In other words, if for each E > 0 and a E X there are 6 1 , 6 2 > 0 such that Bp(a; 6 1 ) C Bd(a; E) and Bd(a; 6 2 ) C Bp(a; E). 2.68. Example. Given a metric space (X, d), we can construct a bounded metric p on the space X by defining (see Example 2.40) p( x, y) = min {I, d( x, y) } . Then, for every E and every Xo EX, we find that {x EX: d(x,xo) < E} C {x EX: p(x,xo) < E} because p( x, y) < d( x, y) for all x, y EX. Therefore, every p-open set is d-open. Conversely, if E > 0 be given then with E' = min{l, E} it follows that {x EX: p(x,xo) < E'} C {x EX: d(x,xo) < E} for every Xo EX. Thus, every d-open set is p-open. Similarly, if d 1 is another metric on X defined by d(x,y) d 1 (x,y) = l+d(x,y) then it is easy to see that d and d 1 are equivalent. Hence, d, p and d 1 are equivalent on X. Now we give an alternate proof for the equivalence of these metrics. Since d 1 (x, y) < p(x, y) < d(x, y) for all x, y E X, 
100 Chapter 2: Concepts in Metric Spaces the identity mapping from (X, d) to (X, p) and the identity mapping from (X, p) to (X, d 1 ) both are (uniformly) continuous. To complete the proof it suffices to show that the identity mapping from (X, d 1 ) to (X, d) is con- tinuous. For € > 0, there exists a 6 = €/(1 + €) > 0 such that d(x,y) < € whenever d1(x,y) < 6, since d1(x,y) < €/(1+€) is equivalent to d(x,y) < €. It follows that all the identity mappings between (X, d), (X, p) and (X, d 1 ) are continuous and hence, d, p and d 1 are all equivalent metrics on X. . 2.69. Remark. For a given metric space (X, d), we recall that there are several new metrics on the same space X which give rise to the same topology for X as d does. Thus, the metric of the given metric space X is not uniquely determined by its topology. For example, consider the metrics doo and d 2 on the same space JR2. Then to say that the metrics d oo and d 2 on JR2 are equivalent, it suffices to note geometrically that inside any disc in }R2 we can find a square, and conversely, inside a square we can find a disc. Similarly, it follows that doo and d 1 on }R2 are equivalent, since inside a square we can find a diamond shaped square and conversely. Therefore, geometrically it is clear that d oo , d 1 and d 2 generate the same topologies on }R2. . The following characterization which gives a sufficient condition for the equivalent metrics is often useful. We omit the proof as it follows easily. 2.70. Proposition. Given a nonempty set X, two metrics d and p on X are equivalent if there exist two positive constants c and C such that (2.71) cp(x, y) < d(x, y) < C p(x, y), for all x, y EX. The converse of this proposition is not true as Example 2.68 points out. Indeed, Example 2.68 shows that cp(x, y) < d(x, y), for c = 1 but the second inequality d(x,y) < Cp(x,y) does not hold, since d(x,y) ,may be arbitrarily large while p(x,y) < 1. Similarly, if X = [1,00) with respect to the usual metric d and p is another metric on X defined by 1 1 p(x,y) = ---, x y then it can be shown that d and p are equivalent metrics. However, it is important to point out that the Cauchy sequence (see Section 2.7) may be used to show that the converse of Proposition 2.70 fails. For this example, 
2.5. Continuity and Equivalent Metrics 101 we note that the sequence {x n }, x n = n, is Cauchy with respect to the metric p but not with respect to the metric d. Consequently, the topological equivalence of two metrics defined on the same set does not necessarily imply that these metrics must be bounded by each other in the sense of the inequalities in (2.71). The situation, however, in the case of metrics induced by norms is simple (see Theorem 5.3). An application of Proposition 2.70 and the discussion in Example 2.32 give 2.72. Corollary. For each p > 1, the metrics d p and d oo defined in Example 2.32 for the space IP(n) are equivalent. In particular, for each a E IRn, (2.34) implies that Bd oo (a; n- 1 / P €) C Bd" (a; €) C Bd oo (a; f). Proof. Recall (2.34) doo(z, w) < dp(z, w) < n 1 / p d oo (z, w). The first inequality shows that the identity mapping from the space (IRn, d p ) into the space (IRn, doo) is (uniformly) continuous. The second inequality shows that if € > 0 is given, then there exists a 6 > 0, namely 6 = €n- 1 / p , such that dp(z,w)<€ whenever d oo (z,w)<6 so that, the identity mapping from (IRn, d oo ) to (IRn, d p ) is (uniformly) con- tinuous. Hence, the metrics are equivalent. _ 2.73. Dense subsets and separability. A subset Y of a metric space X is said to be dense in X if Y = X, Le. every point of X is either a point of Y or a limit point of Y. If Y = Y , then we say that Y is dense in itself. From Definition 2.49, Y is dense in X iff for every a E X and every 6 > 0, B(a; 6) n Y :j:. 0. Equivalently, for each given point a E X and every 6 > 0, there exists a point Yd E Y such that d(a, Yd) < 6. The metric space X is said to be separable if it contains a countable dense subset Y. From the definition, we can say that a separable space X is not too big in the sense that we can approach each of element of the space X through a sequence of elements of a countable set. 2.74. Examples. We list below a list of standard examples of dense subsets, separability and nonseparability. (1) Note that Q is countable, because the members of Q can be put in one-to-one correspondence with N; the set of real numbers, on the contrary, is uncountable. The most well-known examples of dense sets in IR are the set of rationals an d the set of irrationals: Q = IR and IR\Q = III In fact, for each a E IR and every 6 > 0, there exists a point Yd E Q such that la - Ydl < 6. 
102 Chapter 2: Concepts in Metric Spaces (2) The metric space }R with usual metric is separable, since Q is count- able and dense in III (3) Denote by ((r, the set of all points of}Rn each of whose coordinates is a rational number. For each x = (Xl, X2, . . . , X n ) E }Rn, there exist sequences {Xki}kl of rationals such that Xki  Xi as k  00, for each i = 1,2,...,n, and therefore, the sequence {(k} E «r, where (k = (Xkl, Xk2, · · · , Xkn), converges to x, and so X E Qn . Hence, «r is dense in }Rn. Since Q is countable, it follows that Q1 is countable. Therefore, the metric space }Rn with Euclidean metric is separable. (4) Let Y = {q = r + is : r,s E Q} = Q + iQ. Then the one-to-one correspondence between Q x Q and Y is given by the map (r,s) t--+ q = r + is and, since Q x Q is countable, Y is countable. Let a + ib E C \ Y and € > 0 be given. Since Q is dense in IR, there exist rationals x, y such that la - xl < €/2, and Ib - yl < €/2. Then, since Izi < IRe zl + 11m zl for each z E C, it follows that I(a + ib) - (x + iy)1 < la - xl + Ib - yl < €, Le. Y is dense, so that Y is a countable dense subset of C. Consequently, the metric space C with usual metric is separable. The argument of the previous example gives that en is separable. (5) Let Y denote the set of those sequences in IP, 1 < p < 00, which has finitely many nonzero terms and the nonzero entries of the form a + ib with a, b E Q. Then Y is countable. For example, we can choose the countable set as Y = {{rl,...,rn,O,...}: rk = ak +ibk, ak, bk E Q, 1 < k < n}. Let z = {Zk}kl E IP. Then E  1 IZklP is convergent and hence for each € > 0 there exists N such that 00 P L IZk IP < €2 · k=N+l Choose N terms rl,r2,...,rN (rk = ak +ibk, ak, bk E Q for k = 1, 2, . . . , N) such that N P L IZk - Tk IP < €2 · k=l 
2.6. Compactness 103 Now r = {rl, r2, . . . , r N , 0, 0, . . . ,} E Y and for each fixed z E lP we have ( N 00 ) IIp IIp dp(z,r) = L IZk - rkl P + L IZkl P < ( €; + €; ) = € k=l k=N+l so that Y is a countable dense subset of lP. Hence, lP is separable for each 1 < p < 00. (6) The space (1 00 , d oo ) is not separable. For this, we consider a set A consisting of sequences each of whose coordinates consists of zeros and ones, then A is uncountable. Moreover, for each z, w E A, z :j:. w, we have doo(z,w) = 1. Therefore, the family F = {B(a; 1/2) : a E A} of balls of radius 1/2 centered at each a E A is an uncountable family of disjoint open balls. (Note that each open ball in F is singleton.) Suppose that S is a dense subset of 1 00 . Then, each of these balls in F must contain at least one point of S. Our construction shows that the subset S is uncountable so that 1 00 has no countable dense set. (7) From the definition, it follows that if A is dense in B, and B is dense in C, then A is dense in C. . 2.75. Theorem. Continuous image of separable sets are separable. Proof. Let f : X  Y be a continuous map from the metric spaces X into Y and A c X be separable. Then there exists a countable dense subset B of A . Now f(B) is a countable subset of f(A) and, since f(A) c f( B ) c f(B) (see Proposition 2.60), the set f(B) is countable dense in f(A). Therefore, f{A) is separable. _ 2.6 Compactness Let Y be a subset of a metric space (X,d) and let Q = {G a : a E A} be a family of subsets of X such that Y c U Ga. aEA Then we say that the family Q is a covering of Y. If each G a in Q is open, then Q is said to be an open covering for Y. If Q' c Q also covers Y, then Q' is called a s'Ubcovering. H the indexed set A is finite, then Q is said to be a finite covering of Y. We. have the following simple examples: (i) The family Q = {( -n, n) : n E Z+} is a covering of IR and the family Q' = {( -2n, 2n) : n E Z+} is a subcovering of Q. 
104 Chapter 2: Concepts in Metric Spaces (ii) The family :F = {(0,1 - nl ) : n E Z+} is a covering of (0,1) and the family :F' = {(O, 1 - 4(nl» ) : n E Z +} is a sub covering of :F. (Hi) Let Y = {l/n : n = 1,2,...} and 1-/, = {Hn : n E Z+}, where ( 1 1 1 1 ) Hn = n - 2 n + 1 ' n + 2n+l · Then 1-/, is a covering of Y, but it is not the covering of Y U {OJ, since 0  U  1 Hn. On the other hand, if 10 is an open interval which contains the point 0, then H U 10 will be a covering of Y. In fact, 1-/,' = {10,Hl,H2,...,HN} will be a finite sub covering of Y U {OJ, since there always exists a positive integer N such that 11n E 10. A metric space (X, d) is said to be compact or a compact space if X has the property that whenever it is covered by a family of open sets it must be also covered by a finite subfamily. By a compact set in a metric space, we mean a subset of X that is compact when considered as a metric subspace of X. 2.76. Example. Let Y = (0,1) with usual metric of III Is Y com- pact? The family 9 = {(l/n,l): n E Z+} is an open covering of (0,1). Note that this has no finite sub covering of Y. Hence, Y is not compact although it is bounded. N ow consider another space. Let Y = {n : . n E Z +} with usual metric of III Again, it is not possible to find a finite subcover for the open cover given by 9 = {(n-1/2,n+ 1/ 2 ): n E Z+U {OJ}. Thus, Y is not compact. Note that Y is unbounded. . We state without proof the following classical results from real analysis whose proof may be found in standard texts. 2.77. Proposition. (Heine-Borel) IfY is a subset ofJRn, then Y is compact iff Y is closed and bounded. This proposition for n = 1 is really an equivalent form of the following result from real analysis. 2.78. Proposition. (Bolzono- Weierstrass) If Y is a closed and bounded subset of IR, then each continuous function f : Y  IR attains its maximum and minimum. That is, there exist a, bEY such that f(a) = sup f(y) and f(b) = inf f(y). yEY yEY 
2.6. Compactness 105 Does Proposition 2.77 hold in an infinite dimensional space? (see Exer- cise 3.81). Various equivalent formulations for compact metric spaces are available in the literature. For example, we have 2.79. Definitions. A metric space (X, d) is said to be compact if every sequence in X has a convergent subsequence. A subset Y of X is said to be compact if every sequence {Yn} in Y has a convergent subsequence {Ynle} whose limit is an element of Y. The subset Y of X is said to be relatively compact if the closure Y is compact. When various other forms of the definition are being used, compactness of Definitions 2.79 is sometimes referred to as sequential compactness. 2.80. Theorem. We have the following statements: (i) A compact subset of a metric space is separable and bounded. (ii) A compact subset of a metric space is closed. (iii) A compact subset of a metric space is complete. (iv) A closed subset of a compact set is compact. (v) A metric space is compact iff it is sequentially compact. 2.81. Uniformly continuous functions. Let (X, d) and (Y, p) be metric spaces and let f be a mapping of X into Y. In the definition of continuity (see (2.58)) at xo, the number 6 depends on Xo and f. When, given f > 0, there exists a 6 = 6(f), independent of XO, satisfying (2.58) for all Xo E X then we say that f is uniformly continuous on X. Equivalently, f is said to be uniformly continuous on X if for a given f > 0, there exists a 6 = 6(f) > 0 p(f(x),f(y)) < f whenever d(x,y) < 6 for all x,y EX. The notion of uniform continuity is of considerable interest in contexts involving approximating functions by other functions as we do for example in computational analysis and approximation theory. We shall soon see that every continuous function defined on a compact set is uniformly continuous. Clearly every uniformly continuous function of X into Y is continuous at every point of X, but the pointwise continuity does not necessarily imply the uniform continuity. It does not, however, make sense to refer to uniformly continuity at a point. Actually, there are lots of continuous maps which are not uniformly continuous. For examples, let f : (-1r /2, 1r /2)  IR be defined by f(x) = tanx. Then we have f'(x) = (tan x)' = sec 2 x  00 as x  1r/2 
106 Chapter 2: Concepts in Metric Spaces which, by one dimensional Mean Value Theorem, shows that f is continuous but is not uniformly continuous on (-1r /2, 1r /2). In addition to this example, the following functions (i) f: IR  IR, x r-+ x 2 , (ii) f: C  C, Z r-+ Z2, (iii) f: (0, 1)  IR, x r-+ 1/ x, (iv) f:  \{O}  C, Z r-+ l/z, are all continuous in the respective domain of definition with usual metric on IR or C, as the case may be, but are not uniformly continuous therein (see [Po, p.103-105]). In fact, to show that f defined by (i) is not uniformly continuous on IR, it suffices to show that there exists € > 0 such that for all 6 > 0 there exist two. points Xl and X2 with IXI - x21 < 6; but Ix - x I > €. Let € = 1. For any 6 > 0, letting Xl = a - 6/4 and X2 = a + 6/4 shows that 6 IXI - x21 = 2 < 6 but 2 2 6 IXI - x21 = I(XI + X2)(XI - x2)1 = lal 2 > 1 whenever lal > 2/6. Thus, the desired conclusion follows. The same idea would work for the function defined in (ii). Now, for the proof of (iii) and (iv), let 6 be any positive real number. Choose Zl and Z2 such that Zl Z2 = 1 + Zl and consider f(z) = 1/ z. Then Zl, Z2 E (0, 1) and Z2 < Zl, so that Zl < min{l, 6}, IZI - z21 = Zl - Z2 < Zl < 6 and 1 1 --- _ 1 - . Zl Z2 Thus, for every 6 > 0 we have found two points Zl and Z2 in (0,1) (and also in  \ {O}) such that IZI - z21 < 6 and If(ZI) - f(Z2)1 = 1. Alternatively, the conclusion follows if we consider X n = 1/n2 and x = l/n so that X n - x  0 and f(xn) - f(x) = n 2 - n = n(n - 1)  00. We observe that (0,1) and  \{O} are not closed and hence, they are not compact. However, two kinds of continuity is the same when the space X is compact. In particular, is it possible to find a real valued continuous map on a closed interval. that is not uniformly continuous? Can we prove that 
2.6. Compactness 107 every continuous map 1 : (0, 1)  JR, for which the image set 1((0,1)) is an unbounded subset of JR, is not uniformly continuous? 2.82. Theorem. Let (X, d) and (Y, p) be metric spaces, 1 a mapping of X into Y. Assume X is compact. Then, 1 is continuous iff 1 is uniformly continuous. Proof. Assume that 1 is continuous on the compact set X. Suppose on the contrary that 1 is not uniformly continuous. Then there is some € > 0 such that for no 6 > 0 is it the case that p(/(x),/(y)) < € whenever d(x,y) < 6 and X,Y EX. Equivalently, for some € > 0 and for every 6 > 0, there corresponds two points x, y with d(x, y) < 6 and p(/(y),/(y)) > €. Take 6 = 6n =  and denote the values of x, y corresponding to 6n by x n , Yn' Thus, we obtain sequences {x n } and {Yn} such that for every n > 1, d(xn, Yn) <.!. and p(f(x n ), f(Yn)) > f. n As X is compact, we can choose a convergent subsequence X n1c  x as k  00 with x E X, since X is closed. Now, d(Yn1e' x) < d(Yn1e' x n1e ) + d(xn1c' x)  0 as k  00 so that the sequence {Yn1e} converges to x. In particular, the sequence {/(Y n 1e)} does not converge to I(x). This is a contradiction to our hypoth- esis and, therefore, 1 must be uniformly continuous. The other way implication is trivial as it follows from the definition. _ 2.83. Corollary. Suppose I: S I-t IRm is continuous on S, a closed and bounded subset of ]in. Then 1 is uniformly continuous on S. In particular, every continuous function on a bounded closed interval [a, b] is uniformly continuous therein. From the. definition, we note that the composition of two uniformly continuous functions in metric spaces is uniformly continuous. Another simple example of continuous function which is not uniformly continuous is given below. 2.84. Example. Define 1 : IR\ {OJ  IR by f(x) = { r ifx<O if x > 0, 
108 Chapter 2: Concepts in Metric Spaces where r is a fixed positive real number. Then f is continuous on IR\ {OJ but not uniformly continuous. . Now we state the Lipschitz condition which implies uniform continuity. 2.85. Lipschitz condition. A function f : (X, d)  (Y, p) between metric spaces is said to satisfy a Lipschitz condition on X if there exists a constant a > 0 such that p(f(x), f(y)) < ad(x, y) for all x, y E X. We call f a Lipschitz function, or simply a Lipschitzian. The smallest number a satisfying the Lipschitz condition is called Lipschitz constant for f. In the special case a = 1 and X = Y with d = p, we say that f is nonexpansive mapping, see 4.15. In a restricted case p(f(x), f(y)) = d(x, y) for all x, y EX, we say that f is an isometry and we discuss this in detail in 2.100. Clearly, a Lipschitz function is uniformly continuous. If f is real valued function and has a continuous derivative on X = [a, b] with usual metric on IR, then by Mean value theorem f satisfies Lipschitz condition on [a, b]. However, in Chapter 4, we shall discuss more about the additional properties of the Lipschitz functions when a < 1. 2.86. Connected and disconnected sets. We have discussed open sets, closed sets and compact sets. Now, let us briefly discuss the notion of two other classes of sets, namely the connected and the disconnected sets. From the intuitive point of view, a metric space is connected if it consists of a single 'piece'. For example, if we consider the two subsets A and B of IR defined by A  [1,3], B = [1,2] U [4,5] then, according to the intuitive description, we observe that A is connected whereas B is disconnected. Loosely speaking, a set is connected if it cannot be 'split' by two disjoint (open) sets, see Figure 2.12. Our initial setting is a metric space X and let us first transform the intuitive idea of connectedness into a precise definition. A set X is called disconnected if X is the union of two disjoint nonempty open subsets sets Xo and Xl of X: X=XOUX l , x o nx l =0. When this happens, we say that the pair {Xo,X l } disconnects X. If X is not disconnected it is called connected; that is, X is connected iff X cannot be written as the union of two disjoint nonempty open subsets. Equivalently, a metric space X is said to be connected iff either Xo = X and Xl = 0 or Xo = 0 and Xl = X. 
2.6. Compactness 109 ;............. /:11:. ",,:,.'- .... ".. "'. ,:.:( .:, "- ,." . £ " .f':- ':"::\ [' :\ 1\ ::::=>.:........_ _<:1:: , r 1 t.  t: .': .t::)',....:.:«:\. i! ":: .j" \..:- . :" '.,.j,-::.;..."".:::;;.:.:::,;,:,?, '- . '.1 . ":::::";;':::;""'i"",: .. ...... . :.:/ ""'- 1';...: (::. l: f ti V \. ""'" ,:' .' ..:........:"'" ....:::::, :'::::, ?\ :::.:\ ') .) {i .:: .:./)9 ':':::":;,;.0' 11"':'::". l:: A C .. 1\ :;j \.:: ...J \. j \.V .....:.""". '.::'::'-;\:.::' Figure 2.12: Description for connected and disconnected sets Clearly, the empty subset of any metric space is connected. More often the definition of disconnected set is easier for an open set X although the idea is the same. On the other hand, closed sets are more difficult than open sets (e.g. Cantor set) and hence, disconnected sets are more difficult than connected ones (e.g. Cantor set). Moreover, showing that a set X is disconnected is generally easier than showing the connectedness of the same. For, if we can find a point that is not in X, then that point can often be used to 'disconnect' X into two new open sets with the required properties. There is a strong form of connectedness called "path eonnectedness" which is a useful concept. A metric space X is path connected (or arcwise' connected) if every pair of points xo, Xl E X are joined by a curve/path/arc that lies entirely in X. In other words, for every pair of points Xo, Xl of X there exists a continuous mapping 7 : [a, b]  X with 7{a) = Xo and 7{b) = Xl, -00 < a < b < 00. In this definition, without loss of generality, one may choose [a, b] to be the unit interval [0, 1]. Note that if there is a path 71 from a1 to a2 and a path 72 from a2 to a3, then the 7 defined by { 71 (2t) 'Y(t) = 'Y2(2t - 1) for 0 < t < 1/2 for 1/2 < t < 1 is a path from a1 to a3. 2.87. Proposition. In a metric space (X, d) the following are equiv- alent: 
110 Chapter 2: Concepts in Metric Spaces (i) X = Xo U Xl with XO, Xl open nonempty subsets and Xo n Xl = 0. (ii) X = XOUX l with Xo, Xl closed nonempty subsets and XOnXl = 0. (iii) There exists a nonconstant continuous function f : X  Y where Y has the discrete metric. (iv) X is disconnected. Proof. (i){:::::}(ii) follows from the fact that Xo and Xl are both open (or both closed) implies that Xo = X\X l and Xl = X,X o are both closed (or both open). (i)==>(iii): If x E X k (k = 0, 1), then f(x) = k is the required map in (iii) . (iii) ==> (i): Define Xk = f-l(k) (k = 0,1), where the two points of Y are 0, 1. (i){:::::}(iv) is clear. _ 2.88. Proposition. A path connected set is connected. Proof. Suppose that X is path connected but not connected. By the definition of path connectedness it follows that for any pair of points xo, Xl E X, there exists a continuous map 7 : [0, 1]  X with 7(0) = Xo, 7(1) = Xl. By Proposition (2.87)(iii), there exists a continuous map f : X  Y where Y is a discrete metric space. Now, we define a map 4> : IR  [0, 1] by . o ifx<O 4>( x) = X if x E [0, 1] 1 if x > 1. Note that 4> is a continuous function. Now, f 07 0 4> : IR  Y is a noncon- stant continuous function which means IR is disconnected and we obtain a contradiction. _ From Proposition 2.88, it is trivial to observe that discs in C are con- nected sets. For instance, if z E D(a; 6) then the path 7 : [0, 1]  D(a; 6) given by 7(t) = (1 - t)a + tz is a path in the disc joining a to z. 2.89. Example. We list down the following simple examples. We provide the proof for (1) and we leave the rest as an exercise. 
2.6. Compactness 111 (1) Every interval (open, closed, half-open) in }R is connected. Indeed this statement is a consequence of Proposition 2.88. Let I be an interval with a, bEl and a < b. Then [a, b] C I. Consider 1'(t) = (1 - t)a + tb, t E [0, 1]. Clearly, 1'(t) E [a, b] for all t E [0, 1] showing that 1'(t) is a path between a and b lying in I. Thus, I is path connected. Conversely, every nonempty connected subset Y of }R is an interval (possibly degenerate). Indeed, if this is not then there would exist points Y1, Y2 in Y such that Y1 < a < Y2 with a  Y. But then Xo = (-00, a) n Y and Xl = (a, 00) n Y form a disconnection of Y. (2) Every disjoint union of nonempty open subsets of}Rn is disconnected. (3) Any finite or countable set in }Rn (eg. the set of all rationals in JR) disconnected. (4) The Cantor middle third set is disconnected. (5) If X = C and D(a;r) = {z E C: Iz - al < r}, then D(O; 1) U D(2; 1) is disconnected whereas D(O; 1) U D(2; 1) U {I} is connected. . 2.90. Homeomorphism. It is important to know under what condi- tions one can say that two structures are equivalent. As far as set theory is concerned, two sets A and B are equivalent (eg. cardinality) if there exists bijective map which maps A onto B. Now, we shall see when two metric spaces are equivalent. Let (X, d) and (Y, p) be two metric spaces. A function f : X  Y is called a homeomorphism if it is bijective and bicontinuous, i.e both f and its inverse f-1 are continuous. If such a homeomorphism exists, then the two metric spaces are called homeomorphic spaces. In other words, we say that X is homeomorphic to Y, and we may write X = Y. Clearly, the relation "being homeomorphic to" (i.e. = ) is an equivalence relation on the family of metric spaces. A property is called a topological property iff it is preserved by a homeomorphism. Properties of this type are important in topology. A function f : X  Y is called uniformly homeomorphic or uniform homeomorphism if it is bijective and if f and its inverse are uniformly continuous. We have shown that for continuous functions, the inverse images of open sets are open, and the inverse images of closed sets are closed. It is also easy to see that the image of a open (closed) set is not necessarily open (closed). It is then natural to investigate the images of sets under continuous functions. We say that the map f : X  Y is an open mapping if f (G) is open for every open set G in X; it is said to be closed mapping if f(n) is closed for every closed set n in X. Continuous maps need not be open. For example, define f : JR2  IR by f(x) = c for all x E }R2 and for some constant c. 
112 Chapter 2: Concepts in Metric Spaces Then, j, being a constant function, is continuous but for any nonempty open set {} C }R2, we have j ({}) = {c} which is not open in III Similarly, if 9 (x) = x 2 , x E IR, then g((-l,l)) = [0,1) is not open. Consider h defined by 1 h(x) = -, x E [1,00). x Then h([l,oo)) = (0,1] is " not closed in }R although [1,00) is closed in III Another example may be given by the following mapping cP : }R2  IR, ( x, Y) I-t y. Note that cP maps the hyperbola {} = {(x, y) : xy = I} (which is a closed subset of }R2) into the open set ]R \ {OJ. One can see that the image of a bounded set under a continuous function need not be bounded. However, the image of bounded and closed sets under a continuous function is again bounded and closed whenever the function j is defined from }Rn into }Rm (m, n > 1). 2.91. Proposition. (Images of connected sets) The continuous image of a connected set is connected. Proof. Let j be a continuous function from X to Y and X be connected. Clearly, j is continuous from X to j(X), and so without loss of generality, we may assume that Y = j(X). If Y is disconnected, then Y=Y l UY 2 for some disjoint nonempty open subsets Y l and Y2. Therefore, X = j-l(y l ) U j-l(y 2 ) = Xl U X 2 where Xl = j-l(y l ) and X 2 = j-l(y l ) are both nonempty open and disjoint subsets of X. This observation shows that X is disconnected, a contradiction. Hence, Y must be connected. _ 2.92. Proposition. (Images of compact sets) The continuous image of a compact set is compact. Proof. Let (X; d) and (Y, p) be metric spaces, j a continuous mapping of X into Y and let X be compact. Let {Yn} be a sequence in j(X). Then for each n there exists a point X n such that j(x n ) = Yn' As X is compact, 
2.6. Compactness 113 we can choose a convergent subsequence x n "  x as k  00 with x EX, since X is closed. Continuity of I at x implies that Yn" = I(x n ,,)  I(x) as k  00 and the desired conclusion follows by Theorem 2.80(v). . 2.93. Examples. There are continuous bijective maps whose inverse is not continuous. Therefore, we need to illustrate the importance of the condition that the inverse function be continuous in the definition of home- omorphism, (i) Let d be the Euclidean metric on IR and p the discrete metric. Then the identity map from (IR, p) to (IR, d) is a continuous bijection but its inverse is not continuous (see Theorem 2.64). (ii) Next, we give a geometric example. Let X = [0,271") C IR, and Y = {(x,y) E IR 2 : x2 +y 2 = I}, the unit circle in IR 2 and consider the map I : X  Y, I(t) = (cos t, sin t). Clearly, this function is a bijective continuous map of [0,271") to 8 = Y, but its inverse function is not continuous at (1,0). These two examples show that the inverse of a bijective continuous map need not be continuous. Thus, it is natural to look for a suitable additional condition for a bijective continuous function to have the inverse function continuous. . 2.94. Theorem. Let (X, d) be a compact metric space and (Y, p) be a metric space. If f : X  Y is continuous and bijective, then 1-1 is continuous. Proof. Let 9 be the inverse function. We have to show that Yn  y as n  00 => g(Yn)  g(y) as n  00 where Yn, Y E I(X). Let X n = g(Yn) and x = g(y) and assume that Yn  Y as n  00. Suppose on the contrary that x n  x as n  00 is not true. Then there exists an € > 0 such that dx(xn,x) > € 
114 Chapter 2: Concepts in Metric Spaces for infinitely many xn's. We can find a subsequence {x n ,,} that converges to x' EX. But now, we have as subsequence {x n ,,} with x n "  x' as k  00 and dx (x n1c , x) > €. From the second part of the last line we have x :j:. x', whereas from the first part we have f(x') = lim f(x n ,,) = lim Yn" = Y = f(x). k--+oo k--+oo This is a contradiction to our hypothesis that f is one-to-one and, therefore, 9 must be continuous. _ We remark that Theorem 2.94 can also be proved by showing that f is a closed or open map. From the definition, it is clear that f : X  Y is a homeomorphism iff f-l : Y  X is a homeomorphism. Constant map f : IR  IR, x I-t c, is closed but not open in III The map f : IR  IR, x I-t sin x, is continuous but not an open map, since f(IR) = [-1,1]. If f : X  Y is bijective, then we have f-l is continuous <==> f is an open map <==> f is a closed map. This follows from Proposition 2.59(2), since (f-l)-l(G) = f(G) whenever G is open or closed. Similarly, if f : X  Y is bijective then we have homeomorphism <==> continuous open map <==> continuous closed map. Further, we observe that each homeomorphism preserves the convergence of sequences: A one-to-one map f on (X, d) into (Y, p) is a homeomorphism iff, for each Xo EX, X n  Xo in (X, d), iff f(xn)  f(xo) in (Y, p). 2.95. Corollary. If a function f : X  Y between metric spaces (X, d) and (Y, p) is onto and if there exist two positive constants c and C such that cd(x, y) < p(f(x), f(y)) < Cd(x, y), for all x, y EX, then X and Y are homeomorphic. Proof. If f(x) = f(y), then the left inequality, namely, cd(x, y) < p(f(x), f(y)), shows that d(x, y) = 0 which implies x = y so that f is one-to-one. From the given inequalities above, it follows that both f and f-l are Lipschitz functions and therefore, they are continuous. Hence, X and Yare homeo- morphic. _ 2.96. Examples of homeomorphism. Now we present a few exam- ples of homeomorphism: 
2.6. Compactness 115 (i) The function I : IR  IR+ , x I-t exp(x), is a homeomorphism since the function 9 : IR+  IR, x I-t log(x), exists and both are continuous. (ii) The function I : IR  IR, x I-t ax + b (a, b E IR), is obviously a homeomorphism when a :j:. 0, since I is bijective with the inverse function 1-1 : IR  IR is given by x I-t a-Ix - a-lb. Thus, any two open (respectively closed) intervals in IR with usual metric are homeomorphic under this mapping. Indeed, if X = (a, b) and Y = (c, d) (a < b, c < d) then a homeomorphism I : (a, b)  (e, d) is obtained from I(x) = Ax + B by solving c = Aa + Band d = Ab + B. We can rewrite the last two equations as ()=( )() so that () = ab (b 1) () which gives c-d ad-be A= , B= . a-b a-b Thus, the required homeomorphism is given by f(x) = (e - d)x + (ad - be) . a-b Note that I is continuous, bijective and with continuous inverse (a - b)y - (ad - be) yl-t d . c- However, the half-open interval [a, b) is neither homeomorphic to any open interval (c, d) in IR nor to [c, dJ (show!). On the other hand, there is a homeomorphism between the half-open interval [a, b) and [e, d) as well as between [a, b) and (c, dJ. For instance, the map I : [0, 1)  (e, dJ defined by I(x) = (1 - x)d + ex is a homeomorphism. (iii) Now consider the set X = {O, 1, 1/2, . . ., 11n, . . .}. Then X has iso- lated points whereas the set Q has no isolated points. Therefore, it follows that the set of rational numbers is not homeomorphic to the set X. 
116 Chapter 2: Concepts in Metric Spaces (iv) Next, we consider the disc and ellipse defined respectively by D 1 = {(x,y) E JR2 : X2 +y2 < r 2 } and D 2 = {(x,y) E JR2 : x 2 ja 2 +y2jb 2 < I}, where a, b, r > o. The function I : D 1  D 2 , (x, y) I-t r- 1 (ax, by), is a homeomorphism and the inverse function is given by 1- 1 : D 2  D 1 , (x, y) I-t r(a- 1 x, b- 1 y). Finally, we give an important example of a homeomorphism via Stereo- graphic Projection 2.97. Stereographic projection. As we all know, there is a tradi- tional way of representing the extended complex plane as a concrete object, see [Ah, Po]. A natural generalization of this idea in the higher dimensional case is as follows: Let Sn denote the n-dimensional unit sphere whose points are those of JRn+1 which have the distance 1 from the origin: Sn = {x = (Xl, X2, . . . , X n +1) E JRn+1 : X + · · · + X+l - 1 = OJ. Note that for x E JRn+1, ( n+1 ) 1/2 d(x,O) =  x is the distance from x to the origin. Note also that O-sphere in JR1 := JR is then the two points set {-I, I}, the I-sphere is the unit circle in the plane JR2, and the 2-sphere is the surface of a ball of radius 1 in the real 3-space. Let us identify JRn with the set of points of the form (Xl, X2, . . . , X n , 0) E JRn+1 so that we can think of JRn as an object which is sitting inside JRn+1. Let a = N(O, 0,. . .,1) E Sn (This point is called the north pole of the sphere Sn.). Now, we work out a formula for the function II : Sn \ { a}  ]in , called stereographic projection. Actually, we will now show that II is a homeomorphism. We note that which point we have to remove from the sphere Sn is irrelevant because we can always rotate each point of Sn into another point and therefore, for convenience we have chosen to remove the point a = N (0, 0, . . . , 1) from Sn for our discussion. The bijection we shall 
2.6. Compactness 117 produce from Sn \{a} to }Rn is the projection that maps each point P of Sn \ {a} to that point Q of an so that N PQ is a line L in a n + l . Let x = P(XI' X2, . . . , Xn+l) E Sn \ {a}. The line L in }Rn+l that starts from a and passes through x is the set of points of the form Aa + (1 - A)X = ((1 - A)XI, (1 - A)X2,.", A + (1 - A)Xn+I), A E III This line meets an at the point Q provide4 we have zero in their last coordinates. This corresponds to the parameter value, A = - Xn+l (Xn+l # 1 since x # a). 1 - Xn+l Consequently, if x E Sn \ { a }, then L n }Rn , which is the point of intersection of}Rn and the straight line determined by x and a, reduces to the point I1(x) where the coordinates of I1(x) in }Rn may be obtained from the above set of points by substituting \ _ _ Xn+l 1\- , 1 - Xn+l Le. l-A= 1 1 - Xn+l This gives I1(x) = (YI, Y2, . . . , Yn, 0), where (2.98 ) { Xk Yk =  - XnH if k = 1, 2, . . . , n, if k = n + 1. So our projection map II : Sn \ {a}  }Rn is defined by ( Xl X2 xn ) I1(XI,X2,...,X n +l) = 1 ' ,..., 1 ,0 . - Xn+l 1 - Xn+l - Xn+l On any subset of Sn which does not contain point with Xn+l = 1, the projection II defined by this formula is continuous. Indeed, the continuity of II follows from the fact that the component maps IIi : Sn \{a}  a, (Xl, X2,..., Xn+l) t--+ Xi, are continuous, since by the basic properties of the continuity sum, differ- ence, product and reciprocal (where it is defined) of continuous real valued functions are continuous. Conversely, given a point Y = (YI, Y2, . . . , Yn) E }Rn we can always find one and only point x = (Xl, X2, . . . , Xn+l) E Sn \ {a} such that I1(x) = y and arrive at the fact that II has an inverse function II-I : }Rn  Sn \ {a} having the rule for correspondence as . I1-I(y) = (XI,X2,...,X n +I), 
118 Chapter 2: Concepts in Metric Spaces where 2Yk 2 2 2 if k = 1, 2, . · · , n Yl + Y2 + . · . + Y n + 1 2 2 2 1 Yl + Y2 + . . · + Y n - 2 2 if k = n + 1. Yl + Y2 + . . . + Y + 1 The above correspondence is easy to obtain because points on the line joining Q(Yl, Y2, . . . , Yn, 0) and N(O, 0, . . . , 1) are given parametrically by (2.99 ) Xk = {x = Aa + (1 - A)Y = ((1 - A)Yl, (1 - A)Y2,..., (1 - A)Yn, A) : A E IR} and for this point to meet the sphere Sn \ { a }, we must have (1 - A)2y + . . . + (1 - A)2y + A 2 - 1 = O. We note that the term 1 - A (A :j:. 1 because x :j:. a) is a factor of the L.H.S of the last equation and therefore deleting the common factor 1 - A and then simplifying the resulting equation we obtain the solution \ _ EZ=l Y - 1 1\- n 2 ' Ek=l Yk + 1 i.e. 1 - A = En 2 2 l ' k=l Yk + Using this we get the required representation for x = (Xl"", Xn+l) E Sn \ {a} given by (2.99). Thus, we have established the one-to-one corre- spondence between Sn \ { a} and }Rn, Le. II is bijective. Evidently, II-I defined through (2.99) is continuous for reasons similar to those applicable to the function II defined by the formula (2.98). Further, the construction makes it clear that II and II-I are inverses to each other, but one can also directly substitute into the corresponding formula to verify this fact (For a graph when n = 2, see Figure 2.13). From the examples and the defini- tion, we infer that the intuitive meaning of a homeomorphism is that it is a "rubber-sheet transformation". 2.100. Isometry. Let (X,d) and (Y,p) be metric spaces. A function f : X  Y is an isometry iff it preserves the distances: p(f(x), f(x')) = d(x, x') for all x, x' E X. Two metric spaces are said to be isometric iff there exists an isometry of one space onto the other. A property is said to be a metric property iff it is preserved by an isometry, Le. a bijection f : X  Y which preserves distances. More precisely, if X and Yare isometric and if one has the property, then so does the other. In other words, we say that the two metric spaces are "essentially" the same. Note that the algebraic properties of these spaces may differ from each other. It is clear that the relation "being isometric to" is an equivalence relation on the family of metric space. 
2.6. Compactness 119 , , , , , , , , , , . 1J = X2 ((, 1J) = ( + i1J Figure 2.13: Stereographic projection when n = 2 Simple examples of isometries of the Euclidean plane }R2 into }R2 are translation map (T : }R2  }R2, (x, y) I-t (x + a, y + b), where (a, b) is an arbitrary given point of ]R2), reflection map of]R2 in the x-axis (T : }R2  }R2, (x, y) I-t (x, -y)), rotation map (see below) and the glide-reflection map. Further, the metric spaces (}R2, d) and (C, d), where d is the usual metric, are isometric and the isometry is given by f(x,y) =x+iy=z. An easy example from elementary geometry is that every isometry f : }Rn  }Rn (n = 2,3) between two Euclidean spaces (n = 2,3) is a composition of a translation and a rotation about a point. Clearly, since an isometry f : X  Y preserves distances, isometry is injective: f(x) = f(x') => d(x, x') = 0 => x = x'. Therefore, the inverse mapping f-l exists. In fact, d(f-l (x), 1-1 (y)) = d(f f-l(x), f f-l (y)) = d(x, y) so that f-l is also an isometry. Thus, an isometry is clearly uniformly bicontinuous and is always relative to the specified metrics in the two spaces. 
120 Chapter 2: Concepts in Metric Spaces This observation shows that an isometry is a homeomorphism, but not the converse. For example, let 8 E IR. Then the formula To : IR 2  IR 2 , (Xl, X2)  (Xl cos 8 - X2 sin 8, Xl sin 8 + X2 cos 8), or equivalently in matrix form (where 8 is an arbitrarily given angle) T9(Xl,X2) = (:: ()) (  ) represents a counterclockwise rotation of the point (Xl, X2) E IR 2 about the origin (0,0) through an angle 8 with the origin and the coordinate axes remaining unchanged. The relative positions of (Xl, X2) and To (Xl , X2) can easily be indicated pictorially. Note that rotation preserves the distances and hence To is an isometry. Note that To is bijective and bicontinuous. This is an example of a homeomorphism which is an isometry (see also Example 4.6(iv)). On the other hand, the mapping (d - c)x + bc - ad f : [a, b]  [c, d], x  b _ a ' with usual metric both for the domain and codomain spaces, is a homeo- morphism but it is not an isometry unless b - a = d - c. Similarly, the mapping f : IR  IR+, X  eX, is a homeomorphism but not a distance preserving map whenever the metrics involved are the usual metrics. Thus, we have the one way implication: Isometry => homeomorphism => continuity. Our last example of an isometry is the following: Let X n = {z = {Zk}kl E 1 2 : Zk E C, Zk = 0 for k > n}. Define T from the Euclidean space en into X n by (ZlZ2,". ,zn)  {Zl,Z2,... ,zn,O,O,.. .}. Then with Z = (Zl,Z2,... ,zn) and W = (Wl,W2,... ,w n ), we have 00 n d 2 (Tz, Tw) = E I(Tz)k - (TW)kI 2 = E IZk - wkl 2 = d 2 (z, w) k=l k=l and therefore, T is an isometry between en and Xn. 2.7 Cauchy Sequences and Completeness 2.101. Definition. A sequence {xn} in a metric space (X,d) is called Cauchy or fundamental if d(xn, x m )  0 as n, m  00. 
2. 7. Cauchy Sequences and Completeness 121 Equivalently {xn} is Cauchy in (X, d) if for every € > 0 there exists an N = N(€) such that d(xn, x m ) < € for all n, m > N. This is one of the important definitions in Euclidean space (IR, d). For example, a sequence in (IR, d) is convergent iff it is Cauchy. However, as we see below, for a general metric space only one way implication remains valid in general: Suppose that a sequence {xn} in a metric space (X, d) converges to a limit x in X. Then for every € > 0 there exists an N = N(€) such that d(xn, x) < €/2 for all n > N so that € € d(xn, x m ) < d(xn, x) + d(xm, x) < 2 + 2 = € for all n, m > N. Therefore, {xn} is Cauchy and thus we have the following. 2.102. Proposition. Every convergent sequence in a metric space is a Cauchy sequence. Homeomorphism does not preserve Cauchy sequences. See for example, Exercise 2.119: f:IR(-l,l), x x r-+ 1 + Ix!' Does a homeomorphism between semi-metric spaces preserve Cauchy se- quences? Now, we have the following simple and important result. 2.103. Proposition. Every Cauchy sequence (and hence every con- vergent sequence) in a metric space X is bounded. Proof. Let {Xn}nl be a Cauchy sequence in the metric space (X,d). Then for € = 1, there exists an N E N such that d(xn,xm) < 1 for all n,m > N. Then for all n > 1, we have d(Xn,XN) < l+d(Xl,XN)+...+d(XN-l,XN) and the boundedness follows easily. . Consider the sequence {en}nl, en = {O, 0,...,1,0,0,.. .}, of elements from 1 2 . Then for i :j:. j, !!ei - ejll2 = V2 
122 Chapter 2: Concepts in Metric Spaces and, therefore the sequence {en}nl is not Cauchy and hence, it is not convergent. Proposition 2.102 shows that, a necessary condition for a sequence in any metric space (X, d) to converge is that it is Cauchy. A natural question now is whether the converse of Proposition 2.102 is also true: Does every Cauchy sequence converge to a limit? For real or complex sequences in the Euclidean space the answer is yes, bt this is not so in general, as the metric space (Q, d), where d(p, q) = Ip - ql for p, q E Q, shows. Let X n denote the first n decimal approximation of V2. Then d ( x x ) < 10- min{n,m} n, m _ so that X n is Cauchy. But X n  V2, where V2 is not a rational number. Similar reasoning may be applied by taking X n to be the first n terms from the series representation of e: 1 1 X n = 1 + 2! + · · · + (n - 1)! · The Cauchy sequence {1/2n} converges to 0 E Q while the Cauchy sequence {(1 + 1/n)n} converges to the limit e, where e ft Q as e is irrational. In (IR, d) with usual metric, the sequence {1 / n} is Cauchy, because for any € > 0 there exists N > 1/ € so that 1 1 { 11 } --- < max -,- < € forn,m > N. n m n m Thus, {1 / n} is convergent to 0 E IR. On the other hand, if X = (0, 1) or IR \ {O} then the sequence {1 / n} in X with usual metric is Cauchy but does not converge to a point of X as 0 ft X. Another simple example is the sequence of rational numbers X n , where Xl = 2 and X n for n > 2 are obtained inductively from Xn+l =  (xn + :n ) · The first few terms are 2, 1.5,   1.416, :  1.4142, . . . and it is easy to see that this is a Cauchy sequence which does not converge to a rational number(show!). Now, we discuss the method of iteration for computing va, where a > O. Define f : (0,00)  (0,00) by f(x)=  (x+ : ) where a > 0 is the number to which we want to compute the square root. Clearly va is a fixed point of f, Le f(va) = va. Suppose we have some 'initial guess' Xl for the value of va. Then, a/xl is also a good guess for 
2. 7. Cauchy Sequences and Completeness 123 the value of va. If the guess Xl is too big, then a/xl is necessarily too small, and vice versa. Consequently, their average X2 = ! ( Xl + .!!:... ) 2 Xl will be even a better guess than either one. Thus, f(x) provides the average of two approximations of va, one that is too big and the other one that is too small. The same reasoning help us to form a sequence of numbers obtained recursively by iterating the process f(xn) = Xn+l =  (xn + xa n ), for n > 1, where Xl is any reasonable first guess (rational number) at the value of va. Before we start using this process, we must make sure that f(xn) := fn(XI) = fn-l(f(XI)) converges to some point. In fact, we can easily show that if the sequence X n converges then it must converge to the point va because the limit is the solution of the equation X=  (X+ : ), Is {xn} a Cauchy sequence? To check this out, we suppose X n  x. Then we also .have Xn+l  X so that · 2 I.e. X = a. 2 I . I . xn + a X = 1m Xn+l = 1m - n-+oo n-+oo 2x n X2 +a 2x . Solving this for x, we find that X = :i:va. Now, we observe the following: . For all n > 1, we have 2X n (X n +1 - va) - (x + a - 2vax n ) - (- va)2 > 0 so that Xn+l > va for n > 1. . We have Xn+l - X n = -(x - a)/2x n < 0 so that {xn} is decreasing for n > 1 and X n > va for n > 1. · IXn+1 - val = (vx;, - ...[a) 2 /2x n < (vx;, - va)/2a for all n > 1. Thus, we conclude that X n converges and hence, it is Cauchy. Thus, we have a natural question: Under what condition does a Cauchy sequence become a convergent sequence? 
124 Chapter 2: Concepts in Metric Spaces 2.104. Proposition. If a Cauchy sequence in a metric space has a convergent subsequence, then the whole sequence is convergent. Proof. Let {xn1e} be a subsequence of a Cauchy sequence {xn} and X n1e  x as k  00. Since d(xn, x) < d(xn, x n1e ) + d(xn1e' x), the convergence of {xn} follows if we use the standard arguments. _ We shall single out particularly those interesting class of metric spaces in which every Cauchy sequence is convergent: 2.105. Complete metric space. A metric space is called a complete space iff every Cauchy sequence of points in it converges to a point in the space. In particular, one dimensional Euclidean space (]R, d) is complete and we do not give the details of the proof. Using the completeness of (IR, d), one can easily prove the completeness of n-dimensional Euclidean space (IRn, d). We shall discuss the notion of completeness of more general spaces and its application in Chapter 3. However, now we shall discuss some basic properties concerning the completeness. If p and d satisfy the condition (2.71), each sequence which is Cauchy with respect to p is also Cauchy with respect to d" and conversely. Hence we have 2.106. Proposition. Suppose that R and d are two metrics on X satisfying the condition (2.71). Then (X,d) is complete iff (X,p) is complete. The definition of convergence implies that the continuous image of a convergent sequence is a convergent sequence (see Proposition 2.59(1)). From the definition of uniformly continuous, we also have the following 2.107. Proposition. Every uniformly continuous map f from a metric space X into another metric space Y maps a Cauchy sequence in X into a Cauchy sequence in Y. Proof. Let f: (X, dx)  (Y, dy) be uniformly continuous. Suppose that {xn} is a Cauchy sequence in X and that € > 0 is given. By uniform continuity of f, there exists a 6 > 0 such that dy(f(x),f(x')) < € whenever dx(x,x') < 6 and x,x' EX. As {x n } is a Cauchy sequence in X, for 6 > 0, there exists an N E N such that dx(xn,x m ) < 6 for all n,m > N. 
2. 7. Cauchy Sequences and Completeness 125 But then, dy(f(xn),f(xm)) < € for all m,n > N. Since € > 0 arbitrary, it follows that {f(xn)} is Cauchy in Y. . On the other hand, the uniformly continuous image of a complete metric space need not be complete. For example, the map f:]R(-1,1), x x r-+ 1 + lxi' with usual metrics is a homeomorphism as well as uniformly continuous (show!). However, the image f(IR) = (-1,1) is evidently not complete! 2.108. Example. Suppose that there is a sequence {Zn}nl of com- plex numbers (or real numbers) such that IZn+l - znl < alz n - zn-ll for all n > 2, where a E (0, 1) is fixed. Then, by iteration, we find that IZn+l - znl < a n - l lz2 - zll so that for m > n we have IZm - znl - I(zn - Zn+l) + (Zn+l - Zn+2) + ... + (Zm-l - zm)1 m-n < L IZn+k-l - zn+kl k=l m-n < I Z 2 - zll L a n + k - 2 k=l m-n - IZ2 - zlla n - l L a k - l k=l n-l < IZ2-Z11   0 asnoo(sinceo:<l). -a Thus, the sequence {zn} is Cauchy in the complete metric space C (or IR) and hence converges. For example, consider a sequence of real numbers defined recursively as follows: Xl > 0 is chosen, and for n > 1, Xn+l is' defined through the equation Xn+l = (a + xn)-l, n > 1, where a > 1 is fixed real quantity. Then, {xn} converges. Indeed, for n > 2, I I IXn - xn-ll 1 xnH - X n = ( )( ) < 2"lx n - xn-d a+x n a+xn-l a (a=1/a<1) 
126 Chapter 2: Concepts in Metric Spaces which shows that {x n } converges to some x > 0, where I . 1 x = 1m Xn+l = . n-+oo a + hm n -+ oo X n 1 - , a+x Le. x 2 + ax - 1 = o. Hence, the given sequence converges to x = (-a + v a 2 + 4) 12. . 2.109. Proposition. Let (X, d) be a metric space and S e x. (i) If X is complete and S is closed in (X,d), then (S,ds) is complete. (ii) If (S, ds) is complete, then S is closed in (X, d). Proof. (i) Let S be a closed subset of the complete metric space (X, d). Consider a Cauchy sequence {xn} in (S, ds). Then {xn} is Cauchy in (X, d). But, since X is complete, {xn} converges to some x E X. We have either XES or x is a limit point of S. Since S is closed in (X, d), XES. As {xn} is an arbitrary Cauchy sequence, we conclude that S is complete. (ii) Let S be a complete subspace of a metric space (X, d), and let x be a limit point 0f S in (X, d). By Proposition 2.53, there is a sequence {xn} in S\{x} such that X n  x. By Proposition 2.102, {xn} is Cauchy in (S, ds), and since S is complete, we conclude that {xn} converges to some point x' E S. Uniqueness of the limit gives x = x', so that xES and hence S is closed. - We shall see several important applications of completeness property in Chapter 3. 2.110. Theorem. Completeness is preserved under isometry. Proof. Let (X, d) and (Y, p) be two metric spaces and let f : (X, d)  (Y, p) be an isometry. First we assume that X is complete and {Yn} a Cauchy sequence in Y. Then, given € > 0 there exists N such that d(f-l(Yn),f-l(Ym)) = P(Yn,Ym) < € whenever n,m > N so that {f-l(Yn)} is a Cauchy sequence in X and therefore, possesses a limit x E X, since X is complete. Thus, we have f(x) = lim n -+ oo Yn and hence, Y is complete. Similarly, if Y is complete then X is also complete. Hence, X is complete iff Y is complete. - We note that, completeness is not preserved under homeomorphisms, Le. completeness is not a topological property. Indeed, homeomorphisms preserve convergence because of bicontinuity, but they do not necessar- ily preserve Cauchy sequences. For example, consider the usual metric spaces X = (0,1] and Y = [1, (0). Then f : X  Y, x I-t 1/x, is a homeomorphism. We observe that {xn} = {l/n} is Cauchy in X whereas {f(xn)} = {n} is not Cauchy in Y. 
2.8. Completion of Metric Spaces 127 2.8 Completion of Metric Spaces We have already discussed several examples of complete and incomplete metric spaces. A typical way of solving a system of equations is to construct a sequence of approximations to a solution, and then prove that it is a Cauchy sequence. If the space under consideration is complete, then we know that such a sequence of approximations converges to a member of the space. Thus, in this setup, complete metric spaces are more useful than incomplete ones as the incomplete ones are inadequate for many purposes. An intuitive meaning of an incomplete metric space X is that it is, in some sense, a space with hole(s) at "point(s)" where Cauchy sequence(s) should converge there is(are) nothing (For example, we can think of passing from Q to IR by working with Cauchy sequences of rationals as Q with usual metric is not a complete metric space and we also note that Q = IR). From the known examples of incomplete metric spaces, this intuition suggests the possibility of embedding each 'incomplete metric space' as a dense subspace of a larger metric space that is complete. To put it another way, to make any incomplete metric space into a complete metric space, we should be able to "fill the hole(s)-the missing element(s)" by adding to X the new point(s) that serves as the limit of the Cauchy sequences. The resulting space that we end up after filling the hole(s) for all the non convergent Cauchy sequences is called the completion of the given incomplete metric space. Now, our aim in this section is to consider a device by which one can obtain a completion and show that any such completion is unique up to isometry. We remark that the completeness of IR(and hence of C) is assumed throughout the book. 2.111. Definition. A metric space X* = (X*, d*) is called the com- pletion of a metric space X = (X, d) (the completion rather than a comple- tion in contrast to Theorem 2.112) if the following conditions are satisfied: (i) X* is complete (ii) X* contains a dense subspace that is isometric with X. It is easy to list down some simple and well-known examples. With respect to the usual metric of IR, we have (i) The completion of IR is IR itself, (ii) The completion of Q is IR, (iii) The completion of (a, 00) is [a, 00), (iv) The completion of (-00, b) is (-00, b], (v) The completion of each of (a, b), [a, b) and (a, b], -00 < a < b < 00, is [a, b]. 
128 Chapter 2: Concepts in Metric Spaces We shall now discuss the process of completion of a metric space. The completion of a normed space will be dealt in Section 5.8 while the com- pletion of inner product space will be done in Section 6.4. 2.112. heorem. Every metric space has a completion and the com- pletion is unique up to an isometry. Proof. Let X = (X, d) be a given metric space. Let S denote the set of all Cauchy sequences in X. An element of S is then a Cauchy sequence {x n }. Two sequences a = {xn} and {3 = {Yn} in X are called equivalent, written a f"oJ (3, iff lim n -+ oo d(xn, Yn) = O. Step 1: The relation f"oJ is an equivalence relation on S. Indeed, if a = {xn}, (3 = {Yn} and 'Y = {zn} are any three Cauchy sequences in X then we have (i) f"oJ is reflexive, since d(xn, x n ) = 0 for each n so that a f"oJ a (ii) f"oJ is symmetry, since d(xn, Yn) = d(Yn, x n ) for all n so that a f"oJ {3 => (3 f"oJ a (iii) f"oJ is transitive, since d(xn, zn) < d(xn, Yn) + d(Yn, zn) for all n so that a f"oJ {3, {3 f"oJ 'Y => a f"oJ 'Y. Thus, the relation f"oJ decomposes the set S of all Cauchy sequences into equivalence classes, where two Cauchy sequences belong to the same equiv- alence class x. iff they are equivalent. Let X. denote the collection of all these equivalence classes of Cauchy sequences in X with respect to the equivalence relation f"oJ, Le. X. = S / f"oJ. Step 2: If {xn} and {Yn} are two Cauchy sequences in the metric space (X, d), then {d(xn, Yn)} is a convergent sequence of real numbers. Indeed, by virtue of the triangle inequality, we see that d(xn, Yn) < d(xn, x m ) + d(xm, Ym) + d(Ym, Yn) and thus, for all m and n, d(xn,Yn) - d(xm,Ym) < d(xn,xm) + d(Ym,Yn)' Interchanging the role of m and n, it follows that (2.113) Id(xn,Yn) - d(xm,Ym)1 < d(xn,xm) + d(Ym,Yn), for all m, n E N. This observation shows that, if {xn} and {Yn} are Cauchy sequences, then {d(xn, Yn)} is a Cauchy sequence of positive numbers in the complete metric space IR and therefore, the sequence {d(xn, Yn)} is 
2.8. Completion of Metric Spaces 129 convergent in III In particular, the limit lim n -+ oo d(xn, Yn) exists for all CauShy sequences {xn} and {Yn} in X. An immediate consequence of this result is that if {xn} is a Cauchy sequence then, for each x EX, the sequence {d(xn, x)} is convergent, since the stationary/constant sequence {x, X, x, . . .} is a Cauchy sequence in X. Our aim is to complete the following tasks: . to define a metric on X* and make X* is a metric space . to show that X is isomorphic to a subspace Xo of X* . to show that X* is complete . to show that Xo is dense in X*. We remark that the construction of X* from X parallels the construction of the real numbers from rational numbers: for examples the two sequences 3.0, 3.1, 3.14, 3.141, 3.1415, . .. , 22 311 355 3195 3, , 99 ' 113 ' 1017' belong to the equivalence class of rational sequences converging to the real number 1r. Step 3: To construct a function for defining a new metric. Consider two elements x* and y* of X*. Let {xn} and {Yn} be two Cauchy sequences belonging to the equivalence classes represented by x* and y*. We define a function d* on X* x X* as follows: (2.114) d*(x*, y*) = lim d(xn, Yn). n-+oo It is important to show that this definition is well defined. First we observe that the limit always exists by Step 2 and, for this definition to make sense, it is crucial to show that d* (x* , y*) does not depend on the choice of sequences {xn} and {Yn} representing x* and y*, respectively. Indeed, to verify independence of the representatives chosen, we consider {Xn}, {x} E x* and {Yn}, {y} E y*. Then, in view of the definition of the classes x* and y*, we have {Xn}  {x} and {Yn}  {y}, that is (2.115) lim d(xn, x) = 0 = lim d(Yn, Y). n-+oo n-+oo But then, as in (2.113), we obtain Id(xn,Yn) - d(x,y)1 < d(xn'x) + d(Yn,Y) 
130 Chapter 2: Concepts in Metric Spaces so that, by (2.115), Id(xn,Yn) - d(x,y)1  0 as n  00. It follows that lim d(xn,Yn) = lim d(x,y), n-+oo n-+oo as required. This observation proves that the limit in (2.114) is indepen- dent of the choice of Cauchy sequences {xn} and {Yn} representing the equivalence classes x* and y*. Thus, d* is well defined. We obtain in this way a distance function on the set X*. Step 4: The function d* defined in Step 3 is a metric on X*. Note that d is a nonnegative symmetric function and therefore, by the definition of d* in (2.114), the function d* is also nonnegative and symmetric. Since d is a metric on X, we have d(x n , zn) < d(xn, Yn) + d(Yn, zn) for all n. Taking the limits on both sides of this triangle inequality, we see that d*(x*,z*) - lim d(xn, zn) n-+oo < lim {d(xn,Yn)+d(Yn,zn)} n-+oo - lim d(xn, Yn) + lim d(Yn, zn) n-+oo n-+oo - d*(x*, y*) + d*(y., z*) so that the triangle inequality for d* is satisfied. Finally, d* (x*, x*) = 0, and if d*(x*, y*) = 0 then {xn} f"oJ {Yn} so that x* = y*. So, d* defines a metric on X*. Step 5: X is isometric to a subspace of X*. To each x EX, we can associate certain class x* E X*, namely, with the class that contains the stationary / constant sequence { x } := {x, X, . . . , X, . . .}. Let Xo be the set of all such equivalence classes. Clearly, if { x } and { y } are two distinct stationary sequences then x # Y so that d(x, y)  0 and d*({ x },{ y }) - lim d(xn, Yn) n-+oo - d(x,y) # O. (X n = x, Yn = Y for all n > 1) Consequently, { x } and { y } cannot belong to the same equivalent class and therefore, each x* E Xo contains at most one stationary sequence. In view of this, it is natural to identify each element x E X with the equivalence class x* which contains the stationary sequence { x }. This observation im- plies that we can regard the given metric space X as being embedded in X*, where each x E X is represented in X* by the equivalence class of the stationary { x }. 
2.8. Completion of Metric Spaces 131 Now, we define a mapping T : X  Xo by setting T(x) = x* for x E X, where x* is the equivalence class which contains the stationary sequence { x }. This map is onto since, for each x* E X 0, there exists a unique element x E X such that the stationary sequence { x } E x* with Tx = x*. Also, if x,y E X and { x }, { y } are the corresponding constant sequences, then d* (T x, Ty) = d* (x* , y *) = d* ( { x }, { y }) = d (x, y) showing that T is distance preserving surjection from X into Xo. This proves the existence of an isometry between X and Xo C X*. Step 6: Let 'Us show that T(X) = Xo is dense in X*. Consider an arbitrary class x* E X* and an arbitrary € > o. We need to show that the ball Bx. (x*; €) contains at least one point of Xo other than x*. For this, we consider a sequence {x n } E x*. Since {x n } is a Cauchy sequence in (X, d), there exists a positive integer N such that d(x n , x m ) < €/2 whenever m, n > N. In particular, for m = N, we have d(x n , u) < €/2 for all n > N where XN = u E X. Consider the element T(u) E Xo and note that the constant sequence {u,u,u,...,} = { u } E u*(= T(u)), where T is the isometry described in Step 5. By the definition of d*, we see that d* (x*, u*) = lim d(x n , u) < €/2 < €. n-+oo which means that an arbitrary open ball Bx.(x*;€) contains a point u* E Xo. Thus, x* is a limit point of Xo, i.e x* E Xo . Step 7: The metric space (X*, d*) is complete. Let {x} be an arbitrary Cauchy sequence in (X*, d*). Since T(X) = Xo is dense in (X*, d*), for each positive integer n, (2.116) d"(T(xn),x) <.! for some T(xn) E Xo. n We show that {x n } is Cauchy in (X,d). Take € > o. Then there exists a positive number N 1 such that d * ( * * ) € £ N xn,x m < 3 or n,m > 1. Now, for n,m > N 1 , d(xn,x m ) - d*(Txn,Tx m ) < d*(Txn'x) + d*(x,x:n) + d*(x:n,Tx m ) 1 € 1 < -+-+- n 3 m 
132 Chapter 2: Concepts in Metric Spaces Choosing N such that I/N < €/3 and N > N 1 , one has d(xn, x m ) < €, whenever n, m > N. Thus, {xn} is Cauchy in (X,d). Since every Cauchy sequence in X belongs to some element of X*, we can find x* E X* such that {xn} E x*. Now, by (2.116), we have d" (x, x") < d" (x, Tx n ) + d" (Txn, x.) <  + d" (Txn. x"). Also, with x* containing {x n }, we have d*(Txn,x*) = lim d(xn,xm) m-+oo since TX n contains stationary sequence each of whose element is Xn. But {xn} is Cauchy in (X,d), and therefore, the last equality yields that d*(Txn,x*)  0 as n  00. Hence, d*(x,x*)  0 as n  00; that is x  x* in (X*,d*). For an alternate proof of the completeness, see Step 5 in the proof of Theorem 5.89. To complete the proof of the theorem it remains to show that the com- pletion is unique up to isomorphism. _ 2.117. Example. Let X be the set of real numbers {I, 1/2, 1/3,...} and d(x,y) = Ix - yl for all x,y E X. Then (X,d) is a metric space. Clearly, (X,d) is not complete! Indeed, if {xn} C X is a Cauchy sequence then it has a limit point in IR (since ]R is complete). Therefore, we must have either x = l/k for some integer k or x = O. Hence, the completionof X is {O, 1, 1/2, 1/3. . .}. . 2.9 Exercises 2.118. Determine whether the following statements are true or false. Justify your answer. (a) In a metric space (X, d), we have Id(x, y) - d(z, w)1 < d(x, z) + d(y, w) for every x, y, z, w E X. (b) IT Xk (k = 1,2,..., n) are points in a metric space (X, d), then we have n-l d(Xl,Xn) < L d(Xk,Xk+l). k=l 
2.9. Exercises 133 ( c) Let k be a fixed positive real number. Then d on C defined by d(z, w) = min{k, Iz - wi} is a metric on C. (d) A set consisting of a single point x such that d(x,x) = 0 is a metric space. (e) IT d is defined by d(x, y) = Ix - ylA, for x, y E JR, then (JR,:d) is a metric space whenever A E (0, 1]. (f) Let A1,... An be fixed positive real numbers. For x = (Xl,.'" X n ), y = (Y1, · · · , Yn) in }Rn, the function d(x, y) = E=l Ak IXk -Ykl defines a metric on ]Rn . Note: This metric may be called weighted I-metric on }Rn. If A1 = A2 = ... = An = 1, then d(x, y) coincides with the case p = 1 in Example 2.32. (g) For x = (X1,X2), Y = (Y1,Y2) in }R2, the function d(x,y) = I X 1 - Y11 defines a pseudo-metric on }R2 but not a metric on }R2 . (h) Let AI, A2, A3 be three fixed positive numbers. For x = (Xl, X2 , X3), Y = (Y1, Y2, Y3) in ]R3 , the function . ( 3 ) 1/2 d(x,y) =  AklXk - Ykl 2 defines a metric on JR3 if A > 4A1 A3. (i) If X is the set of all bounded real-valued functions and Riemann integrable on [a, b] such that d(f, g) = J: If(t) - g(t)1 dt, then X is not a metric space. (j) The function d on }R defined by d( u, v) = lu 3 - v 3 1 is not translation invariant metric. (k) If d is a discrete metric on ]R2, then the unit circle, i.e. the set of x E }R2 such that d(O, x) = 1, is the punctured plane ]R2 \ {OJ. (I) A discrete metric space is complete. Note: We also observe that a convergent sequence {x n } in a discrete metric space can have only a finite number of points in its range. (m) Every metric space consisting of a finite number of elements is com- plete. (n) Let (X,d) be a metric space and A,B c X. Define the distance between the two sets by p(A, B) = dist (A, B), where p is considered as a function defined from Y x Y into }R+ and Y is the collection of all subsets of X. Then p is not a metric. 
134 Chapter 2: Concepts in Metric Spaces ( 0) In the Euclidean metric space (IRn , d 2 ), the convergence is equivalent to the coordinate wise convergence: The sequence {xk}k>l in IRn converges to x E IRn, where xk = (x,... , x) and x = (Xl"" x n ), iff {X;}kl converges to x p , for p = 1,2,..., n. (p) If p,q > 1, Z = {zn} E lP and W = {w n } E lq, then zw = {znwn} E lr with r = pq/(P + q). (q) In the Euclidean metric space (IR, d), the set Q is not open whereas in the metric space (Q,d), where d(r,s) = Ir - sl, r,s E Q, the set Q IS open. (r) In the Euclidean metric space (IR, d), the set of all irrational numbers is not open. (s) The function d : N x N  IR, d(x, y) = Ix - yl, defines a metric and the ball B(O; 6) is given by N n (-6,6). (t) Let (X, d) be a metric space, A C X and a be a limit point of A. Then for any 6 > 0, the ball B(a; 6) contains infinitely many points of A. In particular, arbitrary finite subset of a metric space is closed. Note: If the intersection B(a; 6) n A were infinite, then the set of all distances d(a, Xi) might not have a minimum in this case. Thus, finiteness is essential to define 6'. (u) Any finite subset A, say A = {Xl, . . . , x n }, of IRn cannot be open. (v) If S is a nonempty subset of a metric space (X, d), then S consists of single point iff d(S) = O. (w) In the Euclidean metric d on C, if A and B are defined by A = {z E C: Izi < R}, and B = {z E C: Iz-2RI < R}, R > 0, then An B = 0 and dist (A, B) = O. (x) If A is closed, B is compact and AnB = 0, then dist (A, B) is positive. In particular, x E A <==> d(x, A) = o. (y) In a metric space (X, d), it is not always true that if A ex, then there exist points x and y in A such that d(A) = d(x, Y)L (z) Let A be a nonempty subset of a metric space (X, d). IT x E A, then d(x, A) = 0 but not conversely. 2.119. Determine whether the following statements are true or false. Justify your answer. (a) Every subset of a metric space is open iff every singleton set is open. (b) In a metric space (X, d), every singleton set is closed. ( c) The convex hull of a closed set in IR is not closed. 
2.9. Exercises 135 (d) In a metric space, the complement of every finite set is open. Note: It follows from this hint that every single point set is open in discrete metric space and hence, by Proposition 2.48(ii) (see also Example 2.65), every subset of a discrete metric space is open (and hence all subsets are closed). We observe that in the space IR with usual metric, the single point sets are not open. (e) The sequence {x n = arctann}nl is Cauchy in (-1r/2,1r/2) whereas {tanxn}nl is not Cauchy. (f) If Y is the set of all sequences in which all but a finite number of terms are zero, then Y is not complete. (g) For every pair of distinct points al, a2 in a metric space X there exist two disjoint balls center at al and a2, respectively. (h) In a metric space X, we have B(a; 6') C B(a; 6) for 0 < 6' < 6. and there exists also an example of a metric space X for which B(a; 6') = B(a; 6) even though 0 < 6' < 6. (i) There exists a metric space in which a open ball B(a; 6) and the corresponding closed ball B[a; 6] may be same. (j) Given a metric space (X,d), there exists another metric p on X such that d(x, y) < p(x, y) for all x, y EX. (k) The metric space (N, d) in Example 2.7 is not complete whereas (N, p) is complete when p(x,y) = Ix-yl, x,y E N, is the usual metric on N. (1) If d and p are respectively the discrete and usual metric on IR, then f : (IR, p)  (IR, d) is not necessarily continuous. (m) If A, B, C are subsets of a metric space X, then the triangle inequality d(A, C) < d(A, B) + d(B, C) is not necessarily true. (n) The real valued function f(x) = l/x is not uniformly continuous on {x : x > O}, and the complex valued function f (z) = 1/ z is not uniformly continuous on {z E C : Rez > OJ. (0) The subset Q of IR is neither open nor closed. (p) There exist continuous mappings that are not open but closed. (q) There exist continuous mappings that are neither open nor closed. (r) There exist open mappings that are neither continuous nor closed. (s) There exist closed mappings that are neither continuous nor open. (t) The map f : (0, 1)  IR, x I-t 10g(x/(1- x)), is a homeomorphism. (u) The map f : (1, 00)  (0,1), x I-t l/x, is a homeomorphism. (v) The map f : IR  (-1,1), x I-t x/(l + lxI), is a homeomorphism but not an isometry. 
136 Chapter 2: Concepts in Metric Spaces (w) The map f : (-1, 1)  IR, x I-t x/(1 - lxI), is a homeomorphism but not an isometry. (x) The composition of finite number of isometric transformations of IR 2 is an isometric transformation. (y) There exist homeomorphisms which are not isometries, but preserve completeness. (z) If d and p are respectively the discrete metric and usual metric on IR, then the identity mapping I : (IR, d)  (IR, p) is bijective and uniformly continuous but not an isometry. 2.120. Give examples in which one encounters with d-function violat- ing one of the three axioms (Ml)-(M3) but satisfying the other two axioms. 2.121. Let d be a metric on a set X. (i) Is lfJ a metric? (ii) Is p = Jd a metric? (iii) Is p defined by p(u, v) = min{d(u, v), I}, a metric? 2.122. Let f(x) = x/(1 + v' 1 + z2). Is d : IR x IR  JRt, defined by (x,y)  If(x) - f(y)l, a metric on IR? 2.123. Let d l and d 2 be two metrics on the same set X. Prove that d l and d 2 are equivalent iff both the identity map from (X, d l ) to (X, d 2 ) and the identity map from (X,d 2 ) to (X,d l ) are continuous. (The identity mapping f from (X, d l ) to (X, d 2 ) is defined by f(x) = x for all x E X. Note that the domain and the range are the same sets but have different metrics). Also, answer the following questions: (i) Is d(u,v) defined by min{d l (u,v),d 2 (u,v)} a metric? (ii) Is p( u, v) defined by p = d l + d 2 a metric? 2.124. IT d is a metric on a non empty set X, for which values of A E IR, is d A is also a metric? 2.125. Let f : IR  JRt be continuous. Define d(a, b) = l b f(t) dt, m = [°00 f(t) dt, M = 1 00 f(t) dt. Show that (IR, d) is a metric space and show also that it is isomorphic to the open interval (m, M) with usual metric. 2.126. Show that each metric p defined on a finite set X is equivalent to the discrete metric d on X. 
2.9. Exercises 137 1  , , , " , " " " , , , " ' " " " , " , " "  1  \           \ \ \ \ \ \ \ o x y y o x Figure 2.14: v 2.127. For X = IR := IR U {oo}, the extended set of real numbers, let Ij(x) (j = 1,2) be given by x/(l+lxl) ifxEIR 11(X)= 1 ifx=+oo, -1 ifx=-oo and arctan (x) if x E IR 12(X) = 1r/2 if x = +00 . -1r/2 if x = -00 If d is defined by d(x,y) = I/j(x) - Ij(y)1 (j = 1,2), then show that v (IR, d) becomes a bounded metric space. Are 11 and 12 one-to-one? If \J Y = [-1, 1] with Euclidean metric, is IR isometric to Y for the first function? \J If Y = [-1r /2, 1r /2] with Euclidean metric, is IR isometric to Y with respect to the metric of second function? Is IR homeomorphic to (-1r /2, 1r /2) under the map I(x) = arctanx? Note: For x, y E IR, d(x, y) = I arctan(x) - arctan(y)I represents the angle shown in Figure 2.14. Note that d(x,y) < 1r for all x,y E III 2.128. For X = IR, define d(x,y) = arctan(x - y) for x,y E III Check whether d is a metric on IR? Note: We need some work to establish the triangle inequality. 2.129. Let X be a nonempty set and let d : X x X  IRt satisfy the following conditions (i) d(x, y) = 0 {:::::} x = Y (ii) d(x, y) < d(x, z) + d(y, z). Show that d is metric. 
138 Chapter 2: Concepts in Metric Spaces 2.130. Let X denote the space of all convergent sequences of complex numbers. For u = {Un}nl' and v = {Vn}nl' define the function d by d(u,v) = l lim (un - vn) l . n-+oo Prove that (X, d) is not a metric space. 2.131. Use €-6 notation to show that the function J.t : (C[a, b], d oo )  (C[a, b], d oo ), x(t) I-t x2(t), is continuous, where d oo is the supremum metric defined in Example 2.38. I 2.132. Let f be a real valued continuous function on [0,00). Suppose that either the restriction of f on [b, 00) is uniformly continuous for some b > o or lim x -+ oo f(x) exists. Show that f is uniformly continuous throughout [0,00). 2.133. Given two metric spaces (X, d) and (Y, p), check whether the following statements are equivalent or not: (i) f: X  Y is not uniformly continuous on X. (ii) There exists an € > 0 such that for every 6 > 0 there are points X6 and x:S in X such that d(X6, X:S) < 6 and p(f(X6), f(x:S)) > €. (iii) There exists an € > 0, and two sequences {xn} and {x} in X such that for every n d(xn, x) < 6 and p(f(xn), f(x)) > €. Note: For example, eX is not uniformly continuous on IR with usual metric because for X n = log n and x = log( n + 1), we have IXn -x1 = Ilog(n/(n+1))1  0 as n  00, If(xn)- f(x)1 = 1 for all n. 2.134. Prove that the function f : (0, 1)  IR defined by f(x) = sin(l/x) is not uniformly continuous on (0,1). Is g(x) = cos(l/x) is uniformly continuous on (0, I)? 2.135. Prove that if the sequences {x n } and {Yn} in a metric space (X, d) converge to x and Y respectively, then the sequence {d(xn, Yn)} con- verges to d( x, y) in IR with usual metric. 2.136. Let (X, d) be a metric space. Define (X x X, p) by p( (Xl, X2), (x, X)) = d(Xl, X) + d(X2, x). 
2.9. Exercises 139 Prove that a sequence {(xf,x)} converge to (xi, x;) iff both {xl} and {x 2 } converge to xi and x; in (X,d), respectively. 2.137. Let I be a continuous map of a metric space (X, d) into itself. Prove that the map T : X  JR, x I-t d(x, I(x)), is continuous, where ]R is equipped with the usual metric obtained from the absolute value. 2.138. If a mapping I : X  Y of metric spaces is continuous and bijective, is 1- 1 necessarily continuous? Justify your answer. 2.139. Find the interior and closure of the subset G of}R2 defined by G = {(x,sin(l/x)) : x E (0, I)} when}R2 is equipped with the usual metric d 2 on }R2 . 2.140. Prove that [0,3] is not homeomorphic to [-1,1) U (2,3]. Show also that there is no continuous onto map from [0,3] to (0,3). 2.141. Assuming}R is complete with respect to the Euclidean metric on JR, show that C is complete with respect to the Euclidean metric on C. 2.142. Let X be a metric space and Y C X. Show that Y is dense in X iff X \ Y has no interior point. (In particular, Q is dense in lll) 2.143. For Zk E C (k = 1,2, . . . , n) and p > 1, prove the inequality ( n ) p n  IZkl < n P - 1  IZkl P , 2.144. If {an} is a sequence of nonnegative real numbers such that E  1 a < 00, then show that ( 00 ) IIp La n=l is a decreasing function of p for p > o. 
Part II BANACH SPACES Banach space theory plays a special role in functional analysis and more interestingly in the infinite dimensional spaces. In Section 3.1, introduce the concept of normed spaces which is fundamental to the development of the theory of Banach spaces. Therefore, the material presented in Section 3.1 is essentially a foundation on this topic. The notion of normed spaces can be thought of as a generalization of the n-dimensional unitary space en with the Euclidean length given by ( n ) 1/2  IZk 1 2 The norm lIull, which is assigned to each element u in a normed linear space V, will then be used to define a metric and hence the convergence Un  u as n  00 in V, by means of the equivalent condition lIu n - ull  0 as n  00. An important observation that we shall see in Section 3.1 is that every normed space is a metric space and therefore a topological space; thus it is natural that the topological concepts such as open subset, closed subset, limit, closure, denseness, compactness, relative compactness, separability, connectivity etc., make sense on a normed space. Further, among the class of normed spaces, the most important ones are the complete normed spaces which are universally known as Banach spaces. We develop these concepts in Chapter 3. Section 3.2 discusses the notion of convexity and complete- ness. Section 3.3 and 3.5 include important examples of Banach spaces. Further, we also illustrate the fact that the same vector space can generate different normed spaces, see Sections 3.3 and 3.4. We also examine several standard examples of Banach spaces in Sections 3.3-3.5. Chapter 4 is devoted fully to the fixed point theory. First, we briefly review some basic facts in fixed point theory. Then, we start with the notion of contraction mappings defined on a metric space (and also on some of its subsets), Le. those mappings such that the distance between the images of any two points is less than the distance between these points. We prove some basic results such as the Banach contraction principle which becomes a useful tool for the proof of various existence and uniqueness theorems, for example in the theory of differential and integral equations. Further, we 
142 also discuss several simple examples in order to understand the application part of fixed point theory. In Section 5.1, we first prove an important result stating that all norms in a finite dimensional vector space are equivalent (whereas this is not the case in infinite dimensional spaces). In Section 5.4, we include the Bernstein proof of the Weierstrass approximation theorem (see Theorem 5.47), and we observe that Bernstein's proof actually displays a sequence of polynomials that approximate a given continuous function in C[a, b]. Moreover, Bernstein proof leads to a powerful Bohman-Korovkin theorem (see Theorem 5.57). In Section 5.6, we study certain linear operators and functionals. An important result about the linear operators is that a linear operator between normed spaces is bounded iff it is bounded on every ball, iff it is bounded on some ball, iff it is continuous at some point, iff it is uniformly continuous, see Theorem 5.66. The set of all bounded linear operators between normed spaces X and Y is denoted by B(X, Y). When Y = F, then the members of B(X,]F) are called functionals on the normed space X. The set of all bounded linear functionals on X is called the dual space of X and this is denoted by X., instead of B(X,]F). We show in Theorem 5.70 that B(X, Y) is a normed space if the addition and the scalar multiplication are defined pointwise. In fact, B(X, Y) becomes a Banach space if Y is Banach (whether X is Banach or not). In particular, the operator norm T I-t IITII makes B(X) = B(X, X) a Banach algebra, Le. a complete normed algebra (A normed algebra is an algebra A together with a norm a I-t lIall satisfying the submultiplicativity: lIabll < lIalillbll for a, b E A). In Chapter 5, we also establish an important result known as the "Open Mapping Theorem" which asserts that an onto bounded linear operator between two normed spaces is an open mapping, Le. it carries open sets onto open sets. 
Chapter 3 Normed Spaces In this chapter we shall first discuss the notion of a norm on a vector space and give several examples of normed vector spaces. These are the linear vector spaces with a length factor or norm defined on them. Throughout Chapter 3 we shall consider real or complex vector spaces only. The study of normed vector spaces requires a vector space together with a "measure of length" called "norm" which is in fact an analogous concept to the length of a vectors in }R3 . Our early examples of normed vector spaces may be divided into three kinds; namely those which are subspaces of}Rn or en (eg. IP(n), 1 < p < 00), those which are subspaces of sequence spaces (eg. IP-spaces, 1 < p < 00), and those which are subs paces of functions spaces (eg. CF[a, b]). We will refer to spaces of the first kind as coordinate spaces, the second kind as sequence spaces, and the third kind as function spaces. Since a coordinate spaces consists of functions on a finite subset of N and the elements of the sequence spaces are functions on N, these two spaces may also be considered as functions spaces. Next, we proceed to the topic of Banach spaces. A Banach space is simply a complete normed space. Thus it contains the limit of all its Cauchy sequences. 3.1 Properties of Norm The concept of norm was introduced in order to give a method for measuring the magnitude of a vector. For example, if x = (-1,2,-3,-7,-11) is in ]R5, then Ilxll = 11 is the vector norm 17 which is the length of the largest 17The number IIull will be read as "norm of u" and will be used throughout this book as various generalizations of the elementary Euclidean distance. 
144 Chapter 3: Normed Spaces coordinate. The Euclidean norm in en will be defined by ( n ) 1/2 IIzII2 =  I Z kl 2 , Z = (Zl, Z2,..., Zn) E en, which is the same as the Euclidean distance of the point z E en from the origin. For example if n = 1 we have IR 1 = IR and C 1 = C so that x E IR => IIxll = Ixl which is the absolute value of the real number x, and that x + iy = z E C => IIzll = Izl = V x 2 + y2 which is the modulus of the complex number z, Le. the length of the vector emanating from (0,0) to (x,y) E IR 2 :: C. This observation shows that the concept of norm which we are going to define explicitly is actually a generalization of the concept of (Euclidean) length that is familiar for the set of real or complex numbers. The set of vector norms on en known as the p-norms is defined by ( n ) IIp IIzllp =  IZk IP For p = 2 this corresponds to the Euclidean norm in en. The norm based on the length of the largest coordinate corresponds to p = 00, which is given by (3.1) if 1 < p < 00. Ilzlloo = max IZkl, lkn see Section 3.3. To obtain a relationship between the algebraic structure and the metric properties of a vector space V, we study a metric on V obtained by means of a norm. The resulting space will be called a normed space. As a first step towards achieving this relation, we introduce the following terminology in which the three axioms (N1)-(N3) are the extension of the familiar properties of the Euclidean length in the plane: 3.2. Definition. Let V be a linear/vector space over the field IF (= C or IR). A norm on V is a mapping/function II · II from V to IRt , 11.11 :VIRt, satisfying the following three axioms 18 : (N1) lIuli = 0 => u = 0 (N2) IIAul1 = IAIliuli for all u E V and all A E IF (N3) lIu + vii < lIuli + Ilvll for all u, v E V. We call the pair (V, II · II), a n ormed space. 19 18Note that 111.£11 is to be thought of as the distance from the zero element Ov to u. A vector of norm 1 is called. a unit vector. 19 Also called a normed vector/linear space. [Positivity] [Homogenei ty] [Triangle inequality] 
3.1. Properties of Norm 145 There are two definitions: the real norm that is applicable to a real vector space and the complex norm that is applicable to a complex vector space. The condition (N2) for A = 0 gives that 11011 = 0, which means that u = 0 => Ilull = o. When we refer to Vasa normed space, it is always assumed that there is a norm II . II defined on the vector space V. A seminorm is one which satisfies the axioms (N2) and (N3), but not necessarily (N1). The seminorm is usually denoted by 1.1, when there is no confusion with the absolute value of a complex number. When the seminorm is given on V, we say that V is a semi normed vector space. A seminorm generalizes the notion of a norm in the sense that vectors other than the zero vector are also allowed to have zero length. Let (V, 11.11) be a normed space and S be a (linear) subs pace of V. Then it is clear that S is also a normed space with respect to the norm II . II, Le. the restriction of the norm II · II to S is also a norm on S. We may denote this restriction by II . lis and call (S, II . lis) a normed linear subspace of (V, II · 11). As usual, when there is no danger of confusion, we refer to S as a normed subspace of V, rather than the more accurate usage (S, II. lis) as a subspace of (V, II . ID. 3.3. Example. The proof that the p-norm on en defined by (3.1) is a norm may be done using the Minkowski inequality (see Lemma 2.26(i)). So, we leave this as an exercise. . For a given normed space V, we can always define a function d : V x V  IRt , called the distance from u to v, by (u, v) t--+ Ilu - vii so that d becomes a metric on V; Le. d defined in this fashion satisfies all the axioms of the metric, see Definition 2.1: (i) II u - v II = 0 iff u = v, (ii) lIu  vII = IIv - ull, (iii) lIu - wll < lIu - vii + IIv - wll, for all u, v, w E V. The metric defined in this way is often referred to as the natural metric induced by the norm. Thus, a normed space is automatically a metric space and, hence a topological space. We shall always assume that a normed space carries this metric and its associated topology, which we call the norm topology. Thus, a normed vector space combines the algebraic structure of a vector space with the topological structure of a metric space. In conclusion, we have 3.4. Proposition. Every normed space (V, II . ID is a metric space with respect to the distance function d(u, v) = Ilu - vII, for u, v E V. 
146 Chapter 3: Normed Spaces 3.5. Simple norms on }R2. On V = }R2, we have a I-norm on }R2 defined by Ilxlh = IXll + I X 21, x = (Xl,X2) E]R2. Still another norm on }R2, known as the elliptical norm II. lie, is given by Ilxll e = X X 2 a 2 + b 2 ' x = (Xl, X2) E JR: , for some fixed a, b > O. The unit ball on the normed space (}R2, II · lie) is then given by { 2 2 } 2 2 Xl X2 {x E IR : Ilx II e < I} = (Xl, X2) E IR : a 2 + b 2 < 1 · If we define IlxliM = max{lIxll e , IIxlh} and IIxli m = min{llxll e , IIxlh} then (1R2, II · 11M) becomes a normed space whereas (}R2, II · 11m) is not (verify!). These examples clearly indicate that a given vector space may have several norms leading to the existence of different normed spaces. 3.6. Converse of Proposition 3.4. While normed spaces give inter- esting examples of metric spaces, there are many interesting examples of metric spaces that do not come from norms. This means that the converse of Proposition 3.4 does nbt hold in general. Indeed, one can give several examples of metric spaces which are not normed spaces, i.e. each of whose metric is not given by any norm in the sense of Proposition 3.4. For ex- ample, consider the discrete metric space defined in Example 2.6. Another simple example is to consider the bounded metric space (X, d) of Example 2.18. This metric space cannot be a normed space because if there exists a norm such that d(u,v) = lIu - vii, then it should satisfy (N2). But it is a simple exercise to see that the condition (N2) is not satisfied for the metric space of Example 2.18. We note that the Chordal metric X(z, w) on C (see Example 2.8) is another example of a metric which is not induced by the norm, as the Chordal metric X(z, w) defined by Iz-wl X(z, w) = J(l + Iz1 2 )(1 + Iw1 2 ) does not satisfy the property (N2) because X(AZ, AW)  IAI X(z, w). Moreover, the natural metric induced by the norm is translation invari- ant, i.e. d(u + W,'V + w) = d(u, v), for all u, v, w E V, and it is also a homogeneous metric because d(AU, AV) = IAld(u, v). 
3.1. Properties of Norm 147 This observation shows that d(u, v) = Ilu - vii = d(u - v, 0) so that the translation invariance property helps in answering many ques- tions by transforming them to the corresponding questions about conver- gence to the zero vector. 3.7. Convergence. We now extend the definition of convergence of sequences in a set of points to functions in the normed space (V, II.ID which may be classified into two categories: those normed spaces in which every convergent sequence in V has a limit in a subspace Y of V and those in which not every convergent sequence in V has a limit in Y. Let 11.11 be a norm on a vector space V over F. We say that a sequence {un} of vectors in V converges to a vector u E V with respect to the norm II .11, written as lim Un = U n-+oo or simply as Un  u, if the norm Ilu n - ull converges to 0 as n  00. The element u is called the limit of the sequence {un} in V. Remember that the limit u must also be an element of V. Thus, from the definition of the normed space, we have lim Un = U <==} lim lIun - ull = 0 n-+oo n-+oo which is equivalent to say that for each f > 0 there exists a natural number N = N(f) such that lIun - ull < f whenever n > N. Note that the convergence depends on the choice of the norms: A given sequence of vectors in V may converge with respect to one norm but not with respect to another norm. Such a situation can happen in an infinite dimensional vector space. Is this possible in the finite dimensional cases? For a detailed discussion on this topic we refer to Section 5.1. The following results are easy to prove. 3.8. Proposition. Let V be a normed space over F. Let Un, v n , U, v E V and An, A E F for n = 1,2, . ... Suppose that lim Un = u, lim V n = v and lim An = A. n-+oo n-+oo n-+oo Then we have (i) u is unique, (ii) {un} is bounded, (iii) lIunll  lIuli as n  00, 
148 Chapter 3: Normed Spaces (iv) AnUn  AU as n  00, (v) Un + V n  u + V as n  00. Proof. (i) If Un  U and Un  u', then the uniqueness part follows from lIu - u'li = lI(u - un) - (u' - un)1I < lIun - ull + lIun - u'li. (ii) As Un  U implies lIun - ull  0, there exists an M > 0 such that lIun - ull < M for all n so that lIunll = lIun - u + ull < lIun - ull + lIuli and hence (ii) follows. (iii) By the triangle inequality (N3), we note that every normed space V satisfies the inequality Iliull - IIvlll < lIu - vii for each u, v E V (This inequality also follows from (2.2) with u = x - y and v = z - y). H {un} is a sequence in V such that Un  u, then from the above inequality we have IlIunll - lIulll < lIun - ull and the desired conclusion follows from this inequality. (iv) This part follows from IIAnUn - Aull < II(An - A)U n + A(U n - u)1I < IAn - Aillunil + IAIliu n - ull. (v) Here the conclusion follows from lI(u n + v n ) - (u + v) II < lIun - ull + IIV n - vII. . 3.9. Corollary. Let V be a normed space and Y C V, a linear subspace. Then the closure Y is a closed linear subspace of V. Proof. Let Y be a subspace of V. We show that Y is a (linear) subspace. For this, we let x, y E Y and A E F. Then, by Proposition 2.53, there are sequences {xn} and {Yn} in Y such that X n  x and Yn  y. By Proposition 3.8, it follows that AX n ...+ Yn  AX + y. Since Y is a linear subspace, we have AX n + Yn E Y for all n. Therefore, letting n  00, we note that AX+Y E Y (see Proposition 2.53) and therefore, Y is a subspace of V. . 3.10. Geometry of norms. As with metric spaces, it is possible to understand the concept of norms from a geometrical point of view. For 
3.1. Properties of Norm 149 instance, the open and closed balls (with center a and radius 6 > 0) in a normed space (V, II · II) are defined by the sets B(a;6) = {x E V : IIx - all < oJ, B[a;6] = {x E V: IIx - all < 6}. In particular, the open and the closed unit balls in V are then defined by B := B(O; 1) = {x E V: Ilxll < 1}, B := B[O; 1] = {x E V: Ilxll < 1}, respectively. It is important to note that in a normed space, B[a; 6] = B(a; 6), where B(a; 6) denotes the closure of the open ball B(a; 6). It can be easily shown that B(a; 6) C B[a; 6] since B(a; 6) C B[a; 6] and B[a; 6] is a closed set. Indeed, since B(a; 6)' is the union of B(a; 6) and its limit points and since B(a; 6) C B[a; 6], it suffices to prove that all the limit points of B(a; 6) belong to B[a; 6]. Assume that x is a limit point of B(a;6). Then there exists a sequence {x n } in B(a;6) such that X n  x. Thus, IIx - all < IIx - xnll + IIxn - all < Ilx - xnll + 6. As IIxn - xII  0, this implies that IIx - all < 6, i.e. x E B[a; 6] which proves that B(a; 6) c B[a; 6]. To prove the reverse inclusion, we choose x E B[a; 6]. Then, IIx - all < 6. If Ilx - all < 6, then x E B(a; 6) C B(a; 6). Thus, it suffices to consider the points x E B[a; 6] with IIx - all = 6. Define X n = (1 -  ) x + : . Then, for all n, we see that IIxn - all = (1 -  ) (x - a) = (1 -  ) 8 < 8. Therefore, X n E B(a; 6) for all n, and 1 6 IIxn - xII = -lix - all = -  0 ' as n  00 n n so that X n  x as n  00. Thus, each x E B[a; 6] is a limit point of a sequence in the open ball B(a; 6) which shows that B[a; 6] is contained in the closure of B (a; 6). 
150 Chapter 3: Normed Spaces 3.11. Proposition. A normed space X is homeomorphic to the open unit disc B = {y EX: lIyll < I}. Proof. For x EX, Y E B, we consider x y f(x) ::: 1 + IIxll and g(y) = 1 -llyll ' Then, we have 1 1 · IIf(x)1I < 1, 1 -lIf(x)1I = 1 + IIxll ' 1 + IIg(y)1I = 1 -llyll I(x) g(y) · g(f(x)) = 1 -lIf(x)1I = x, f(g(y)) = 1 + IIg(y)1I = y. Thus, I : X  B is bijective with 9 = 1-1. Since the norm II · II is a continuous function, X n  x implies that IIxnll  IIxli. In particular, I(x n )  I(x) so that I is continuous. Similarly, we see that 9 is continuous. Thus, I is a homeomorphism. - 3.12. Observations. In a normed space it is important to note the following observations: (a) If V :F {OJ and if II · II is a norm on V, then all · II is also a norm for each a > O. (b) When we work on more than one norm on the space V (see 3.5) that is under consideration, then we may write these norms by 11,111, 11.112, etc. Similarly when discussing more than one space with respect to the same norm, the associated norm with respect to the spaces V, W may sometimes be denoted by 11.llv, 11.llw respectively, if necessary. (c) From Proposition 3.8(iii), it follows t"hat the norm II · II : V  IRt , u  lIull, is a uniformly continuous function. . It is also important to observe that the interior of a set depends on the choice of the metric. For example, consider the closed interval J = [0, 1]. Then we have the following simple observation: (a) Int J = (0, 1), when J is considered as a subset of IR with usual metric. (b) Int J = J, when J is considered as a subset of IR with discrete metric. (c) Int J = 0, when J is considered as a subset of (IR2, 11.112). For normed spaces, we have the following important result concerning the interior of subspaces. 
3.1. Properties of Norm 151 3.13. Proposition. Every proper subspace of a normed space has empty interior. Proof. Let Y be a proper subspace of a normed space (X, 11.11). Assume on the contrary that, int Y # 0. Then there exists an element a E Y and an open ball B(a;6) c Y for some 6 > O. Then, for each 0 # x E X, we have 6 x y = a + 211xll E B(aj 6), i.e. lIy - all = 6/2 < 6, which shows that y E Y, and therefore, 2 6 (y - a)lIxll = x E Y. Thus, every x E X is a point of Y, contradicting the assumption that Y is proper subspace of X. . From the idea of the proof of Proposition 3.13, we can obtain the fol- lowing general property for normed spaces. 3.14. Corollary. A subspace Y of a normed space X is either dense (i.e. Y = X) or nowhere dense (i.e. Y does not contain an open ball which is equivalent to say that int Y = 0 ). Proof. Suppose Y # X. Then we must show that Y does not contain an open ball. Assume on the contrary that, B(a; 6) c Y for some 6 > 0 and for some point a E Y . Then, as in the proof of Proposition 3.13, it follows that every x E X is also in Y , contradicting the assumption that Y # X. . For a metric space (X, d), we have already shown that 0 and X are both open and closed. Further, every subset of a discrete metric space is both open and closed. However, for normed spaces we have the following precise information. 3.15. Proposition. In a normed space (X, II · II), the only subsets which are both open and closed are the empty set 0 and the whole space X. Proof. Suppose not. Then there exists a proper nonempty subset G of X which is both open and closed. Therefore, GC is also both open, closed and nonempty. Further, X = GuGc is a disjoint union of nonempty proper open subsets of X and so, X is disconnected. Let us take x E G and y E GC. Consider the map f : [0, 1]  [x, y], A r-+ AX + (1 - A) y , 
152 Chapter 3: Normed Spaces where [x, y] : = Lx,y = {Ax + (1 - A) Y : A E [0, I]} eX. (N ote that, in this proof [x, y] is not to be considered as closed interval unless X is a subset of IR). Clearly, f is continuous and, since [0, 1] is connected, it follows that [x, y] is connected. Therefore, [x, y] is either in G or in GC which is not possible because x E G and y E GC. Thus, we must have either GC = 0 or G = X. This contradicts our initial assumption. Thus, if G # 0 is a subset of X which is both open and closed, then G=X. Alternately, by writing [x, y] = ([x, y] n G) U ([x, y] n G C ), we can quickly see that [x, y] is disconnected, a contradiction. . From Proposition 3.15, we observe that every normed space is connected whereas a metric space is not necessarily connected as the discrete metric space demonstrates. 3.2 Convexity and Completeness Recall that every normed space is a metric space and is therefore a topo- logical space. Therefore, we can make use of some basic definitions and results from general topology to give a topological structure to the normed space. We shall later see that the concept of norm equivalence will be the . same from the point of view of continuity and convergence, see Proposition 3.8. First, we review some basic concepts from general topology: As dis- cussed partly in 3.10, the open ball, closed ball, sphere of center Xo with radius 6 > 0 are all easy to state from the respective definitions of these concepts on metric spaces. From Definition 1.47, we note that the concept of convexity does not depend on the vector space under consideration. On the other hand, every ball in a normed space, irrespective of whether it is open or closed, will depend not only on 6 and a, but also on the particular norm that is being considered on the space. As we see in Proposition 3.16 below, every ball in a normed space is convex for any choice of the norm. Consider the vector space V = IR2 and define { (lxlP + lyIP)l/p II (x, y)lIp = max{lxl,lyl} ifO<p<oo if p = 00, where (x, y) E IR2. We have already discussed the open unit balls, B(O; 1), for p = 1,2,00 (see Examples 2.46). They are precisely the interior of the curves r P for p = 1,2,00 in Figure 3.1. We observe that for p = 1 
3.2. Convexity and Completeness y 1 -1 -1 153 p= 00 p=3 p=2 p=l P -! -2 1 x Figure 3.1: Description of r p = 8(0; 1) for p = 1,2,00,1/2 and p = 00, the boundary of the unit ball in each case is a square, while for p = 2 it is a circle; in each of these three cases the interior of r p is convex. It is easily seen that, as p increases from 1 to 00, the open unit ball grows steadily as a convex domain. However, we note that if p E (0,1), the convexity conclusion fails. See Figure 3.1, for the case p = 1/2, where the region bounded by the curve rl/2 is obviously not convex. Note that the function 11(., .)llp for 1 < p < 00 defines a norm on }R2. What about for o < p < I? Finally, on }R2 , define a new function x _ { (lxl2 + lyI2)1/2 II( · y}1I - max{lxl.lyl} if xy > 0 if xy < O. Clearly, this function II · II defines a new norm on }R2 and, with respect to this norm, }R2 becomes a normed space. Now, the open unit ball B(O; 1) in this normed space is described as follows: (x, y) E B(O; 1) if (x, y) satisfies x 2 + y2 < 1 max{-x,y}<l max{x, -y} < 1 ifxy > O if x < 0, y > 0 if x > 0, y < o. 
154 Chapter 3: Normed Spaces y (-1,1) 1- - -- - "" " 1 " I '\ 1 \ \ I \ , \ 0 x 1 \ " I " I "'" "- _--_ - I - --.- (1, -1) Figure 3.2: The open unit ball for a new norm on ]R2 Geometrically (see Figure 3.2), it is clear that B(O; 1) is a convex subset of ]R2 . 3.16. Proposition. The open ball B(a; R) = {x: IIx-all < R} in a normed space is convex. In particular, the function f defined by f(x) = IIxll 1S convex. Proof. If x, y E B(a; R) and A E (0, 1), then IIAX + (1 - A)Y - all < Alx - all + (1 - A)lIy - all < AR + (1 - A)R = R and the conclusion follows. . Note that Proposition 3.16 does not hold in general in metric spaces. For example, consider the Frechet metric d on the space of all sequences of complex numbers considered in Example 2.18: d(z, w) = f 3- n IZn - wnl . - n=l 1 + IZn - wnl Let Z = {p,O,p,O,p,...} and w = {q,q,q,.. .}, where p and q are fixed positive real numbers with q E (0,3). (For example, z = {I + (_I)n-l }nl 
3.2. Convexity and Completeness 155 and W = {I, 1, . . . , } so that p = 2 and q = 1). Then pool d ( z , O ) =  . 1 + p  3 2n - 1 n==l But, since (1 - x)-l = E:=o x n , 00 1 00 1 00 1 3 9 3 L 3 2n - 1 = L 3 n - L 3 2n = 2 - 8 = 8 n=l n=O n=O and therefore, 3p d(z,O) = 8(1 + p) ' Similarly, we have q  1 q d(w,O) = 1 + q  3 n = 2(1 + q)' Thus, z, W E B[O; q/2(1 + q)] whenever 3p q . 3p 4q 8(1 + p) = 2(1 + q) , I.e. q = p + 4 or p = 3 _ q. However, AZ + (1 - A)W ft B[0;6] for all A E (0,1) with 6 = q/2(1 + q). If we let ( = AZ + (1 - A)W = {(n}nl, then we have { (I - A)q (n = AP+ (1- A)q for n > 2 is even for n > 1 is odd so that (l-A)q ( 1 ) Ap+(l-A)q ( 3 ) d((,O) - 1+(1-.\)q 8 + 1+.\p+(1-.\)q 8 1 1 3 - 2 8[1 + (1 - A)q] 8[1 + AP + (1 - A)q] - f(A), say. It is a simple calculation to verify that q d((,O) > 2(1 + q) for all .\ E (0,1) which shows that ( ft B[O; q/2(1 + q)]. In fact, in the special values p = 2, q = 1, one can quickly obtain that 1 - A 3(1 + A) 1 d((,O) = 8(2 _ .\) + 8(2 + .\) > 4 {=::} .\(1-.\) > 0, 
156 Chapter 3: Normed Spaces which shows that AZ + (1- A)W ft B[O; 1/4] whenever Z,W E B(O; 1/4). To complete the proof for other values of p > 0 and q E (0,3), we must verify that 1(>") > 1(0) for all >.. E (0,1). Since p = 3 4q > 0, we have -q 134 1(>") > 1(0) {=::} 1 + (1 - >..)q + 1 + >"[4q/(3 - q)] + (1 - >..)q < 1 + q l+q 3 4 {=::} 1 + (1 - >..)q + 1 + (>..q/(3 - q)) < {=::} { l+q I } 3 { 3-q I } 0 1 + (1 - >..)q - + 3 - q + >..q - < {=::}A { 1 3 } 0 q 1 + (1 - A)q - 3 - q + Aq < qA(l - A) > 0 {=::} [1 + (1 - A)q][3 - q + Aq] {=::} A(l - A) > 0, since A > 0 and q E (0,3). A similar conclusion may be drawn with 2- n in place of 3- n in the metric expression of d(z, w). 3.17. Notion of Banach spaces. The sequence {un} in a normed space (V, II · II) is called a Cauchy sequence if for every € > 0 there exists a positive integer N = N(€) such that lIu m - Un II < € whenever m,n > N. We note that Cauchy sequences play a vital role in the theory of normed spaces. The normed space (V, II . II) is said to be complete if V is complete as a metric space with the metric d(u,v) = lIu - vII for u,v E V. In other words, (V, 11.11) is called a complete normed space if for every sequence {un} in V such that d(un,u m ) = lIu n - umll  0 as n,m  00, there exists an element u E V such that d(un,u) = lIun - ull  0 as n  00. A Banach space is a complete normed space. 20 Note that a given sequence {un} in V may be a Cauchy sequence with respect to one norm but not necessarily with respect to another norm. 3.18. Example. Recall that Q is a vector space over the field Q itself and one can easily see that Q is a normed subspace of IR with the 20The name Banach is referred to the famous Polish mathematician S. Banach who extensively investigated the properties of these spaces from the year 1922 onwards. 
3.2. Convexity and Completeness 157 Euclidean norm: IIxll = Ixl for x E Q. Further, the sequence { 0.1 , 0.101, 0.101001, . . . , } is a Cauchy sequence of rational numbers converging to a limit which is a irrational number. From this observation it follows easily that Q is not a Banach space. . .A reformulation of Proposition 2.102 for normed spaces is the following. 3.19. Proposition. Every convergent sequence in a normed space is a Cauchy sequence. Note that every Cauchy sequence in a normed space is bounded. How- ever, as in Proposition 2.102, the converse of the Proposition 3.19 is not true in general. Thus, a Cauchy sequence need not always be a convergent sequence in a normed space (V, 11.11) which is not complete as we see in the following example. 3.20. Example. Consider the vector space Coo = {{zn} E 1 00 : {zn} has only a finite number of nonzero terms}. We continue to use this notation throughout the book, see also page 167. Then Zn = {1, 1/2, 1/3, . . ., 1/n, 0, 0, . . .} E Coo, for each n E N and that {Zn}nl is a Cauchy sequence with respect to the supnorm. For n > m, it follows that Zn - Zm = { o, 0, .. . , 1 l ' 1 2 '''''.!.' 0, 0, .. . } m+ m+ n so that { 1 1 1 } 1 IIZn - Zmlloo = sup 1 ' 2 " . . , - = 1  0, m+ m+ n m+ as m  00. But Zn  Z = {1, 1/2, 1/3,..., 1/n,...} ft Coo. Note that Coo is a subspace of both 1 2 and 1 00 . Also, Z E 1 2 C 1 00 . . By Proposition 2.104, we see that if a Cauchy sequence in a normed space has a convergent subsequence then the whole sequence is convergent. By Proposition 3.8, we observe the following: 3.21. Corollary. In a Banach space, a sequence is convergent iff it is Cauchy. 
158 Chapter 3: Normed Spaces We recall the notion of convergence of a series from 2.17 for the normed space settings. Given a sequence {Xn}nl in a normed space X, ",.e may form the sequence of partial sums n Sn = LXk. k=l If Sn  x in X as n  00, we say that the series converges to x or has a sum x. The (unique) limit x = lim n -+ oo Sn is called the sum of the series and we write 00 LXk = x. k=l If E %" 1 Xk = x for some x EX, then we say that the series E %" 1 Xk is convergent in X or that the series E  1 Xk converges in X. If the numerical series E  1 Ilxkll is convergent, then we say that the series E  1 Xk is absolutely convergent. Once again, we remark that a convergent series in a normed space need not be absolutely convergent. 3.22. Example. The series 00 L(-1)n+1.! n n=l converges but does not converge absolutely. The convergence of the series follows directly from the alternating series test. To recall (probably done in a Calculus course) that the series is not absolutely convergent, consider 1 f(x) = - x with Xk = k + 1, Xk = Xk - Xk-l, for 1 < k < n and the partition p = {XO, Xl, . . . , x n }. Since 1 M k = sup f(x) = k ' XE[Xk-t,XIc] we have (n+l n n 1 logn < log(n+ 1) -log1 = 11 f(x) dx < U(P, f) = L MkXk = L k 1 k=l k=l 00 1 so that, since logn  00 as n  00, the series L n does not converge. n=l 00 Alternatively, to show that the harmonic series L  is not convergent, it n=l 
3.2. Convexity and Completeness 159 suffices to show that the sequence of partial sums is not Cauchy. For this, we notice that 1 1 1 n 1 S2n - Sn = n + 1 + n + 2 + · . . + 2n > 2n = 2 and therefore, {sn} does not converge. 00 Moreover, for p > 1, the series '" -.!:.. is convergent because L....J n P . n=l {n. dx _ 1 [ 1 _ 1 ]  1 as n  00. J 1 x P - 1 - p n P - 1 p - 1 How about 0 < p < I? . 3.23. Example. We provide some simple examples to demonstrate the facts described above for the convergence of 'series. (i) Consider the Banach space X = IR with the Euclidean norm. If we set Xk = (k 2 + k)-l, then n n ( 1 1 ) 1 Sn =  Xk = L - - = 1 - -  1  k=l k k + 1 n asnoo so that 00 1 . L k 2 k =1. k=l + (ii) Consider the Banach space X = C with the Euclidean norm and let Zk = a k - 1 , where a is some fixed complex number with lal < 1. Then n I n L n 1 -a Sn = Zk = 1 + a + · · · + a - = I-a k=l which shows that 1 . Sn - 1 -a = lain  0 as n  00 for lal < 1. 11- al Therefore, 00 1  a k - 1 = (Ial < 1).  I-a k=l 
160 Chapter 3: Normed Spaces (Hi) If we consider the space X = C[0,3/4] with respect to the supnorm and fk(X) = xk, kEN, then we have the estimate n L k X X - I-x k=l ( 1 - x n ) X - x I-x -1-x Ixln+l I-x (3/4)n+l < 1/4 and therefore, E  1 fk(X) converges to the function x/(1 - x). . Note that, an absolutely convergent series need not be convergent in normed spaces, in general. For example, consider the space of all polynomi- als p defined on [0, 1] with respect to the supnorm IIplioo = sUPxe[O,l] Ip(x)l. The series E  °  is absolutely convergent, but not convergent in the space of all polynomials with respect to the supnorm since eX is not an ele- ment of this space. Also, we can construct examples by choosing a sequence {fn(X)}nl of functions in C[O, 1] such that IIfnll =  for each n > 1 and that E  1 fn is not continuous. However, from our next result, we see that every absolutely convergent series in a Banach space is convergent. 3.24. Proposition. A normed space X is Banach iff every abso- lutely convergent series in X is convergent. Proof. Assume first that X is a Banach space. If the absolute conver- gence of a series E  1 x n is assumed for some (countable) sequence {x n }, then E  lllxnll < 00. Let Sn = EZ=l Xk. Now, since n n IIsn - smll = L Xk < L Ilxkll ---+ 0 as n > m  00, k=m+l k=m+l it follows that the sequence {Sn} of partial sums is a Cauchy sequence in X. As X is Banach, the sequence {sn} has a limit in X and hence the series E  1 xk is convergent. Conversely, suppose that X is a normed space and every absolutely convergent series in X is convergent. Given a Cauchy sequence {x n }, we can choose an increasing subsequence {nk} keN of positive integers such that Ilxn - xmll < 2- k as n,m > nk. Thus, the series X nl + (x n2 - x n1 ) + (x n3 - x n2 ) + · · · 
3.2. Convexity and Completeness 161 converges to some x (Note that the k th partial sum of this series is x n ",). Indeed, 00 IIx n1 + (x n2 - x n1 ) + (x n3 - x n2 ) + ...11 < IIxn11l + L 2- k < 00 k=l so that the series 00 X n1 + L(Xn"'+l - x n ",) k=l is summable absolutely, and therefore converges (by hypothesis). As the k th partial sum of this series is X n "', we have x n1c  x. Since {xn} is Cauchy and has a convergent subsequence {x n1c } which converges to x, by Proposition 2.104, X n  x and the space is complete. _ 3.25. Definition. A family {Xa}aEA of elements in a normed space is called s'Ummable to x, written EaEA Xa = x, if for each f > 0 there exists a finite subset J(f) of A such that if J is a finite subset of A containing J(f) then one has L Xa - x < f. aEJ The family {Xa}aEA is said to be absolutely summable if {lixall}aEA is summable in III Note that if :F denotes the collection of all finite subsets of A, then EaEA Xa is nothing but LX a = sup LXa. aEA JEF aEJ In the case A = N, this is equivalent to the convergence of the series. We state the following results without proof as it is routine. 3.26. Proposition. If {Xa}aEA and {Ya}aEA are two summable families in a normed space X with sums x and y respectively, then {xa + Ya}aEA is a summable family with sum x + y. 3.27. Poposition. In a Banach space (X, 11,11), the family {Xa}aEA is summable to x iff for each f > 0, there xists a finite subset J(f) of A such that, for every finite subset J of A distinct from J(f), one has L Xa - x < f. aEJ 
162 Chapter 3: Normed Spaces 3.3 The Banach Spaces lP(n) (1 < p < 00) The norms of the spaces lP(n) and lOO(n) (these norms are usually called lP-norm and loo-norm, respectively) are given by Ilzllp = ( n ) lip  IZkIP max IZk I lkn if 1 < p < 00, if p = 00, where Z = (Zl, Z2, . . . , zn) E en. The triangle inequality (N3) is a con- sequence of the Minkowski inequality (see Lemma 2.26). Note that (see 2.32) max IZkl = lim Ilzllp. lkn p-+oo The loo-norm II . 1100 on the space lOO(n) is called the supnorm or the max- imum norm or uniform norm. We note that 1 2 (n) is the n-dimensional unitary space (also called complex n-space) and, when we deal with IRn instead of en, this is simply the n-dimensional Euclidean space (also called real n-space) (see 2.32). The fact that Ilzlloo, Z = (Zl, Z2,... , zn), is a norm follows easily. Indeed, (i) We note that Ilzlloo = 0 <==} IZkl = 0 for all kEN <==} Zk = 0 for all kEN <==} Z = o. (ii) IIAZlloo = max IAZkl = IAI max IZkl = IAlllzlloo. lkn lkn (iii) The triangle inequality (N3) follows from Lemma 2.28(i). Now, to show that it is a Banach space, we assume that {(k} is a Cauchy sequence in lOO(n), where (i = (Zl (i), . . . , zn(i)). Then for all i, j > 1, we have the inequality IZk(i) - zk(j)1 < lI(i - (jlloo = max IZk(i) - zk(j)l, 1 < k < n, - lkn so that {zk(i)}il is a Cauchy sequence in F (IR or C) for each k with 1 < k < n. Since 1F (C or IR) is complete, the classical Cauchy convergence theorem implies that {zk(i)}il is convergent and hence zk(i)  Zk for each fixed k = 1,2,..., n, as i  00. If ( = (Zl,.'., zn), then lI(i - (1100 = sup IZk(i) - zkl  0 as n  00 lkn 
3.3. The Banach Spaces IP(n) (1 < p < 00) 163 and therefore, we obtain that (i  ( as i  00. Hence lOO(n) is complete. More precisely, we can write Jim lI(i-.(lloo = Jim { max IZk(i) - Zkl } = max { .lim IZk(i) - zkl } = 0 -+oo ---too l:5k:5n l:5k:5n -+oo and conclude that lOO(n) is complete. Let CF(X) denote the space of all continuous functions on an arbitrary compact set X over the field F. Then, we see that the space lOO(n) is a special case of the CF(X) space, where X = {I, 2, . . . , n} is a discrete compact space. In the above discussion, we have defined II . lip under the assumption that 1 < p < 00. Suppose that 0 < p < 1 and use the same definition of II · lip on en. Then, we see that II . lip, 0 < p < 1, does not satisfy the triangle inequality (N3) and is therefore not a norm on en for 0 < p < 1, unless n = 1. This observation is easy to verify, for example if n = 2, u = (1,0) and v = (0, 1) then lIuli p = 1 = Ilvll p , Ilu + vll p = 11(1, 1)llp = 2 l/p > 2, as p < 1, and therefore (N3) cannot hold for 0 < p < 1. We now show that the space IP(n) = (en, II · lip) for 1 < p < 00 is complete and hence a Banach space. This is in fact equivalent to show the completeness of this space with respect to the metric dp(z, w) of Examples 2.32 defined by ( n ) IIp dp(Z, w) = dp(z - w, 0) = Ilz - wll p = E IZk - wkl P · k=l To see this, we suppose that 1 < p < 00. Let {(i}, (i = (zl(i),...,zn(i)), be a Cauchy sequence in IP(n) with respect to the norm (called IP-norm or p-norm on en) ( n ) IIp IIzllp =  IZkl P · Then for each 1 < k < n and for all i,j > 1, we have IZk(i) - zk(j)1 < dp((i, (j) := II(i - (jllp  0 as i,j  00 so that {zk(i)}il is a Cauchy sequence in F (C or IR) with the usual metric, for every 1 < k < n. Since F (C or IR) is complete, {zk(i)}il is convergent and thus it has a limit Zk, where Zk = limi-+oo zk(i). Therefore, if ( = (Zl, . . . , zn) then using the inequality in 2.32 we deduce that lI(i - (lip < n 1 / P d oo ((i. () = n 1 / p ln IZk(i) - zkl. which shows that (i  ( as i  00. Hence lP(n) is complete. 
164 Chapter 3: Normed Spaces 3.4 The Sequence Spaces lP (1 < p < 00) For sequences Z = {Zn}nl and w = {Wn}nl belonging to the space X of the space of all sequences of complex numbers, we consider the associated metric d p (., .), 1 < p < 00, defined in 2.32 so that the corresponding norm is given by dp(z,O) := IIzlIp, 1 < p < 00. The III-norm (or simply p-norm) and loo-norm (or simply supnorm) are given by IIzllp = ( 00 ) l/p  IZkIP sup IZk I lk<oo ifl < p<oo if p = 00 Now, we define IP = {z = {Zn}nl : IIzllp < oo}. where Z = {Zn}n>l E lP. Now it should be pointed out that, in general, (E  1 IZkIP)l/P n;ed not always be finite for 1 < p < 00. However, in the definition of lP-spaces, we have restricted ourselves to the case for which this sum is finite. The norm properties are left to the reader for verification. We now show that the space lP for 1 < p < 00 is complete and hence a Banach space. Let {(i}, (i = {zk(i)}kl E lP for i = 1,2,..., be a Cauchy sequence in the normed space (lP, II. lip). Then for each kEN and for all i,j > 1, we have (3.28) IZk(i) - zk(j)1 < dp((i,(j) = dp((i - (j,O) = lI(i - (jllp ---+ 0 as i, j  00. Set Zk = limj-+oo Zk(j) for each k > 1. This provides a sequence ( = {Zk}kl' To complete the proof, we need to answer the following two questions: . Is ( E lP for 1 < p < oo? . Is lI(i - (lip  0 for 1 < p < oo? We first consider the case p = 00. Since {(i} is Cauchy in 1 00 , inequality (3.28) gives the following: Given € > 0 there exists a natural number N = N(k) such that (3.29) IZk(i) - zk(j)1 < doo((i,(j):= lI(i - (jlloo < € whenever i,j > N. Now fixing i and letting j  00, in (3.29), we see that IZkl < IZk(i) - zkl + IZk(i)1 < € + IZk(i)1 < € + lI(ilioo for i > N(k). Note that {(i}, being a Cauchy sequence, is bounded (see Proposition 2.103). Thus, ( E 1 00 . Further, by (3.29), we have lI(i - (1100 = sup IZk(i) - zkl  0 as i  00. k 
3.4. The Sequence Spaces lP (1 < p < 00) 165 Next, we consider the case 1 < p < 00. We shall first show that ( E lP and then lI(i - (lip  0 as i  00. We note that if M is a positive integer, then M M E IZkl P = .lim E IZk(j)IP and J-+OO k=l k=l M E IZk(j)IP < II(jll. k=l Thus, since {(i} is a Cauchy sequence in lP, given f > 0 there exists an N such that M E IZk(i) - zk(j)IP < lI(i - (jll < f whenever i,j > N. k=l Now, fixing i and letting j  00, we obtain M E IZk(i) - zkl P < f whenever i > N. k=l Since this is true for every M, on letting M  00, we get lI(i - (II < f for i > N or equivalently, lI(i - (lip  0 as i  00. Now, by the triangle inequality in lP (see Lemma 2.26), 1I(lIp = ( 00 ) IIp  IZkIP ( 00 ) IIp ( 00 ) IIp < h IZk - zk{i}IP + h IZk(i}IP , - II( - (illp + lI(illp < flIp + lI(ilip for i > N, shows that ( E lP. Hence, lP is a Banach space for 1 < p < 00. Thus, we arrived at the following result. 3.30. Theorem. The space lP is a Banach space for 1 < p < 00. We note that the elements of the Banach space 1 00 are called bounded sequences in the field IF whereas the elements of the Banach space 1 1 are called absolutely summable sequences in IF. Similarly, the elements of the Banach space 1 2 are called absolutely square summable sequence whereas the elements of the Banach spaces lP for each p in (1,00) are called p-th summable sequences in IF. 
166 Chapter 3: Normed Spaces 3.31. Theorem. For 1 < p < q < 00, we have the strict inclusion IP C lq. .... Proof. Let z E IP (z :j:. 0), where 1 < p < q < 00. If we define z e = II z II p = {a 1 , a2, · · .}, then lIeli p = 1, by (N2); that is E  1 lanl P = 1 and therefore, we have lanl < 1 for each n > 1. If q < 00, then -Ianl < 1 implies that lanl' > lanl q , Le. 1 = Ilell > lIell, which gives Ilell q < 1, i.e. Ilzllq < Ilzllp. Moreover, equality can occur in the last inequality only if z is a multiple of en for some n; that is, z can have at most one nonzero term. Here en E 1 00 where only the n-th coordinate is 1 and others are zero. If q = 00, then the desired inclusion is clear. Indeed, as Ian I < 1 => sup Ian I < 1, nl by (N2), it follows that IIz II 00 IIzllp = II ell 00 < 1, i.e. IIzlloo < IIzllp. Thus, for 1 < p < q < 00, we have shown that IIzllq < IIzllp and hence the inclusion lP C lq. Finally, in order to prove that the inclusion is proper, we choose a point s in the open interval (p, q), q < 00. Then the element Zo given by Zo = {n-l/S}nl belongs to ,q because 00 IIzoll = E n- q / s < 00. n=l On the other hand, 00 IIzolI = E n- p / s n=l which is divergent. Note that Zo = {(I/k 2 10g 2 k)l/ q } 
3.4. The Sequence Spaces lP (1 < P < 00) 167 also serves our purpose if q < 00. If q = 00, then in this case we see that the constant sequence {Zk}, Zk = C :j:. 0, belongs to 1 00 but not to lP for each P < 00. The same holds for the sequence {Zk}, where Zk = 1/ log k for k > 2, whose term approaches zero as k  00. Hence, the spaces lP and ,q for p :j:. q are not equal. _ 3.32. Corollary. If Z = {Zk}kl belongs to lP for some p < 00, then Ilzlloo = lim p .-+ oo Ilzllp. Proof. If Z E lPo, then Z E lP for Po < P < 00. Since IZkJ  0, there exists a largest IZkl, say IZNI, so that Ilzlloo = IZNI. Note that if ZN = 0, then Z = 0 and therefore, the result follows. So, we may assume that ZN :j:. O. Further, since IZk/ZNI < 1 and Ek IZklP < 00, it follows that the series E IZk/ zNIP converges. Hence there exists c > 0, independent of p, as II. lip < II. II Po for Po < p < 00, such that E IZk/ZNIP < c. Therefore, we have IZNIP < IIzlI = IZNIP E IZk/ zNI P < CIZNIP and the conclusion follows if we take p-th root on both side of the last inequality (note that c is independent of p) and allow p  00. _ 3.33. Remark. For 0 < p < 1, define (IP, d) as in Remark 2.36. Then (IP, d) is a metric space but not a normed space. . There are a few important subs paces of 1 00 , namely c, Co and Coo, where c - {z = {Zn}nl E 1 00 : lim n -+ oo Zn exists and is finite} , Co { z = {Zn}n>l E c: lim Zn = O } - n.-+oo and Coo C ' 00 is the vector space of all finitely supported sequences {Zn}nl, Le. "support [{ Zn}nl] = {n : Zn :j:. O}" is finite. The supnorm on Coo is (then Ilzlloo = maxlznl, for Z = {Zn}nl E Coo. nEN With the notation lP = lP(N) and ' 00 = lOO(N) in mind, we can introduce more general Banach spaces lP(A) and Co(A) for an abstract set A, where lP(A) = { f : A  F: E If(a)IP < oo } aEA with the norm (EaEA If(o:)IP)l/ P . Note that E If(a)IP = sup { E If(a)IP : J a finite subset of A } . aEA aEJ 
168 Chapter 3: Normed Spaces In view of this definition, it can be shown that Co(r) and lP(r), p E [1,00], are nonseparable for r uncountable (see Examples 2.74). 3.34. Example. Let X = lP (1 < p < (0) or Co. Given a sequence {Zn}nl in the normed space X, we have the representation 00 {Zn}nl = L Znen n=l where {en}n>l is the usual sequence of standard unit vectors in X. Note that the abo;e representation cannot be extended to 1 00 , for if {zn} E 1 00 is such that Zn does not tend to zero, then the corresponding sequence of partial sums given by n Sn = LZkek k=l is not Cauchy and therefore, cannot converge. . 3.35. Corollary. The subspaces c and Co are closed subspaces of 1 00 (hence are Banach spaces), but Coo is not complete. Proof. First, we want to prove that the space c is a closed subspace of 1 00 . This follows from the fact that the uniform limit of convergent sequences is a convergent sequence. Indeed, in order to prove that the uniform limit of a convergent sequences in c is convergent, we consider a sequence {(i}, where (i = {zk(i)}kl E c, that converges to ( = {Zk}kl E 1 00 . This means that for every € > 0 there exists an JV E N such that lI(i - (1100 < €/3 for all i > N. In particular, this in turn implies that for each kEN and for each given € > 0 and for all i > N, (3.36) IZA,(i) - zkl < lI(i - (1100 := sp IZk(i) - zA,1 < ; and then, fix an i satisfying the last inequality. For such a fixed i, since {Zk (i)} kl E c is a convergent sequence, there exists an N 1 such that IZp(i) - zq(i)1 < ; for all p, q > N 1 . Also, for p, q > N 1 , Izp - zql - Izp - zp(i) + zp(i) - zq(i) + zq(i) - zql < Izp - zp(i)1 + IZp(i) - zq(i)1 + IZq(i) - zql < 211(i - (1100 + IZp(i) - zq(i)1 € € < 2 3 + 3 = €, by (3.36), 
3.4. The Sequence Spaces lP (1 < p < 00) 169 showing that {Zk} is a convergent (scalar) sequence (since every Cauchy sequence in C is convergent). The closedness property of c follows. Propo- sition 2.109 immediately yields the fact that "a closed subspace of a Banach space is Banach". This observation shows that c (with X = 1 00 and S = c in Proposition 2.109) is a Banach space under the supnorm. The same reasoning gives that the space Co is Banach with respect to the supnorm. Finally, we recall from Example 3.20 that Coo is not complete. For this, it suffices to show that Coo is not closed in Co. Consider the sequence {Zn}nl in Coo, where Zn = {1,1j2,1j3,...,1jn,0,0,...} E Coo for each n = 1,2,.... Then, it is Cauchy (see Example 3.20) with respect to the supnorm and, therefore {Zn} converges to Z = {1, 1 j 2, 1 j 3, . . . , 1 j n, . . .} E Co \ Coo. The fact that Z E Co follows from the completeness of the space Co and IIZn-Zlloo= { o,o,..., 11 ' 1 2 ,... } n+ n+ 00 1 1  0 as n  00. n+ Hence, we conclude that Coo is not complete. Also, we observe that Zn  Z in 1 00 and, since '00 1 IIZn - ZII = I: k 2  ° as n  00, k=n+l Zn  Z in 1 2 . This observation shows that Coo is not a closed subspace of 1 2 and ' 00 although Coo is a subs pace of both 1 2 and 1 00 . . 3.37. Example. We provide one more example of an incomplete space, this time as a subspace of 1 2 . We consider Coo as a subspace of 1 2 . Now, with the understanding that Coo = (Coo, II · 112), Coo is again a vector space consisting of all finitely supported sequences from 1 2 . Clearly, with the norm inherited from 1 2 , Coo is in fact a normed space. However, Coo is not complete. For this, it suffices to consider the sequence {Zn}nl in Coo, where . Zn = {1,1j2,1j2 2 ,...,1j2 n ,0,0,...} E Coo for each n = 1,2,.... However, we remark that the sequence considered in Corollary 3.35 can also be used to show that Coo is not complete. Now, for n > m, Zn - Zm = { 0,0, .. · , 2 ' 2H ' · .. , 2n1_1 ,0,0, .. .} 
170 Chapter 3: Normed Spaces so that 00 1 1 1 IIZn - Zmll < L 2 2k = 34 m - 1 ---+ 0 as m  00 k=m so that {Zn} is a Cauchy sequence with respect to 2-norm in the Banach space 1 2 and therefore, {Zn} converges in 1 2 to Z = {1, 1/2, 1/2 2 , . . . , 1/ 2 n , . . .} E 1 2 \ Coo. Consequently, Coo is not complete. Is Zn  Z in lOO? . 3.38. Remark. By Theorem 3.31, we have the strict inclusion lP S; 1 00 for 1 < p < 00. By the definition of Co, we get c C 1 00 as the convergent sequences are bounded. Is it true that Co C lP for 1 < p < oo? Observe that, as n- 1 / p  0 as n  00, Z = {n-l/P}nl E Co. But 00 1 IIzllP =  - P n n=l which is divergent and therefore, z  lP. This observation shows that the inclusion Co C IP fails to hold for 1 < p < 00. However, if z = {Zn}n>l E Coo then Zn = 0 for large n so that E  1 IZn 1 P < 00 which shows that Z E lP for 1 < p < 00. Thus, the inclusion Coo C lP holds for all 1 < p < 00. By definition, Coo C 1 00 . . We end this section with the following theorem which gives a necessary and sufficient condition for a normed space to be complete. 3.39. Theorem. Let X be a normed space. Then X is a Banach space iff the unit sphere S(O; 1).= {x EX: Ilxll = 1} is a complete metric space (under the induced metric d(x, y) = IIx - ylI). Proof. (=»: Let {Yn} be a sequence in S(O; 1) such that Yn  y. Then, since the norm is a continuous function, Yn  y implies that IIYnll  Ilyli. Further, as IIYnll = 1 for each n, it follows that lIylI = lim llYn II = 1, Le. y E S(O; 1). n-+oo Therefore, S(O; 1) is a closed subset of the complete space X and hence, S(O; 1) is complete. 
3.4. The Sequence Spaces IP (1 < p < 00) 171 ( {::): Let the unit sphere S(O; 1) be complete, and let {xn} be a Cauchy sequence in X. Then IIxm - X n II  0 as n, m  00. If it has a subse- quence {x nk } converging to 0, then it is easy to see that the sequence {xn} converges to 0 because {xn} is a Cauchy sequence. So, without loss of generality, we assume that no subsequence of {xn} converges to o. Thus, there exists a N E Nand 6 > 0 such that Ilxnll > 6 for all n > N. This observation suggests that it suffices to consider {xn} for which X n :j:. 0 for each n and does not have any subsequence converging to O. Now, define a sequence {Yn} in S(O; 1) by X n Yn = IIxn II ' X n =I- O. Then IIYm-Ynll - xmllxnll - xnllxmll IIxn IIlIxm II (xm - xn)lIxnll + xn(llxnll-lIxmID Ilxnllllxmll < IIxm - xnll + IlIxnll-lIxmlll IIxm II IIxm II < IIxm - xnll + IIxm - xnll IIxm II IIxm II 211 x m - X n II IIxm II Since a Cauchy sequence of nonzero elements in a normed space is bounded, we can find a 6 > 0 and an M > 0 such that 6 < Ilxl: II < M for all k = 1,2, . ... Thus, IIYm - Ynll < 211 x m {}- xnll  0 as m,n  00 so that {Yn} is a Cauchy sequence in the complete space S(O; 1) and hence, there exists Y E S(O; 1) such that X n IIxnll = Yn  y. Recall that if {xn} is Cauchy, then so is the sequence {lIxnll} of positive real numbers, since IlIxm II - IIxn III < IIxm - X n II. Since IR is complete, IIxnll  a for some a E IRt and consequently, X n  ay E X. Hence, X is conaplete. _ 
172 Chapter 3: Normed Spaces 3.5 The Function Space C(X) Let us start with a well-known function 1 (1 - Z)Q+l ' By Taylor's theorem, we see that Z # 1. 1  r(n + a + 1) n (1 - z)aH =  r(o: + l)r(n + 1) z, Izi < 1, so that lirn fn(z) = (1 1) H ' n-+oo - Z Q  r(k + a + 1) k fn(z) = t:o r(o: + l)r(k + 1) z · As it stands this convergence is for each individual z such that Izl < 1 which we shall soon refer to as the pointwise convergence. To fix the idea, let us briefly discuss the topic of sequences of real valued functions since the same idea can be easily extended to complex valued functions or vector valued functions. Suppose that (X, d) is a nonempty metric space and suppose that, for each n E N, In is a function from X into III We are interested in the convergence of the sequence of functions {In}nl and, in particular, in determining what properties are possessed by a function that is the limit of a sequence of continuous functions. There are two main ways in which we can discuss the notion of convergence. One is through "pointwise convergence" and the other is through "uniform convergence" . 3.40. Definition. For functions In' I: X  ]R, we say that the se- quence {In} converges pointwise to the function I (or I is a pointwise limit of {In}) iff for each x EX, In(x)  I(x) as n  00. In other words, In  I if for a given x E X and an € > 0, there exists a natural number N = N(x, €) such that I/n(x) - l(x)1 < € whenever n > N. A simple observation from basic 'calculus' is that the notion of point- wise convergence does not preserve any important properties of In' e.g, see Example 3.45. 3.41. Definition. For functions In' I: X  ]R, we say that the sequence {In} converges 'Uniformly to I on X iff given € > 0, there is N = N(€) such that for all x E X I/n(x)'- l(x)I' < € whenever n > N. 
3.5. The Function Space C(X) 173 ",---.... " " , \ \ , "    /   I(x) In (X) ,  € ! ,'..../ ./ € ! '''''--'''''''' / , I \ , " '...._",  /    , '...._-' Figure 3.3: Description for uniform convergence The difference between these two definitions is that, in the uniform convergence, N depends only on €, meaning that the same N works for each x EX, whereas in the pointwise convergence, x E X is given and N can depend on x as well as €. Clearly, the definition of uniform convergence is equivalent to "A sequence of functions {In} defined on X converges uniformly to a function I on X iff In  I in the sup-metric/supnorm; that is sup I/n(x) - l(x)1 = II/n - 11100  0 as n  00, xEX see Figure 3.3." From the two definitions, it is clear that the uniform convergence is a stronger property than the pointwise convergence, as the following results and the examples show. 3.42. Proposition. The uniform convergence implies the pointwise convergence but not conversely. Recall that if X is a compact set and C(X) denotes the space of all continuous IF-valued functions on the compact space, then for each I E C(X), we define the norm, called uniform norm or supremum norm or simply supnorm, by 11/1100 = sup I/(t) I. tEX Now, for the convenience of the reader, we state and prove the following important result, known as the Uniform Convergence Theorem. 3.43. Proposition. The limit of a uniformly convergent sequence {In} of continuous functions on X is also continuous therein. Proof. Let {In} converge uniformly to I on X. We need to show that f is continuous at every point of X. Let a E X be arbitrary and € > 0 be given. Since In  I uniformly on X, there exists an N = N(€) such that for all x EX, I/n(x) - l(x)1 < €/3 for n > N(€). 
174 Chapter 3: Normed Spaces Continuity of In at a shows that there exists a 6 > 0 such that In(Bx(a; 6)) c By(ln(a); f/3). Then for x E Bx{a; 6) and all n > N{f), we have I/(x) - l{a)1 < I/(x) - In(x)1 + I/n(x) - In(a)1 + I/n(a) - l(a)1 showing that I{Bx{a; 6)) C By(/(a); f). The desired conclusion follows. . 3.44. Corollary. Suppose that a sequence of functions {In} con- verges pointwise on a X to I and that each In is continuous at a point a E X. If I is not continuous at a, then the sequence {In} does not con- verge uniformly on X to I. Most often this corollary is useful in a way to prove that the convergence of a given sequence is not uniform. 3.45. Example. Let X = [0,1] and In{t) = tn. Then X is compact and each In is continuous on X for n > 1. Clearly, {In} converges pointwise on X to the function f(t) = {  ifO < t<1 if t = 1, since, for t E (0, 1), In(t) = t n  0 as n  00 and In(l) = 1. Clearly, I is not a continuous function, so the convergence is not uniform by Corollary 3.44. Thus, point-wise convergence of a sequence cannot guarantee the continuity of the limit function. One can easily verify this fact by a direct method. Indeed, we first observe that I/n(t) - l(t)1 = t n for t E (0, 1) so that n log{l/t) Ifn(t) - f(t)1 < € <==} t < € <==} n > log(l/€) = N(t,€). But the quantity N(t, f) depends on t as well as f and hence, {In} may not converge uniformly on [0, 1]. Alternately, we may simply consider sup I/n(t) - l(t)1 = 1 tE[O,l] and it suffices to observe that this does not approach to O. . 
3.5. The Function Space C(X) 175 3.46. Example. Let n = {x = (Xl, X2, . . . , X n ) E }Rn : IIxll < 1} and a = (a1, a2,.. '.' an) E }Rn be a fixed element such that lIall < 1 (example lIall = 1/2, 1/3 etc). Here II · II denotes the Euclidean norm II · 112 on an (see Section 3.4) defined by ( n ) 1/2 nXll2 =  I X kl 2 · Define In: n c an  }R2 by In(x) = (lIxll n , (x · a)n) =: (g(x), h(x)) where x · a denotes the usual dot product on }Rn defined by n x.a=Lxkak. k=l Then the component functions 9 and h satisfy { 0 1 g(x) = IIxlin  for IIxll < 1 for Ilxll = 1 and, by Cauchy inequality, h(x) = (x.. a)n < Ilxllnllali n < lIalin  0 as n  00 for all x E n, respectively. Therefore, {In (x)} converges pointwise to the function { (0, 0) f(x) = (1,0) for IIxll < 1 for IIxll = 1, which is equivalent to writing f(x) = (cjJ(x), 0), cjJ(x) = { for IIxll < 1 }; for IIxll = 1. Note that cP : n  ]R is not continuous on n and therefore, f is not continuous on n. By Corollary 3.44, a sequence of continuous functions cannot converge uniformly to a discontinuous function. In view of this, we conclude that {In} does not converge to I uniformly on n. Alternatively, as { (lIxlln, (x. a)n) fn(x) - f(x) = (IIxlln _ 1, (x . a)n) for IIxll < 1 for IIxll = 1, 
176 Chapter 3: Normed Spaces it is simply sufficient to observe that sup II/n(x) - I(x) II > sup II/n(x) - l(x)11 > sup Ilxlin = 1 +0 IIxll::;l IIxll<l IIxll<l and the conclusion readily follows. . 3.47. Example. Define In : I = [0, 1]  }R by { 1-nx fn(x) = 0 for 0 < x < 1/ n for 1/ n < x < 1. Clearly, In E C[O,I] and In(x) converges pointwise on [0, 1] to the function f(x) = {  for 0 < x < 1 for x = 0, which is not continuous on [0,1]. . 3.48. Example. Define In : [0, 1]  }R by nx x fn(x) = n+x = l+x/n ' Then In(x)  x as n  00 and so {In} converges pointwise on [0, 1] to I(x) = x. Now, to verify the uniform convergence, we compute -x - n+x x 2 = cP(x), n+x say. I/n(x) - xl = nx Note that ,p'(x) = x(2n + x) (n + x)2 so that cP is increasing on [0, 1] and therefore, 1 ,p(x) < ,p(l) = n + 1  0 as n  00. Hence, we concltIde that the convergence is uniform on [0, 1]. . 3.49. Example. Let {} = {(x, y) E }R2 : 0 < x, y < I} and define In: {}  }R2 by In(x, ) = ( l+n y , l +nx ) = ( l/n+ y , x+l/n ) . y n + x n + y 1 + x/n 1 + y/n Clearly, {In} converges pointwise on {} to the function I(x,y) = (y,x). Moreover, ( 1 - xy 1 - x Y ) ( 1 1 ) In (x, y) - I (x, y) = , = (1 - xy) , n+x n+y n+x n+y 
3.5. The Function Space C(X) 177 and therefore, with respect to the Euclidean norm II . 112 on }R2, it follows that II/n(x, y) - I(x, y)ll - (l-xy)2 Cn:x)2 + (n:y)2 ) < (l-xy)2 ( 2 + 2 ) 2(1 - xy)2 n 2 So, sup IIfn(x,y) - f(x,y)1I2 < V2  0 as n  00 (x,y)EO n showing that {In} converges uniformly to I on n. . 3.50. Supnorm on continuous functions. The space CF[a, b], a distinguished member of the class of infinite dimensional spaces, is a well- known example of a function space. For each I E Cera, b], we consider the function II · 1100 defined by 11/1100 = sup I/(t)l. tE[a,b] It can be easily shown that II . 1100 defines a norm on Cera, b]. Indeed, the null vector (J := 0 is the function identically zero on [a, b]. It is obvious that 11/1100 > 0 and 11/1100 = 0 iff I(t) = 0 for all t E [a, b]. The axiom (N2) follows from the relation sup IA/(t)1 = sup IAII/(t)1 = IAI sup I/(t)l. Finally, we see that the triangle inequality (N3) follows from Lemma 2.28(ii). Thus, (Ce[a, b], II · 1100) is a normed space and it becomes a metric space with respect to the metric induced by the norm II · 1100.: doo(/, g) = III - glloo = sup I/(t) - g(t)l, for I, 9 E Cera, b], tE[a, . see Example 2.38. Further, Figure 3.4 shows that the graph of a function 1 E C[a, b] rises to a height h l and sinks to a distance h 2 so that the norm 11/1100 is defined by the larger of these two numbers. The height to which the graph rises and the depth to which it falls are both 0 whenever 11/1100 = 0, thus (N1) holds. If 11/1100 = Cl and IIglioo = C2, then from the definition of norms on 1 and 9 it follows that for all t E [a,b], I(t) E [-Cl,Cl] and g(t) E [-C2, C2] which give f(t) + g(t) E [-Cl - C2, Cl +C2]' This proves (N3). Finally, we show that the space (Ce[a, b], II · 11(0) is complete and hence a Banach space. Let {In} be an arbitrary Cauchy sequence in (Ce[a, b]., 11.11(0). 
178 Chapter 3: Normed Spaces s - II I f""'4 I  , I I I , I I 0 a b t Figure 3.4: Description for 11/1100 Thus, for each € > 0 there exists an N = N(€) such that for n,m > Nand all t E [a, b], we have f IIfm - fnlloo = sup Ifm(t) - fn(t)1 < 3 tE[ a,b] so that (3.51 ) € Ifm(t) - fn(t)1 < IIfm - fnlloo < 3 for all n,m > N. In particular, for each fixed t E [a, b], the inequality (3.51) shows that {fn(t)} is a Cauchy sequence in C. Since C is complete under the usual metric topology, there exists a function f : [a, b]  C such that fn(t)  f(t), for each fixed t E [a, b]. Letting m  00 in (3.51) shows that for each fixed -t- E [a, b] (3.52) € If(t) - fn(t)1 < 3 ' for all n > N. But N is independent of t and, as t is arbitrary, we may take the supremum in (3.52) to obtain (3.53) € Ilfn - flloo < 3 ' for all n > N. Thus, {fn} becomes a sequence of uniformly convergent (continuous) func- tions on the compact set [a, b] with limit f(t). Hence, fn -+ f uniformly, Le. IIfn - flloo  0 as n  00. It remains to show that f is a continuous function of t on [a, b]. Proposition 3.43 implies that the limit function J is continuous in [a, b]. Hence, (Cc[a, b], II · 11(0) is a Banach space Indeed, for the proof of the continuity of f, we must show that t n  t implies that f(t n )  f(t). By (3.53), given € > 0, there exists an N E N such that € IIf - fNlloo < 3 
3.6. Basic Results on LP-Spaces 179 and therefore, for n sufficiently large, I/(t n ) - l(t)1 < I/(t n ) - IN(tn)1 + I/N(t n ) - IN(t)1 + I/N(t) - l(t)1 < €/3 + I/N(t n ) - IN(t)1 + €/3 < €. So, I is continuous. 3.54. Proposition. Let Y = {I I I : [a, b]  [c, d] is continuous}. Then the subspace (Y, II .11(0) c (C[a, b], 11.1100) is complete. Proof. Since (C[a, b], 11.11(0) is complete, by Proposition 2.109, it suffices to show that Y is closed. If {In} is a sequence in Y such that In  I for some I E C[a, b], then for each fixed t E [a, b], I/n(t) - l(t)1 < II/n - 11100  0 as n  00. Further, since In(t) E [c, d] for each fixed t E [a, b], we must have I(t) E [c, d] as [c, d] is a closed subspace of the complete metric space III Thus, lEY so that Y is closed. _ Analogously, we can prove the following result. So, we omit the details. 3.55. Theorem. The supnorm makes the space CF(X) into a Banach space, where X is a compact space. 3.6 Basic Results on V-Spaces Let a and b satisfy -00 < a < b < 00. We denote by A2[a, b] the set of all real valued functions I on [a, b] that satisfy the condition lab If(xW dx < 00, where the above integral is in the sense of Lebesgue. We call A2[a, b], the space of all real valued square integrable functions on [a, b]. 3.56. Example. If I is bounded on [a, b], then there exists an M > 0 such that I/(x)1 < M for all x E [a, b] and therefore lab If(x)1 2 dx < M 2 (b - a) < 00. Thus, A2[a, b] contains all bounded functions on [a, b]. Consider the function 9 : [-1, 1]  IR defined by g(x) = { Ixl-l/ 0 2 for x E [-1,1] \{O} for x = o. 
180 Chapter 3: Normed Spaces Note that the corresponding integral is an improper integral. In fact if € > 0 is small, then we have r- f dx + 1 1 dx = 2 1 1 dx = 2 log ( ! ) J -1 Ixl f Ixl f x € which approaches 00 as €  O. Thus, 9 is not a square integrable function on [-1,1]. Hence, 9  A2[-I, 1]. However, it can be easily seen that the function cP defined by cP(x) = Ixl- 1 / 3 belongs to A 2 [-I, 1] although x = 0 is a singular point. . Let II. U2 be a function from A2[a, b] into IR defined by ( b ) 1/2 11/112 = 1 1/ (x W dx · It is easy to see that the space A2[a, b] forms a vector space with respect to the usual addition and scalar multiplication, but the function II · 112 does not define a norm on it because there exist functions 9 :j:. 0 in A2[a, b] such that IIgl12 = O. For example, if g(x) = {  for x E (a, b] for x = a then IIgl12 = 0, but 9 :j:. O. Observe that 9 is not continuous on [a, b] and 9 = 0 almost everywhere. 21 However, the Minkowski inequality enables us to show that 11.112 is a semi-norme<;l space and it does become a norm if we identify functions which differ only on a set of measure zero on [a, b], Le. two such functions are equal almost everywhere. Thus, by defining 11/112 = IIgll2 whenever I = 9 a.e., we can partition the space A2[a, b] into equivalence classes based on equality almost everywhere. Let L2[a, b] denote the re- sulting space of equivalence classes [I] of A 2 [a, b] (also called functions by convention). Thus, L2[a, b] is the set of all equivalence classes of functions [I] such that if 11, 12 E [I], then 11 = 12 a.e. With the addition and scalar multiplication defined as [I] + [g] = [I + g] and [0:/] = 0:[/], respectively, the set L2[a, b] forms a vector space. With the norm of the equivalence class [I] defined as ( b ) 1/2 11[/]112 := 11/112 = 1 1/ (X)1 2 dx , 21 We use the phrase "almost everywhere", abbreviated a.e., to mean "except on a set of measure zero". Hence " f g a.e." means that {x : f(x) #- g(x)} has a measure zero. 
3.6. Basic Results on LP-Spaces 181 (L2[a, b], II .112) is a normed space and is in fact a Banach space. However, we do not discuss the completeness of L2[a, b]-space in detail. Now, we proceed to discuss the LP[a, b]-spaces and some of their properties. We consider a measure space (X, S, J.t)-that is, a triplet consisting of a set X, au-algebra S and a measure J.t on S. For 1 < p < 00, we define LP(X, J.t) to be the set of all measurable functions I : X  IR such that I/IP is integrable with respect to p" Le. !If(x)IP dx < 00. Remember that elements of LP(X, J.t) are equivalence classes of measurable functions which are equal 'almost everywhere. The LP(X, J.t) spaces are usually studied in the theory of measure and integration and is therefore beyond the scope of this book. However, when X = [a, b] and J.t is the Lebesgue measure, we write LP[a, b] for the corresponding space. Thus, the space LP[a, b] is the vector space of equivalence classes of functions such that in each class two functions are equal almost everywhere. Indeed, by Lemma 2.28(iii), it follows that II ::l: glP < 2 P (I/IP ::l: IgJP) so that . I, 9 E LP[a, b] => I::l: 9 E LP[a, b] . I E LP[a, b] => Ag E LP[a, b] for A E IR. The Holder inequality helps us to show that 1 1 - + - = 1. p q We may also define LOO[a, b] which is a generalization of the space of all bounded functions I on [a, b]. A real valued measurable functions defined on [a, b] is said to be essentially bounded, or simply bounded, on [a, b] if I E LP[a, b], 9 E Lq[a, b] => Ig E L 1 [a, b], I/(x)1 < m a.e. on [a, b] where m > 0 is finite and is called an essential upper bound for III. IT III has an essential upper bound, then there is a least upper bound. The least such bound is denoted by ess sup III. We define LOO[a, b] to be the space of all (essentially) bounded functions from [a, b] into IR. As with the LP-spaces, we are considering the equivalence classes of I as elements of LOO[a, b]. On LOO[a, b], we define the essential supremum 11/1100 := esssup[a,b]I/(x)1 = inf{m > 0: I/(x)1 < m a.e. on [a, b] }. We show that essential supremum is actually a norm on LOO[a, b]. For this we simply observe the following facts: 
182 Chapter 3: Normed Spaces · 11/1100 > 0 for every I E Loo[a, b] and 11/1100 = 0 iff I = 0 a.e. · For A E JR, IIA/lioo = IAlll/lioo . From the inequalities II (x) + g(x)1 < I/(x)1 + Ig(x)1 < 11/1100 + Ilglioo which is true for almost every x E [a, b], it follows that III + glloo = ess sup II + gl < 11/1100 + IIglioo Thus, LOO [a, b] is a normed space under the essential supremum norm. H I has no essential bound, then its essential supremum is defined to be 00. Here is an example to clarify this idea. 3.57. ExampleCi Consider I,g : [-1, 1]  ]R defined by I (x) = x 2 , x E [-1, 1], and { x2 g(x) =  for x \E [-1, 1] \ { 0, :i: 1/3 } for x = 0 for x = :i:1/3. Note that 9 agrees with I except at the points 0, :i:1/3. Then, sup Ig(x)1 = 5, sup I/(x)1 = 1 tE[-l,l] tE[-l,l] but esssuplg(x)1 = 1 =esssupl/(x)l. We remark that I and 9 are different functions on [-1,1] but are considered to be the same, because I = 9 a.e., as elements of L oo [-1, 1]. . HIE LP[a, b] with 1 < p < 00, we define a function II · lip by II · lip : P[a, b]  IR, [/]:= 1 t-+ 11[1] lip = (l b I/(x)IP dx ) IIp =: III lip. Then, for 1 < p < 00, we note that · II I lip > 0 and, II I lip = 0 iff I = 0 a.e. · IIA/lip = IAlll/lip for A E IR · III + g!lp < III lip + IIgllp, by the Minkowski inequality. 
3.6. Basic Results on LP-Spaces 183 This observation shows that, for 1 < p < 00, the function II · lip defines a norm on LP[a, b], called LP-norm, which makes the space LP[a, b] a normed space. It is an infinite dimensional Banach space. In fact, with the help of the Holder and the Minkowski inequalities, it can be shown that LP[a, b] forms a Banach space. However, we shall prove a weaker form of this result in the next section. Moreover, the proof for the case 1 < p < 00 is similar to that for the case p = 1. Note also that space LOO[a, b] becomes Banach space under the essential supremum norm defined above. Also, the following result gives a motivation for the use of the notation II · 1100. We leave the proof as an exercise. 3.58. Proposition. 11/1100 = lim p -+ oo II/lIp. Complex LP-spaces are defined in the following way. Let I : [a, b]  C be a complex valued function defined on [a, b]. Then we have the decom- position I(x) = Re I(x) + iIm I(x) where Re I and 1m I are real valued functions defined on [a, b]. For each complex number z = x + iy, we have the following well-known inequalities Ixl Iyl } < Izl, Izi < Ixl + Iyl which shows that I/IP is integrable iff IRe liP and 11m liP are both inte- grable. Thus, a complex valued function I defined on [a, b] is said to belong to LP[a, b] iff Re I and 1m I are both belong to LP[ a, b]. As with the spaces  Cera, b], the complex LP[a, b] space may be denoted by Lt[a, b], if necessary. The complex Banach spaces Lt[a, b] and the real Banach spaces LP[a, b] are collectively referred to as the Lebesgue or LP spaces. We obtain the following simple properties of LP[a, b]: . We have ( b ) IIp ( b ) IIp l lf (X)IP dx < ll1fll dx < IIflloo (b - a)l/P which gives the inequality l b If(x)IP dx < IIfll(b - a). . From the above inequality, we have LOO[a, b] S; LP[a, b] for 1 < p < 00. By Proposition 3.58, we observe that if I E LOO [a, b] then 11/1100 = lim III lip. P-+OO 
184 Chapter 3: Normed Spaces . For 1 < p < 00, the inequality i b I/(x)g(x)IP dx < IIgll i b I/(x)IP dx shows that II 19 lip < IIg II 00 II/lIp, that is, I E LP[a, b] and 9 E Loo[a, b] implies that Ig E LP[a, b]. The case for p = 00 is easy. Another fundamental property of LP-spaces is the following inclusion result. . 3.59. Theorem. For 1 < p < q < 00, we have Lq[a, b]  LP[a, b]. Proof. Let 1 < p < q < 00 and I E Lq[a, b]. Then, we have i b (1/(x)IP)qlp dx < 00 so that I/IP E Lq/P[a, b]. By the definition of LP-norm, we have III/I P · 1111 = i b I/(x)IP dx = 11/11: and, by Holder's inequality (Lemma 2.26(iii) with u = I and v = 1), we see that II/II - 111/1" 1111 < (i b (1/(x)IP)m dx ) 11 m (i b In dx ) 11n [ (i b I/(x)lmp dx ) lImp] P (b _ a)l / n 11/11p (b - a)l-l/m. If we choose m and n such that n m = and mp = q, n-1 then the last inequality is equivalent to II/II < II/II (b - a)l-P/q, or equivalently 1 1 - + - = 1, m n II/lIp < 1I/IIq (b - a)l/P-l/q and the desired inclusion result follows. The case q = 00 has been consid- ered already. _ 
3.7. Norms on C[a, b] 185 3.7 Norms on C[a, b] The aim of this section is to show that C[a, b] is not a Banach space with respect to certain norms (other than the supnorm discussed in the previous section). We first consider the II · lip-norm on C[O, 1] defined by ( b ) IIp II/lip = 1 1 /(t)IP dt · Let us first consider the case p = 1. It is then a simple exercise to see that (C[a, b], II · 111) is a normed space. Indeed, since each I E .C[a, b] is a continuous function on [a, b], the function III is Riemann integrable and therefore, the function II .111 given by II/Ill = l b 1/(t)1 dt is well defined on [a, b]. Let us first verify the axiom (Nl). Clearly I = 0 implies that 11/111 = O. We show that 111111 = 0 => I = O. Suppose to the contrary that 11/111 = 0 but I :j:. o. Then, since I is not identically zero on [a, b], there exists a number c E [a, b] such that I/(c)1 > O. Continuity of I shows that I/(t)1 > 0 for all t in some interval [0:,.8] which lies completely in [a, b] with c E [0:, .8]. But then II/Ill = l b 1/(t)1 dt > I: 1/(t)1 dt > 0 which is a contradiction to the assumption that 11/111 = o. Thus (Nl) follows. The axiom (N2) is a straightforward calculation whereas (N3) is a con- sequence of the triangle inequality. Before we prove a general result about the incompleteness property of C[a, b] with respect to the p-norm (see Theorem 3.60), we present the standard arguments to show that C[O, 1] is not a Banach space with respect to the LP-norm for the special cases p = 1,2. To see this, we define In(t) on [0, 1] by o if 0 < t < 1/2 In(t) = at + b if 1/2 < t < (n + 1)/(2n) 1 if (n + 1) / (2n) < t < 1. Note that the graph of In consists of three line segments, the middle one joining the points (1/2,0) and ((n + 1)/2n, 1). For In to be continuous, we 
186 Chapter 3: Normed Spaces s  , II OJ  : \ :  o 1 2 1+1m 2 2 1 + 1n 2 2 t Figure 3.5: A Cauchy sequence in C[O, 1] without limit must have a a(n+ 1) fn(1/2) = b + 2 = 0, fn((n + 1)j(2n)) = 2n + b = 1 which, by solving for a and b, give a = 2n and b = -n so that at + b = n(2t - 1) for 1/2 < t < (n + 1)j(2n). With these values of a and b, each element of the sequence {fn(t)} is a con- tinuous function and hence belongs to C[0,1]. Now, we show that {fn(t)} is a Cauchy sequence in C[O, 1] but does not converge to a limit in C[0,1], see Figure 3.5. With m > n, we have IIfm - fnll = ( r(nH)/2n + r 1 ) Ifm(t) - fn(t)1 2 dt 11/2 1(n+l)/2n < ( r(nH)/2n + r 1 ) (1 _ fn(t))2 dt, 11/2 1(n+l)/2n since 0 < fm(t) - fn(t) < 1 - fn(t), j (n+l)/2n - (1 - fn(t))2 dt, 1/2 since 1 - fn(t) = 0 for (n + 1)/(2n) < t < 1, j (n+l)/2n - (1 - n(2t - 1))2 dt 1/2 j (n+l)/2n < dt = 1j2n 1/2 
3.7. Norms on C[a, b] 187 so that 111m - Inl12 < 1/V2ii < t: for m > n > N > 1/2t: 2 and therefore {In(t)} is Cauchy. Finally, we show that the sequence {In} does not converge to a continu- ous function. Suppose, on the contrary, that {In} converges to some contin- uous function I on C[O, 1] with respect to the L2- norm . Then Il/n - 1112  0 as n  00, and therefore, r 1 / 2 r 1 / 2 10 If(tW dt = 10 Ifn(t) - f(t)12 dt < IIfn - fll -t 0 so that I(t) = 0 for t E [0,1/2], because of the continuity of I. On the other hand, given a real number a > 1/2, one can always find a large n so that n+l < . > 1 2n - a, I.e. n - 2a _ 1 · This observation implies that for a given a E (1/2, 1), there exists N E N such that n > N => In(t) = 1 for t E [a, 1]. Hence, for n > N, i 1 11 - f(tW dt = i 1 Ifn(t) - f(tW dt < IIfn - fll -t 0 as n  00. Note that the first integral is independent of n and therefore has to be zero, which implies that I(t) = 1 for all t E [a, 1]. By the continuity of I, I(t) = 1 for all t E (1/2,1] since a > 1/2 is arbitrary. This is a contradiction to the assumption that I is continuous. Hence, C[O,I] is not a Banach space with respect to the L2- norm . Finally, we show that the normed space C[O, 1] provided by the integral norm, Le. Ll- norm (or simply I-norm), defined as IIfll1 = l b If(t)1 dt is not complete. To see this, we consider the same sequence In(t) E C[O,I]: With m > n, we have IIfm - fnll1 - 1 1 Ifm(t) - fn(t)1 dt j (n+l)/2n I/m(t) - In(t)1 dt 1/2 j (n+l)/2n 1 < 2 dt = -, 1/2 n 
188 Chapter 3: Normed Spaces since I/m(t) - In(t)1 < I/m(t)1 + I/n(t)1 < 2, so that the sequence {/n(t)} is Cauchy. Next, we show that there is no continuous function to which the sequence {In} converges with respect to Ll-norm. Suppose on the contrary that In converges to a limit I E C[O,l] in Ll-norm. As before, it can be shown that 1 1/2+1/2n lim I/n(t) - l(t)1 dt = 0, n-+oo 0 and Le. I(t) = 0 for t E [0,1/2) lim (l 11 - f(t)1 dt = 0, i.e. f(t) = 1 for all t E (1/2, 1]. n-+oo J 1/2+1/2n These observations show that I cannot be continuous at t = 1/2. Hence, C[O,l] is not a Banach space with respect to the L1-norm. More generally, we have 3.60. Theorem. The space C[a, b] with respect to the LP-norm is not complete for every p with 1 < p < 00. Proof. Let {In(t)} be a sequence of continuous functions taking values between 0 and 1 with the following two properties: (i) In  0 on every interval [a, c - €] (ii) In  1 on every interval [c + €, b], where c is a fixed point in (a, b) (For instance, the sequence of continuous functions {/n(t)} on [a, b] defined by { 0 for a < t < c - l/n In(t) = n 1 (t - c + l/n) for c - l/n < t < c , c E (a, b), for c < t < b satisfies the conditions (i) and (ii) with € = l/n and c = (a + b)/2). Note that I/n(t)1 = In(t) for each n. Then, for any given € > 0 and with m > n, we have (l C - f l c+f l b ) II/n - Imll = + + I/n(t) - Im(t)IP dt < € + 26 + € = 4€ a C-f C+f which converges to 0 as n, m  00. Therefore, {In} is a Cauchy sequence. Suppose on the contrary that In converges in p-norm to a limit I, where I E C[a, b]. We may first note that if a sequence {In} converges in p-norm to a function I E C[a, b] and also converges uniformly to a function 9 on the subinterval [c, dJ C [a, b], then I(t) = g(t) on [c, dJ. Indeed, if we restrict functions in C[a, b] to the interval [c, dJ then we have . IIfn - fll = l d Ifn(t) - f(t)IP dt < i b Ifn(t) - f(t)IP dt ---+ 0 as n  00, 
3.7. Norms on C[a, b] 189 s s = In(t) / -1 _1 0 n 1 n 1 t Figure 3.6: Graph of s = fn(t) and d IIfn - gll = r Ifn(t) - g(t)IP dt < (d - c) max Ifn(t) - g(t)IP -t 0, lc tE[c,d] as n  00, so that I(t) = g(t) on [c, d], since the limit is unique. Thus, in particular, if the sequence {/n(t)}, which we have assumed at the beginning, converges to I with LP-norm, then we must have f(t) = {  for a < t < c for c < t < b ' c E (a , b). But I is not continuous and hence we arrive at a contradiction. - 3.61. Remark. In the space CF[a, b], we note that In  I in sup- norm iff In(t)  I(t) uniformly on [a, b]; Le. for a given € > 0 there exists a natural number N, independent of t, such that for every t E [a, b] I/n(t) - l(t)1 < € for n > N. . 3.62. Example. Consider the sequence {In(t)} of continuous func- tions on [-1,1] each of whose graph is given as in Figure 3.6. Note that { 0 for t fj; [-1 In, 1 In] In ( t) = 1 + nt for -1 In < t < 0 , t E [-1, 1]. 1 - nt for 0 < t < 1 In Let X = (C[-I, 1], II . 111) and Y = (C[-I, 1], II · 11(0)' Then for all n > 1, we have 11/111 = / 1 Ifn(t)1 dt = .!. -1 n which is the area of the isosceles triangle with vertices at (-l/n,O), (0,1) and (1/n,O). Also, II/nlloo = 1. Then In  0 in X but {In} does not converge to 0 in Y. 
190 Chapter 3: Normed Spaces s s = gn(t) 1 -1 o 1. n 1 t Figure 3.7: Graph of s = 9n(t) Suppose that {gn(t)} is a sequence of continuous functions on [-1,1] each of whose graph is as shown in Figure 3.7. Geometrically, it is easy to see that 119n - 9mlh = ill 19n(t) - 9m(t)1 dt  0 as n,m  00. Clearly, {gn} cannot converge to any continuous function because its point- wise limit f is given by 9(t) = {  if-1 < t<0 ifO<t < l which is clearly not continuous on [-1, 1]. . 3.63. Remark. Suppose that we are given a norm or norms on a vector space V. Is it possible to find new norms? Yes, it is! A sum of several norms and a positive constant multiple of a norm on V is again a norm on the same vector space V. Further, if II · 111 and II · 112 are two different norms on V then a new norm 11.11 on V may be defined by lIuli = max{ lIull1' lIuI12}' Can this be generalized if we are given a finite number of norms on V? Do we get a new norm when we replace max by min? . 3.64. Remark. Some important subspaces of B[a, b] are C[a, b], [a, b] (the space of all differentiable functions on [a, b]), the space of all polynomial functions on [a, b], and R[a, b] (the space of all Riemann inte- grable functions [a, b]). Each of these subspaces are normed spaces with respect to the supnorm. However, only the subspaces C[a, b] and R[a, b] are closed, and therefore, they are the only Banach spaces here. . 
3.8. Exercises 191 3.8 Exercises 3.65. Determine whether the following statements are true or false. Justify your answer. (a) It is always possible to define a norm on each vector space. (b) The set {x: IIx - all < R} on IRn is convex where ( n ) 1/2 II x - a II = E (x k - a k ) 2 , X = (Xl, · · · , X n ) , a = (a 1 , · · · , an). k=l (c) If {x n } is a sequence in a normed space X, then X n  a implies that Yn -+ a, where Yn = (Xl + X2 + · · · + xn)/n. (d) If II .1 h and II .112 are. two norms on a vector space X, then II .11 defined by IIxll = IIxlh + IIxll2 for x E X is also a norm. (e) If X is any nontrivial vector space, then there is no norm which induces the discrete metric. (f) The metric space (C,d), where d(z,w) = Iz - wl/(l + Iz - wI), does not define a norm on C. (g) The metric space (C[O, 1], d), d(f, g) = max x E[O,l] arctan I!(x) - g(x)l, does not define a norm. . (h) Let (IR, d) be a metric space. Define d* by 1 + d(x, y) if one and only one of x, y, d* (x, y) = is strictly positive, d( x, y) otherwise. Then d* defines a metric on IR and there is no norm which induces the metric d*. (i) Any metric induced by a norm is always unbounded. (j) The set of all convergent sequences in a normed space (V, 11.11) forms a vector space as well as a normed space. (k) The space of polynomials p(z) on [a, b] with the supnorm is a normed space but not a Banach space. Note: The space of polynomials p(z) on [a, b] with the supnorm is a subspace of (CF[a, b], II . 11(0) with respect to the supnorm IIplloo = SUPtE[a,b] Ip(t)l. Any contiuous function can be uniformly approxi- mated by polynomials on [a, b], by the Weierstrass theorem. (1) If X = R[a, b], the space of all functions f : [a, b]  IR such that If I is Riemann integrable, then the function II . 111 defined by II/Ill = lab 1/(t)1 dt 
192 Chapter 3: Normed spaces is not a norm on R[a, b]. (m) The space Pn[a, b] of polynomials p(t) on [a, b] is a normed space but not a Banach space with respect to the Ll-norm IIpll1 = lab Ip(t) I dt. (n) If X = C1[a,b], then the formula 11/111 = lab 1!,(t)1 dt, I E C 1 [a, b], defines a seminorm on X but not a norm whereas the modified formula 11/111 = 1/(0) I + lab II' (t) 1 dt defines a norm on X. (0) In a normed space (X, 11,11), we have S(xo; 6) = Xo + 6B(O; 1), where Xo E X, 6 > O. (p) For ReA> lip, the sequence {n-A}n>l belongs to lP. In particular, {n-A}nl E 1 2 for A E IR with A E (1/2,00). (q) The p-norm II · lip on r satisfies the inequalities IIzllq < IIzllp and IIzllp < n (q-p)/pq IIzllq for p < q. (r) The subset 8 = {f E C[O,l] : f(x) > 0 for all x E [0,1]} is open in (C[O, 1], II. 11(0)' ' (s) The subset S = {f E C[O,l] : f(O) = O} is a closed linear subspace of (C[O, 1], II .11(0)' (t) For the subset 8 = {f E 0[0, 1] : f(O) = 0, If(t)1 < 6}, the element g(t) = 6 E (C[O, 1], 11.11(0) is not a limit point of S. (u) If X = (C[a, b], 11.11(0), then the subset y = {I EX: lab I(t) dt < 1 } is open in X. (v) In any nontrivial normed space, there exist subsets that are not open. (w) The sequence {fn(t)} of functions in C[a, b], where o fn(t) = n(t - c + 1/n) 1 for a < t < c - 11n for c - 1 I n < t < c C E (a b) , , , for c < t < b 
3.8. Exercises 193 is Cauchy with respect to the L1 and L2-norms but does not converge to a limit in C[a, b] with respect to these norms. Note: See Theorem 3.60. (x) The sequence {fn(t)} of functions in C[O, 1], where { 2ntn+1 In(t} = 1 - 2 n (1 - t}n+1 for 0 < t < 1/2 for 1/2 < t < 1 ' C E (a, b), is Cauchy with. respect to the L1 and L 2 -norms but converges to a discontinuous function f with respect to these norms, where o forO < t<I/2 f(t) = 1/2 for t = 1/2 1 for 1/2 < t < 1. (y) If n = {(x, y) E IR 2 : 0 < x, y < I}, then the function fn : n  JR2 defined by ( ny nx ) In(x,y}= n+x ' n+y is uniformly convergent on n to g(x,y) = (y,x). (z) The map T : LP(IR)  LP(IR) defined by f(t) I-t t- 2 / p f(l/t), t E IR, is an isometry. 3.66. If a E C with 0 < lal < 1 and Zk = {ank}no. Then prove or disprove that {Zk} E 1 2 for each kEN and Y = {Zk : kEN} is dense in 1 2 . 3.67. Define G = {{X n }n>l E , 1 : IXkl < l/k2, for k = 1,2,. . .}. Verify whether G is an open subset of 1 1 or not. 3.68. Let n = {x = (Xl, X2, . . . Xn): 0 < Xi < 1 for 1 < i < n} and let cPn : n  JRn be defined by A,. ( ) _ ( nXn nXn-1. . . nX1 ) 'Pn X - n , n " n · n + E "'=1 Xk n + E "'=1 Xk n + E "'=1, Xk "'n kn-l "'1 Is cPn convergent uniformly to cP(x) = (X n ,X n -1,...,X1) on n? 3.69. Let n = {x = (Xl, X2, . . . Xn): 0 < Xi < 1 for 1 < i < n} 
194 Chapter 3: Normed spaces and let cPk : {}  }Rn be defined by cPk(X) = (x ,x,... ,x). Does cPk converge pointwise to a function cP on {} as k  oo? If so, find the limit function cPo Is the convergence uniform? 3.70. Determine which of the following defines a norm on X: (i) IIxll = 2 [x] , X = }R (Here [x] denotes the largest integer < x). (ii) IIxll = log lxi, X = IR. (iii) Ilxll = exp x, X = IR. (iv) lI(x1,x2)1I = IX11, X =}R2 and (X1,X2) E }R2. (v) II(X1,X2)1I = max{21x11 +3Ix21,3Ix11 +2Ix21}, X =}R2 and (X1,X2) E }R2. (vi) II(X1,X2)1I = IX1 +x21, X = }R2. (vii) II(Zl,Z2"..,zn)1I = IZ11, X =en. (viii) II(Zl, Z2,..., zn)1I = EZ=l I Z kI 2 , X = en. (ix) IIfll = U:U' (t))2 dt) 1/2, X = {J E C1 [a, b] : f(a) = f(b) = O}. 3.71. Let A1,. . . An be fixed positive real numbers. Define (i) II(Zl,...,zn)lh = EZ=l Ak l zk l, (ii) lI(zl"",zn)112 = (E Z- 1 Akl z kI 2 )1/2, (iii) II ( Z 1, . . . , Zn) 112 = (E Z= 1 Ak I Z k 1 1 /2) 2 , (iv) II(Zl,.", zn)lIoo = max1kn AklZkl. Determine which of the above defines a norm on en . 3.72. If x = (1, -1,2) E }R3, then find IIxli p when p = 1,3,5,7,00.. 3.73. Let V denote the set of all complex-valued functions I(z(t)) = u(t) + iv(t), t E [a, b], which are continuously differentiable on [a, b]. For I E V, define (i) II/lh = maxxE[a,b] I/(x)l, (ii) 11/112 = maxxE[a,b] {1/(x) I + II' (x)I}, (iii) 11/113 = maxxE[a,b] II' (x) I, (iv) 11/114 = maxxE[a,b] I/(x)12. 
3.8. Exercises 195 Determine which of the above defines a norm on V. Verify also the com- pleteness property for those which are normed spaces. 3.74. If X = LP[O, 1], I(t) = t and g(t) = t 3 , then find III lip and IIglip for 1 < p < 00. Verify lim p -+ oo II/lIp = Ilglloo and lim p -+ oo IIglip = IIglioo. 3.75. Show that the functions II · 111,00 and II .111,1 on en [a, b] defined by 11/111,00 = sup I/(t)1 + sup 1/'(t)l, tE[a,b] tE[a,b] and IIflh.1 = i b {If(t)1 + If'(t)l} dt, are norms on en [a, b]. 3.76. Show that the normed spaces LP[a, b], 1 < p < 00, are complete. Note: We have already observed this result in Section 3.6 but without a detailed proof. 3.77. Let V = A, the space of all analytic functions I in the unit disc  = {z E C: Izl < I} and continuous on the closure  = {z E C: Izl < I}. Show that (V, 11.11) is a normed space with the norm 11/11 = maxlzl=1 I/(z)l. 3.78. In a normed space (V, 11,11), show that the mapping cP : V x V  V, (u,v) I-t u+v, as well as the mapping 1/J: C x V  V, (A,u) I-t AU, are continuous. -- 3.79. Let I : C  IRt be defined by the formula f(z) =  14Rez + iImzl 2 , Does this define a norm on C? Draw the picture of the unit ball B(O; 1). 3.80. Let (X, d) be a metric space, a E X and r > O. Prove that the subset S(a;r) = {x EX: d(x,a) = r} is closed. Give an example of a situation where X is not bounded with respect to d but S(a; r) = 0. Show that if X  {OJ is a real vector space and d is determined by a norm 11,11, then S(a; r) :j:. 0 for all a E X and r > o. 3.81. Show by an example that the Heine-Borel Theorem (see Propo- sition 2.77) does not extend to infinite dimensional normed spaces. 
Chapter 4 Contraction Mappings and Applications In this chapter, we present several classical theorems on fixed point prop- erties and in particular, Banach fixed point theorem, Peano's theorem on differential equations. We include a large number of examples for motivat- ing the fixed point theory. 4.1 Discussion on Fixed Point Problems Given a nonempty set X and an operator T on X into itself, the problem of finding a vector jpoint x E X such that Tx = x is called a fixed point problem and the solution x is called fixed point (or invariant point) of the operator T. The space X is said to have a fixed point property if each continuous operator T : X  X has a fixed point. A natural question is under what conditions on X and T, a fixed point exists? Theorems which establish the existence (and uniqueness) of such points are called fixed point theorems. There are a number of versions of these theorems and especially, when X is a complete metric space. These have several simple and fun- damental results which receive far reaching applications. In this section, we present the simplest and more widely used version of the fixed point theorem and some of its consequences via completeness. The theorems of this type often enable us to solve the existence of the solutions of operator equation satisfying certain conditions. For example, Fredh<;>lm and Volterra integral equation, two point boundary value problems in differential equa- tions as well as in eigenvalue problems including approximation theory and variational inequality. Indeed, an operator equation Tx = y may be equivalently transformed to fixed point formulation: Sx = x, with Sx = x + Tx - y 
198 Chapter 4: Contraction Mappings and Applications b ,  ,,' // , "'\ " ''0 , " , , a , " , , , , , , , , , , , , , , , " , " , , , , ,,' , " , , , , y = f(x) o a b Figure 4.1: Contraction: f([a, b]) C [a, b] and therefore finding a vector x such that Sx = x is same as solving the equation Tx = y. Thus, many problems involving operator equations can be put in the form of finding fixed points. Often, in classical real analysis, finding the zeros of a real valued function g(x) defined on an interval is same as finding the fixed points of f(x), where x - f(x) = g(x). In order to illustrate the fact, consider the quadratic polynomial g(x) = x 2 - 5x + 4 which has x = -4, -1 as its zeros. If we rewrite the equation g(x) = 0 as x - f(x) = g(x), f(x) = x 2 + 4 , 5 then it is clear that the problem of finding the roots of g(x) = 0 is same as finding fixed points of f(x). A classical example of this type follows from the Intermediate Value Theorem 22 . More precisely, we have the following (see Figure 4.1): 4.1. Proposition. Every selfmapping23 f of bounded interval [a, b] has a fixed point in [a, b]. 22Intermediate value theorem says that if 1 : [a, b]  IR, a < b, is continuous such that l(a)/(b) < 0, then there exist a point c E (a, b) such that I(c) = o. It is important to note that the result depends on 1 taking values in IR rather than in ]R2 or C. For example, the map 1 : [0, 211"]  C defined by I(t) = e it is continuous, 1(0) = 1, 1(11") = -1 but I(t) f:. 0 for all t. 23By a self-mapping T of a space X we mean a single-valued continuous mapping T from X into itself. 
4.1. Fixed Point Problems 199 Proof. Consider the metric space X = [a, b] with the usual/Euclidean metric, and assume that f : [a, b]  [a, b] is continuous, so that 4>(t) = f(t) - t is continuous. Since f([a, b]) C [a, b], we have f(a) > a and f(b) < b so that 4>(a) = f(a) - a > 0 and 4>(b) = f(b) - b < O. If 4>(a) = 0 or 4>(b) = 0, then it is clear that either a or b is a fixed point of f. If this is not the case, then we must have 4>(b) < 0 < 4>(a). Since 4> is continuous, Intermediate value theorem guarantees that there exists a number c in (a, b) such that 4>(c) = 0, Le. f(c) = c. Hence, f has a fixed point on [a, b]. . The proof of Proposition 4.1 shows that "every continuous and bounded map f from the unbounded interval [0, 00) into itself has a fixed point in [0,00)". This fact follows easily if we define 4>(t) = f(t) - t, where f(t) < k for all t > 0, and observe that 4>(0) > 0, 4>(k + 1) = f(k + 1) - k - 1 < -1. We see that the boundedness condition in this example is essential for the existence of a fixed point. For example, there exists a continuous map f from [0,00) into itself without having a fixed point in [0,00). In fact, the mapping f(t) = t 2 + a, where a > 1/4 is an arbitrary fixed real number, does this job, as the inequality f(t) - t = (t - 1/2)2 + a - 1/4 > 0 holds for all t and hence has no fixed point. Next, we state (without proof) a result which extends Proposition 4.1 for higher dimensional Euclidean spaces proved by Brouwer in 1912: 4.2. Proposition. Every self mapping f of a closed unit ball B[O; 1] in }Rn has a fixed point in B[O; 1]. It is natural to ask whether there exists a larger class of other mappings in metric spaces which have fiXed points. Now, we consider the fixed point problem (4.3) x=Tx, T:XX, for a metric space X. A close examination of (4.3) suggests a method of constructing a sequence of approximations to the exact solution x: let Xl be the first approximation to x. If Xl were the exact solution of (4.3), then TXl = Xl; otherwise Xl ':j:. X so that TXl will not equal either Xl or 4 x. Therefore, we may set X2 = TXl is the second approximation for x. Repeating the arguments,...we end up with a sequence Xn+l = TXn, n E N. Then we say that the map T assigns on X a method of successive approxi- mations or the iteration method. The sequence {xn} so obtained is referred to as an iterative sequence. Now the problem is to ask whether {xn} con- verges to x. This question led to the introduction of contraction mapping in metric spaces. 
200 Chapter 4: Contraction Mappings and Applications f:X--tX Figure 4.2: Contraction 4.2 Contraction Mapping Principle 4.4. Definition. Let (X, d) be a metric space and T : X  X be a mapping which maps X into X. A point x E X is called a fixed point of T iff Tx = x. The mapping T is called a contraction mapping of X (or simply T is a contraction), iff there is a constant 0 < a < 1 satisfying the Lipschitz condition (4.5) d(Tx,Ty) < ad(x,y), x,y E X. The smallest a, denoted by a(T), for which (4.5) holds is said to be Lipschitz constant for T (see also Definition 2.85) and in this case, the map T is called a a(T)-contraction with respect to d. Some authors use the notation ad(T) to denote the Lipschitz constant of T with respect to the metric d. Since a E (0, 1), we see that the distance between the image points Tx, Ty contracts by at least a factor of a under the mapping T and so is called with the name 'contraction mapping' for T, as illustrated in Figure 4.2. For example, to solve the equation x 2 = a (see p. 122) we may rewrite this equation into a fixed point problem x = f(x) =  (x + : ). Then for x, y > 0, If(x) - f(y)1 = Ix - yl 1 -  2 xy and therefore, If(x) - f(y)1 < ,8lx - yl if 1 _.!:. < 2,8 with ,8 < 1. xy 
4.2. Contraction Mapping Principle 201 Thus, f is a contraction mapping if one necessarily has a < (2{3 + l)xy which means that we have to exclude y that are too small. This observation suggests that we must try a metric space of the form ( [a, 00 ) , d), a > O. Note that ( [a, 00 ) , d) is a closed su bspace of the complete metric space (JR, d) and hence complete. (For example, let a = 2. As discussed in p. 122, the sequence Xn+l = f(x n ), n E N, converges to a solution of x 2 = 2, where Xl = 2 can be chosen as the first approximation to x.) However, to make a good choice of a and apply fixed point theorem, we may write that 1 ( a ) (x-va>2 f(x) = 2 x + x = va + 2x > va for each x > o. Moreover, for x,y > va, we have 11- (a/xy) I < 1 with Lipschitz constant 1/2 and hence, f is a contraction. 4.6. Examples. We shall now present a list of simple examples, comments and remarks: First we recall that a map T : X  Y between two normed spaces is called an isometry (into) if IITx - Tx'ily = Ilx - x'lIx for all x, x' EX. If, in addition, the map T is linear, then it suffices to check that IITxlly = IIxlix for every x E X. (i) Clearly, since a is independent of x, y EX, a contraction mapping is uniformly continuous. However, it is interesting to oote that the converse is not true. For example, if k > 1 is fixed then f(x) = kx is uniformly continuous on IR but not a contraction map. Here f'(x) = k so that f has a bounded derivative on III Let us look another example. Consider f : (IR+ , d)  (IR+ , d), x 1-t..[X, d-usual metric. Then for all x, y E R+ , we have { x-y I..[X - v'Y12 = x + Y - 2VXY < y-x ifx > y ifx<y = Ix - yl so that T is uniformly continuous (with 6 = €2). On the other hand, for instance d(O, 1/3) = 1/3, d(f(O), f(1/3)) = 1/v?' = v?'d(0,1/3) and therefore, T is not a contraction. Note that T'(x) = 1/(2..fi) and therefore, this function has unbounded derivative in R+ . (ii) Let a < 1 be fixed. Then the mapping T : lP  IP, Z I-t {aZn}nl, is clearly a contraction. Similarly, each mapping T defined on the n-dimensional Euclidean space (IRn, d), x I-t ax, is a contraction. 
202 Chapter 4: Contraction Mappings and Applications (Hi) The translation mapping T on the Euclidean space (IRn, d), x I-t a+x (a :j:. 0), has no fixed points. Note that d(Tx, Ty) = d(x, y) and ]in is convex but not compact (compare with Exercise 4.29). But T: (IRn,d)  (IRn,d), x I-t a+x/4 (a  0), has a fixed point x = 4a/3 on an. (iv) The mapping T on the Euclidean space (IR 2 , d) such that _ ( Xl ) ( a -e ) x - I-t X, X2 e a a,e E IR, (a - 1)2 + e 2 :j:. 0, has only one fixed point (0,0) E IR 2 . In fact, this is a linear map and hence, (0, 0) is always a fixed point. Also, as (a - 1)2 + c2 :j:. 0, it is unique. If x = (Xl, X2) and y = (Yl, Y2), then we have TX-Ty= ( a c -e ) ( X l - Yl ) . a X2 - Y2 It is easy to show that IITx - TylI = IIx - yll (a 2 + e 2 ). In fact, because of the linearity of T, it suffices to show that IITxll = IIxll (a 2 + c 2 ). Notice that IITxll - lI(axl - CX2, eXl + x2)11 - (axl - CX2)2 + (CXl + aX2)2 _ IIxll (a 2 + e 2 ). This observation shows that, if a 2 + c2 = 1 then IITxl12 = IIxll2 and therefore, T is an isometry. (v) Consider the space X := R = {z E C: Izl < R} with respect to the Euclidean metric. Then T : R  R, Z I-t zn (n > 2), where R = (a/n)l/(n-l) with a < 1 being fixed, is a contraction. In particular, the following special cases are interesting to state separately. (a) T: l/4  l/4' Z I-t Z2, is a contraction with a = 1/2. (b) T: l/6  l/6' Z I-t Z2, is a contraction with a = 1/3. (c) T: l/V6  l/V6' z I-t Z3, is a contraction with a = 1/2. (vi) Define T : [0, 1]  [0, 1], X I-t x 2 - x + 1. Then x = 1 is the unique fixed point but T is not a contraction mapping on [0, 1]. Since Tx - Ty = (x - y)(x + y - 1) 
4.2. Contraction Mapping Principle 203 and since sUPOx,yl (x + Y - 1) = 1, we have a = 1 which shows that T is not a contraction. Alternatively, this fact can also be proved as follows: Suppose on the contrary that T is a contraction mapping. Then there is a constant a < 1 satisfying the Lipschitz condition ITx - Tyl < alx - yl for all x, y E [0, 1]. Thus, for x = y + h, where h, y > 0 are small enough, we have T(y + ) - Ty < a with a < 1 which, by taking h -+ 0, gives IT'(y)1 < a < 1, for all y E [0,1). But T' (0) = -1 and this contradiction proves our assertion. (vii) The mapping T : ]R2  }R2, (x, y) I-t (0, y), has all points of the y-axis as fixed points and hence, has infinitely many fixed points. Similarly, the map T : }R2  }R2, (x, y) I-t (x,O), has infinitely many fixed points. . Recall that, since all distances under contraction mapping" are reduced by at least a factor of a, repeated application of T in (4.5) will shrink distances drastically so that the existence of unique fixed point becomes plausible. Thus, we have the following simple version of "Banach con- traction principle" which is perhaps the most frequently cited fixed point theorem in many branches of mathematical analysis: 4.7. Theorem. (Contraction Mapping theorem) Every contrac- tion mapping on a complete metric space has a unique fixed point. . Proof. Let Xo be arbitrary. Form a sequence of points {Xn}nl recur- sively by Xn+l = TXn, Le. Xl = Txo, X2 = TXl = T(Txo) = T2xo, ..., Xn+l = Tn+lxo (This procedure of constructing a sequence of elements starting from a point Xo is called the method of iteration). Since T is a contraction mapping, the definition of X n gives d(Xn+l,Xn) = d(Txn,Txn-l) < ad(Xn,Xn-l) for all n > 1, and therefore, by induction, it follows that d(Xn+l' x n ) < and(xl' xo) for all n > 1. Hence, for n > m, we have from the Triangle inequality d(xm,xn) < d(Xm,Xm+l) + d(X m +l,X m +2) +... + d(Xn-l,Xn) < d(Xl, xo)[a m + a m + l + . . . + a n - l ] am < d(xl,xo) l_a  0 as m  00, since a < 1. 
204 Chapter 4: Contraction Mappings and Applications Note that if € > 0 is arbitrary, then we have am 1 ( (1 - a)f ) d(Xl, xo) 1 < € {::::::} m > I log d( ) . - a oga Xl,XO Therefore, there exists an N such that d(xm,xn)<€ forn,m>N which shows that {xn} is a Cauchy sequence in X and, since X is complete, {xn} converges to a point x in X. We now claim that x is a fixed point of T. Clearly, T is continuous and since every continuous function in a metric space preserve the convergence (see Proposition 2.59), we have Xn+l = TX n  Tx, but Xn+l  x, which implies Tx = x. As for the uniqueness, suppose that x, x' are fixed points of T. Then d(x', x) = d(Tx', Tx) < ad(x', x) and since a E (0, 1), the last inequality is possible only when d(x', x) = 0, · h ' I.e. w en x = x. . The proof of the preceding theorem contains a result which is interesting enough to be stated separately. This form of Theorem 4.7 is often more convenient. 4.8. Corollary. Let 8 be a closed subset of a complete metric space in (X, d) and T : 8  8 be a contraction mapping. If Xo be any point of 8, and Xn+l = TX n for every n > 0, then the sequence {x n } converges to the fixed point of T. Proof. From the proof of Theorem 4.7 and hypothesis, it follows that {xn} is a Cauchy sequence in 8. But, since every closed subset of a complete metric space is complete, the sequence {xn} converges to a limit point x E 8, as in Theorem 4.7, we have T(x) = x. . In applying this corollary one has to keep in mind the two importC¥lt facts: T is not only mapping closed subset 8 into itself but is also a con- traction on 8. For example, if X = IR is a Banach space with usual metric, if 8 is an open unit ball B = {x E IR: Ixl < I} and if p E IR is such that Ipl = 1, then contraction map T: B  B, x+p Xl-t 2 ' 
4.2. Contraction Mapping Principle 205 has no fixed point in B. This example shows that the condition that S is closed in Corollary 4.8 is necessary. Further, under the hypothesis of Corollary 4.8, we also see that d( ) < d(xo, Txo) Xo,x - 1- a(T) holds. Here a(T) denotes the Lipschitz constant. For, d(xo, x) - lim d(xo, Tnxo) n-+oo n-l < lim  d(Tkxo, Tk+lxO) n-+oo  k=O 00 - Ld(Tkxo,Tk+lxO) k=O 00 < Lakd(xo,Txo) k=O 1 1 _ a d(xo, Txo). 4.9. An application. H T is contraction with Lipschitz constant a, then Tn, where n is a positive integer, is clearly a contraction mapping with Lipschitz constant an. However, the converse is not true as the following example points out. Define T: (C[O, 1], 11.11(0)  (C[O, 1],11,11(0), f  Tf, where (T f)(x) = 1 z f(t) dt (x E (0,1)). Note the variable limit of integration in contrast to the Fredholm operator, see p. 216. Then (T2f)(X) = (T(Tf)(x) = 1 z {I t f(U)dU} dt, where 0 < u < t and t < x. Interchanging the order of integration gives (T 2 f)(x) = 1:1: lZ f(u) dtdu = 1 z f(u) (lZ dt) du = 1 z (x - u)f(u) du from which, by induction, we deduce that (Tn f)(x) = (n  1)! 1 z (x - ut- 1 f(u) duo 
206 Chapter 4: Contraction Mappings and Applications In fact, if this formula holds for n = 1,2,..., m, then for n = m + 1, we have (T m +1 f)(x) - T ( m  I)! 1 t (t - u)m-l f(u) du ) (x) - 1 x [ (m  I)! LX (t - u)m-l dt] f(u) du 1 1 x - -, (x - u)m f(u) duo m. 0 If we choose f1 = 1 and f2 = 0 on C[O, 1], then we have d oo (f1,f2) = 1 = d oo (Tf1,Tf2) which shows that T is not a contraction on C[O, 1]. On the other hand, for f, 9 E C[O, 1], we have d oo (T2 f, T2g) - sup 1 x (x - u)(f(u) - g(u» du xe[O,1] 0 1 < 2 sup If(x) - g(x)1 xe[O,1] 1 - 2 d oo (f,g) and therefore, T2 is a contraction. A more general integral operator which we encounter often is the follow- ing type which is called Volterra Integral Operator T with k as its kernel: T : (C[a, b], II .11(0)  (C[a, b], II · 1100) where f(x) t-t (Tf)(x) = 1 x k(x,t)f(t)dt, a < t < x < b, and k : [a, b] x [a, b]  IR is continuous on the triangular region {(x, t) : a < t < x < b} and zero for t > X. As above, it is easy to see that 1 {X (Tn f)(x) = (n _ I)! lr. k(x, t)(x - u)n-l f(u) duo Therefore, technically speaking, the following version is a slightly general- ized form of Theorem 4.7. 4.10. Corollary. Let (X, d) be a complete metric space. 1fT: X  X is such that Tn is a contraction map for some integer n > 1, then Tx = x has a unique solution. 
4.2. Contraction Mapping Principle 207 Proof. Assume that 8 = Tn is a contraction for some n. By Theorem 4.7, 8 has a unique fixed point Xo, say; that is 8xo = Xo so that T8 x o = Txo. Now, since T(8x) = 8(Tx) (= Tn+1x), for all x E X, we have 8Txo = T8xo = Txo which shows that Txo is a also a fixed point of 8. But by uniqueness of the fixed point of 8, we have Txo = Xo. In other words, we say that T has a fixed point Xo. It remains to show that Xo is unique. Since Tnxo = Tn-1(Txo) = Tn- 1x o = ... = Txo = Xo, it follows that a fixed point of T is also a fixed point of 8 = Tn which must be unique as 8 is a contraction on the complete metric space X. Thus, the map T has a unique fixed point. - 4.11. Example. Consider the usual metric d(x,y) = Ix - yl on III Define T : (]R, d)  (]R, d) by x I-t 1 + x/4. Then T has a fixed point at 4/3. IT we let 8 = [0,2], then 8 is a closed subset of the complete metric space (]R, d) and 1 d(Tx, Ty) = ITx - Tyl = 4 1x - yl for x, y E 8. Note that T(8) = [1,3/2] C 8 and thus, T maps 8 into 8. Therefore, T must have a fixed point in 8. On the other hand, for the closed subset 8 1 = [0, 1] of (JR, d), we have T(8 1 ) = [1,5/4] which is not a subset of 8 1 and therefore, in this case, T is not a mapping of 8 1 into 8 1 , Note that T has no fixed point in 8 1 . Finally, T : (0, 1/2]  (0, 1/2], x I-t x 2 , is a contraction map with a = 1/2 but has no fixed point in (0, 1/2]. This is because of the fact that (0, 1/2] is not a complete metric space. Note that Tx = x 2 = x gives x is either 0 or 1 but neither of them belongs to (0, 1/2]. This example establishes the fact that the Banach contraction principle (Theorem 4.7) does not hold for incomplete metric space. . Next, we examine a classical applications of calculus for giving informa- tion about the iteration using the concept of derivative. Let f be differen- tiable at a point a E IR such that f(a) = a. Then with x = a + h, Ihl < 6, we have f(a + h) - f(a) = f(x) - a  hf'(a). Thus if If'(a)1 < 1, where h is small enough, then it seems plausible that f(x) approaches a than x was, and if If'(a)1 > 1, then f(x) should be 
208 Chapter 4: Contraction Mappings and Applications further away. This observation shows that the quantity If' (a) I is helpful in classifying the fixed points into two categories, namely attractive and repulsive. For example, the real root of x 3 + x-I can be regarded as a fixed point of the maps f(x) = 1 - x 3 or g(x) = (-2x 3 + 3x + 2)/5. In the first map, the fixed point turns out to be repulsive whereas in the second map it is attractive. 4.12. Example. Let 8 be either a closed interval in ]R (not neces- sarily bounded) or 8 = III Assume f : 8  8 is a differentiable function on 8, If'(x)1 < a < 1 on 8. Then by Mean value theorem, for x,y E 8, there is a c in the open interval (x, y) such that If(x) - f(y)1 = If'(c)llx - yl < alx - yl, for all x,y E 8, which is the desired Lipschitz condition. Thus f is a contraction. Further, 8 is a closed subspace of the complete metric space ]R with usual metric, and so 8 itself is a complete metric space. According to Theorem 4.7, f has a unique fixed point x E 8, and is given by x = lim x n , n-+oo where Xo is any point of 8 and X n = f(Xn-l) for every n > 1. . This example motivates us to have the following simple result. 4.13. Proposition. A differentiable map f : [a, b]  [c, d] is a contraction on [a,b] iff there exists an a < 1 such that If'(x)1 < a on [a,b]. Proof. First, we assume that f is a contraction on [a, b]. Then If (x + h) - f(x)1 < al(x + h) - xl for x, x + h E [a, b] so that, for h  0, we have f(x + h) - f(x) h < a. As h  0, we have If'(x)1 < a. The converse part follows from Example 4.12. . 4.14. Example. In this example, we show that every differentiable function f on [a,b] with f(a) < 0 < f(b) and f'(x) E [r,R] C (0,00) for all x E [a, b], has a solution c such that f(c) = O. Note that, by Intermediate value theorem, f(x) = 0 has solution in [a, b] whenever f is continuous and f(a)f(b) < O. Our objective here is to find iteration converging to 
4.2. Contraction Mapping Principle 209 the solution. Thus, in order to find c such that f(c) = 0, we convert the problem to a fixed point problem by setting g(x) = x - af(x) where a > 0 will be chosen in such a way that 9 is a contraction mapping of [a, b] into itself. Note that, a point y is zero of f iff y is fixed point of g. Since f(a) < 0 < f(b), we have g(a) = a - af(a) > a, g(b) = b - af(b) < b and therefore 9 is a differentiable map from [a, b] into [a, b]. Further, since f'(x) E [r, R], 1 - aR < g' (x) = 1 - af' (x) < 1 - ar < 1 and so if we choose a in (0,2/ R), we have -1 < g'(x) < 1 for all x E [a, b]. This shows that Ig'(x)1 < 1 on [a, b], and, in particular, there exists a A E (0,1) such that Ig'(x)1 < A for all x E [a, b]. Therefore, by Example 4.12, it follows that 9 has a unique fixed point c in [a, b] which means that f(c) = O. For such A, the solution c can be obtained by iteration. More precisely, this can be achieved by choosing an arbitrary point Xo E [a, b] and then forming the sequence of iterations {x n }, X n = 1- af(xn-l) = g(Xn-l), converging to the unique fixed point c. Finally, we discuss the example of solving the following equation for 8: 8 - e sin 8 = 4>, where 4> is a fixed real number in the interval (0, 21r). We encounter this equation in orbital mechanics, where e is the eccentricity of an orbit of some satellite and 8 is the central angle from pregee. Also, we may note that if P and t the period and the time for pregee respectively, then 4> = 21rt/ P. If we define f(8) = 8 - e sin 8 - 4>, then f(O) = -4> < 0 < 21r - 4> = f(21r) and f'(8) = 1 - ecos8 so that f'(9) lies in the closed interval [1 - e,l + e]. Therefore, from the above discussion, we conclude that there is a unique fixed point in [0,21r] provided a E (0,2/(1 + e)). Indeed, with r = 1 - e and R = 1 + e, we have g(9) = 8 - af(8) = 9 - 0.[8 - e sin 8 - 4>] = (1 - 0.)9 + a(e sin 8 - 4» 
210 Chapter 4: Contraction Mappings and Applications y d(z, s) o x o u z-plane w-plane Figure 4.3: Rotation about Zo in X = {z : r < Iz - zol < R} so that if we choose for example a = 1/(1 + e), then it follows that (8) = ( e ) 8 + e sin 8 - cP . 9 1+e 1+e This observations shows that g'(8) E "[0, A] with A = 2e/(1 + e). Hence, for e < 1, we can solve the given equation by successive approximations. . What happens if a = 1 in the definition contraction? (see Examples 4.6(iii)). Note that the function f(x) = Ixl satisfies Lipschitz condition with a = 1 but not differentiable at x = O. As another example., we consider a closed annulus x = {z E C : r < Iz - zol < R}. Let T be a rotation about the center zoo Then, d(Tz,T() = d(z,() for z,( E X, z # (, so that the contraction condition (4.5) holds with a = 1 (see Figure 4.3). Note that T is continuous on X, where X is compact but not convex. On the other hand, such rotations in general have no fixed points. Thus, a < 1 in Definition 4.4 is essential. What happens to Banach contraction principle if ad(T) = 1? We begin our discussion of this case with the following definition and several simple examples. 4.15. Definition. A mapping T : (X, d).  (X, d) is called contrac- tive (or strictly contractive) if (4.16) d(Tx,Ty) < d(x,y), x,y E X, x  y, and T is called nonexpansive if (4.5) holds with a = 1, i.e. if d(Tx,Ty) < d(x,y), for all x,y E X. 
4.2. Contraction Mapping Principle 211 The translation map f : IR  IR, x I-t x + a with a > 0, is nonexpansive, but f has no fixed point. The case a = 0 shows that the identity map is a nonexpansive map and each point of the domain is a fixed point. Clearly, the basic properties of contraction mappings do not extend to nonexpansive mappings. Thus, it is natural to look for nonexpansive mappings having fixed points (see Exercise 4.30). First, we recall the one way implication: contraction => contractive => nonexpansive. We remark that the weaker property (4.16) does not guarantee the existence. of a fixed point as the example T : IR  IR, x I-t 10g(1 + eX), with usual metric, shows. Thus, Theorem 4.7 fails to hold under this weaker condition. In this example, we note that T'(x) = eX /(1 + eX) < 1 for all x E IR and therefore, by Mean value theorem, T satisfies the weaker inequality d(Tx, Ty) = ITx - Tyl < d(x, y) = Ix - yl, for x, y E IR, x :F y. But T has no fixed point in IR, because ITx - xl = Ilog(l + eX) - xl > 0, for all x E IR. Similar reasoning shows that the function T defined by T:IRIR, eX xl-tx+ 1 ' +e x has no fixed point in IR, where T'(x) = 1- eX /(1 + e X )2 < 1 for x E IR, and ITx - Tyl < Ix - yl for x, y E IR, x :F y. Finally, it is easy to see that the function T defined by T:IRIR, 1 x I--t Ixl + 1 + lxi ' is strictly contractive and has no fixed points. How about the example T : [1, 00)  [1,00), x I-t x + l/x, with usual metric? Can we say that this map has no fixed point? Can we say that this map is not a contraction whereas T : [1,00)  [1,00), with x I-t (75/76)(x + l/x), a contraction map? All of the above examples show that 'A contractive self mapping of a complete metric space need not have a fixed point. ' If Tx = sin x or cosx, then T'(x) exists and IT'(x)1 < 1 for all x E III Therefore, by Mean value theorem, T satisfies the Lipschitz condition so that T is uniformly continuous on III In fact, a direct calculation gives that I sinx-sinyl = 21 sin((x-y)/2) cos((x+y)/2)1 < 21 sin((x-y)/2)1 < Ix-yl 
212 Chapter 4: Contraction Mappings and Applications (Also, we remark that this inequality is a simple consequence of the Mean Value Theorem). Therefore, we obtain that T : IR  [-1, 1], x I-t sin x, is a nonexpansive map, and hence uniformly continuous on ]R (Recall that j(z) = sin z is not uniformly continuous on C). Clearly, this function is not distance preserving and so not an isometry. However, Tx = sinx 2 is not uniformly continuous on IR, because for x n = V2n1r and Yn = y 2n1r + 1r /2 we have IX n - Ynl  0, and ITx n - TYnl = 1 for n > 1. This observation shows that the function Tx = sin x 2 cannot be a contrac- tion on IR. Note that T : IR  IR, x I-t cos x or SIn x, is not a contraction, but the map T : IR  IR, x I-t 0.999 cos x or 0.999 sin x, is a contraction. We remark that in the above examples satisfying the weaker property (4.16), the corresponding metric space is not compact. Further, T satisfying the property (4.16) need not have a fixed point. However, if X is compact then T has a unique fixed point (see Exercise 4.29). 4.3 Applications to Differential Equations The most important applications of the contraction mapping theorem are to function spaces. Consider the ordinary differential equation for the Cauchy problem: (4.17)  = f(x,y) subject to y(xo) = Yo. Let j be a continuous function in a neighbourhood G of U o = (xo, Yo) in ]R2, say G = [a,b] x IR. Then we say that Y = cP(x) is a solution of (4.17) if following conditions are satisfied. . cP is differentiable on [a, b] . cP'(x) = j(x, cP(x)) for x E [a, b] · cP(xo) = Yo. 
4.3. Applications to Differential Equations 213 Thus, the Fundamental Theorem of Calculus is equivalent to the following: 4.18. Proposition. The function 4> : [a, b]. IR is a solution of (4.17) iff 4> satisfies the integral equation ,p(x) = Yo + (Z f(t, ,p(t» dt, x E [a, b). lzo 4.19. Theorem. Suppose that f is a continuous function on the neighbourhood G = [a, b] x IR C IR 2 and satisfies ( the Lipschitz condition in yon G) If(x, y) - f(x, y')1 < a:ly - y'l fo all (x, y), (x, y') E G. Assume that Xo E ( a, b) and Yo E III Then the ordinary differential equation 4>' (x) = f(x, 4>(x)) with initial condition 4>(xo) = 4>0, has a unique solution 4>(x) in the function space ( C [a, b], II · II 00 ) · Proof. Let X = (C[a, b], II · 11(0)' Then the differential equation with the given initial condition is equivalent to the integral equation described by the map T : X  X such that Tg(x) = ,po + (Z f(t, g(t» dt. lzo H we let Sg(x) = J: o f(t,g(t)) dt, then the discussion in 4.9 clearly implies that sng = (n  1)! 1: (x - t)n-l f(t, g(t» dt and hence, we have Tng(x) = Tn-lr/Jo + (n  1)! 1: (x - t)n-l f(t, g(t» dt. Now, it can be easily shown that for all n E N and for g,h EX, ITng(x) - Tnh(x)1 < o:lx -,x o1n doo(g, h). n. Indeed ITng(x) - Tnh(x)1 - 1: (n- ))l [f(t, g(t» - f(t, h(t») dt {Z Ix - tl n - 1 < 1zo (n _ 1)! o:lg(t) - h(t)1 dt < d ( h) Ix - xoln a: 00 g, , n. < a:lb - al n d ( h) , 00 g, . n. 
214 Chapter 4: Contraction Mappings and Applications But, for sufficiently large value of n, alb - al n 1 , <, n. and so Tn is a contraction on C[a, b]. Thus, T has a unique fixed point in C[a, b], which is the unique solution to the differential equation and the proof is complete. _ 4.20. Example. Note that the solution of the differential equation in Theorem 4.19 is the limit function in the iteration defined by </>n+l (x) = </>0 + 1 x f{t, ,pn{t)) dt, for n = 1,2, . . .. XQ For example, consider the simple initial value problem dy 2 - = m x + my subject to y(O) = 0, dx where m is a nonzero real number. Here Xo = 0 and f(x, y) = m 2 x+my and it is easy to see that Theorem 4.19 is applicable on any closed interval [a, b] which contains the origin. Hence, this differential equation has a unique solution in IR. Indeed the above iteration formula (with </>0 = 0) yields that {X {X (mt)2 ,pl - 10 f{t, r/Jo) dt = 10 m 2 t dt = 2 ' {X {X . (mt)2 (mt)3 ,p2 - 10 f{t,,pl (t)) dt = 10 [m 2 t + m,pI(t)] dt = 2 + 3! and so on. In this case, we obtain a series for exp(mt) - 1 - mt which converges uniformly on bounded closed intervals. . As another example of the applications of the contraction mapping the- orem to differential equations, we state and prove Peano's (also Picard's) theorem which asserts the basic existence and uniqueness of the solution of ordinary differential equations under appropriate hypotheses. This the- orem is a generalization of the Theorem 4.19. We use a slightly different argument to complete the proof of this theorem. 4.21. Theorem. (Peano-Picard's heorem) If f is a continuous function in a neighbourhood G of U o = (xo, Yo) in IR 2 satisfying (the Lips- chitz condition in y on the neighbourhood G) If(x, y) - f(x, y')1 < aly - y'l for all (x, y), (x, y') E G, 
4.3. Applications to Differential Equations 215 then, on some neighbourhood I = [xo - 6, Xo + 6], there exists a unique solution cP(x) of the ordinary differential equation for the Cauchy problem: cP' (x) = f(x, cP(x)), x E I with cP(xo) = cPo. Proof. Let 6 > 0 be such that 0 < 6 < 1 and B[U o ; 6] c G. Choose 6 > 0 such that 0.6 < 1 and {(x,y) : Ix - xol < 6, Iy - Yol < a6} = I x J c B[U o ;6], where J = [Yo - 0.6, Yo + 0.6]. Now, since f is ,continuous on the compact set B[U o ; 6], f is bounded on B[U o ; 6]; that is, there exists a number M > o such that If(x,y)1 < M for all (x,y) E B[U o ;6]. We show that the differential equation has a solution on I. Let X be the set of continuous function 9 : I  J (Le whose graph is contained in the rectangle I x J) satisfying the initial condition g(xo) = cPo: X = {g : 9 is continuous on I, Ig(x) - g(xo)1 < M6 for x E I, g(xo) = cPo}. Then, by Proposition 3.54, X becomes a complete metric space under the supremum metric doo(g, h) = sup Ig(x) - h(x)l, g, hEX. xEI For 9 EX, we consider the map T : X  X defined by Tg(x) = ,po + r f(t,g(t)) dt 1xo and seek a function cP that satisfies the integral equation ,p(x) = ,po + r: f(t, ,p(t)) dt, 'i.e. T,p = ,p. 1xo We first check that if 9 E X, then so does Tg (so that T has the stated range). Clearly, T is continuous and ITg(x) - ,pol < r If(t, g(t))l dt < (Z Mdt < Mix - xol < M  1 x o 1xo so that Tg E X. Thus, T is a mapping from X into itself. Next we show that T is contraction mapping with k < 0.6 < 1. Indeed, for g, hEX, we have doo(Tg, Th) sup (Z [j(t, g(t)) - f(t, h(t))] dt xEI 1xo < sup (Z If(t, g(t)) - f(t, h(t)) I dt xEI 1 Xo 
216 Chapter 4: Contraction Mappings and Applications < o:sup r Ig(t) - h(t)1 dt, by Lipschitz condition, xEI J Xo < 0.6 sup Ig(x) - h(x)1 xEI - a6d oo (g,h) so that, since 0.6 < 1, T is a contraction mapping. By contraction mapping theorem, there exists a unique function 4> in X such that T 4> = 4>, or ,p(x) = ,po + r f(t, ,p(t)) dt Jxo which is exactly equivalent to a solution of the given initial value problem, namely, 4>'(x) = f(x, 4>(x)) on I with 4>(xo) = 4>0, and the proof is complete. . 4.22. Example. Consider the Fredholm Integral Equation for an unknown function f : [0, 1]  JR: (4.23) f(x) = ,p(x) + 1 1 k(x, y)f(y) dy, where 4> E C[O, 1] and k : [0,1] x [0, 1]  IR are given functions. Let C[O,I] be equipped with the maximum metric d oo (see also Example 5.79). Let an operator T be defined by T : C[O, 1]  C[O, 1], g(x) t-+ ,p(x) + 1 1 k(x, y)g(y) dy. Then the integral equation (4.23) is a fixed point equation in the form Tf = f. If M = sUPO<x<1 f01 Ik(x, y)1 dy < a, then for each g, h E C[O,I], we find that - - doo(Tg, Th) < sup (1 k(x, y)(g(y) - h(y)) dy xE[0,1] Jo < sup (1 Ik(x, y)llg(y) - h(y)1 dy xE[0,1] J o < M sup Ig(y) - h(y)1 0y1 - Mdood(g,h). This observation shows that if there exists an a < 1 with M < a, then the method of successive approximation will produce the solution to the Fredholm Integral Equation since the map T in this case is contraction on the complete metric space (C[O, 1], II · 11(0)' Hence there is a unique continuous function f satisfying (4.23) whenever M < a < 1. . 
4.4. Exercises 217 4.4 Exercises 4.24. Determine whether the following statements are true or false. Justify your answer. (a) The function f(x) = x n for x E (0, 1) and n E N is a contraction. (b) The function f(x) = cosx has a unique fixed point in [0,1r/2]. (c) The mapping T: [a,b]  [a,b], x I-t log x, has a unique fixed point. (d) Let f be an analytic function in a domain D C C. Let S C D be a connected compact subset of D and f maps S into itself such that If' (z) I <.1 for all z E S. Then there exists a unique fixed point in S. (e) The function f : [0, 1]  [0, 1], x I-t 1/(2 + x), is a contraction with the corresponding Lipschitz constant a = 0.25. (f) IT T l , T 2 : (X, d)  (X, d), then we have the inequality a(Tl 0 T 2 ) < a(T l )a(T 2 ), where a(T) is the Lipschitz constant of a mapping T. (g) Let Tl and T 2 be two a-contraction mapping of a metric space (X, d) such that d(Tlx, T 2 x) < /3 for all x E X and for some /3. IT Xl and X2 are two fixed points of Tl and T 2 respectively, then we have the inequality (3 d(Xl,X2) < 1 · -a (h) Let d and p be two metrics on X and suppose that there exist two positive constants c and C such that cp(x,y) < d(x,y) < Cp(x,y), for all x,y E X. IT T : (X, d)  (X, p), then Hm [ad(Tn)]l/n = lim [a (Tn)]l/n, n-+oo n-+oo P where ad(S) stands for the Lipschitz constant of S with respect to the metric d. (i) The mapping T : (0, 00)  (0,00), x I-t e- x + x, with usual metric is not a contraction. (j) IT a > 0 and X = [y'a/2,00), then he map x I-t (1/2)(x + a/x) is a contraction. (k) The map Tl : ]R -+ JR, x I-t cos( cos x), is a contraction but not the map Tl : IR  JR, X I-t sin (sin x). 
218 Chapter 4: Contraction Mappings and Applications (I) If a E IR with lal < 1, then the continuous map T: IR 2  IR 2 , X = (Xl,X2) I-t X = (2 + asinxl,acosx2), is a contraction with usual metric. (m) The mapping T : IR  IR, X I-t X - arctan( x) + 1r /2, has no fixed point in III (n) The map 1 : IRt  IRt , x I-t y a + x2, has no fixed point. Here a is a fixed positive real number. (0) If Xo > 0, then the sequence {xn} defined by 1 Xn+l = 1 + X n converges. (p) A nonexpansive map may not have a fixed point, or it may have more than one fixed point. (q) If B = {z E 1 2 : IIzll2 < I} is the closed unit ball on , 2 and T : B  B is defined_by z ' {Zn}nl t-+ h/ l-lIzlI,Zl,Z2,...,Zn'."}' then T has no fixed point. (r) If X = {z E 1 2 : IIzl12 < I}, then the continuous map ( 1 - IIzll ) T : X  X, {zn} I-t 2 ' Zl , Z2, · .. , has no fixed point. (s) If X = {/(t) E C[O,l] : 0 = 1(0) < I(t) < 1(1) = I} and if T : X  X is defined by I(t) I-t tl(t), then T is a nonexpansive map. (t) There exists a nonexpansive map T : X  X which has a fixed point in X. (u) There exists a nonexpansive map T : X  X having more than two fixed points in X. (v) If a space X has a fixed point property, then so does the -space Y which is homeomorphic to X. (w) If 1 : IR  IR is a contraction map, then there exists a closed interval [a, b] in IR such that 1 maps [a, b] into itself. (x) If a E IR is a fixed real number such that lal > 1, then the mapping 1 ; IR  IR, x I-t 1/(x 2 + a 2 ), is a contraction. 
4.4. Exercises 219 (y) Let T be a self map of a complete metric space X. If S : X  X and if STS-1 is a contraction map, then T must have a unique fixed point. (z) If T is a self map of a = {z E C: Izi = I} defined by T:aa, zt--+z n , nEZ, then, for n  1, T has In - 11 distinct fixed points given by {e 21rki /(n-1): k= 1,...,ln-ll}. 4.25. Use the Banach contraction principle to solve the equation g(x) = 0, where g(x) = x 2 - 4x + 1. 4.26. Let I(x) = x 3 + x 2 - 1. Using the fixed point theory idea, find the solution of the equation I(x) = 0 in the interval (0,1). 4.27. Let a, b be fixed positive real numbers with b < a + 1 < 2 and X = [1, 00 ). Does the mapping 1 from the usual metric space (X, d) into itself with x t--+ ax + (b/x) satisfy the hypothesis of contraction mapping principle on X? What is the fixed point of I? 4.28. Define T : }Rn  }Rn be defined by x t--+ Ax + b where T is an n x n matrix and b E }Rn is a given n x 1 matrix. Consider the following metrics (i) doo(x,y) = max1kn IXk - Ykl, ( n ) 1/2 (ii) d 2 (x, y) = E(Xk - Yk)2 , k=l (iii) d 1 (x, Y) = EZ=l IXk - Ykl, where x = (Xl, X2, . . . , x n ), Y = (Y1, Y2, . . . , Yn) E }Rn. In each case, find a condition on the matrix A which ensures that the corresponding operator T is a contraction. 4.29. Let (X, d) be a compact metric space. If T : X  X is strictly contractive, then T has a unique fixed point. 4.30. Find a condition under which nonexpansive mappings have fixed points. 
Chapter 5 Linear Operators on N ormed Spaces As we see in Section 5.1, the finite dimensional normed spaces are much simpler than the infinite dimensional normed spaces. Indeed, the complete- ness property is inherent in all finite dimensional spaces, see Theorem 5.18. However, in the subject of analysis, infinite dimensional spaces are more important and therefore, special attention must be paid to the norms in question. We start with the properties of finite dimensional spaces in Sec- tion 5.1 in which we also discuss characterization theorems for equivalent norms. In Section 5.2 we introduce the notion of direct sums, complemen- tary subspaces and projections on to subspaces. In Section 5.3, we prove Riesz Theorem and its consequences. In Section 5.4, we briefly discuss the Weierstrass approximation theorem whereas in Section 5.5, we introduce the notion of Schauder basis. Recall that a mapping from a normed space X into a normed space Y is called an operator. A mapping from X into the scalar field IF is called a junctional. In Section 5.6, we study certain bounded linear operators and in particular bounded linear functionals. These are in fact continuous maps and therefore the collection of all bounded linear operators possess the structure of a vector space, see Theorem 5.70. There are various types of special operators which are not linear (eg. 'sub linear' operators) although we do not discuss them in this book. We consider the completion of normed spaces and the quotient spaces in Sections 5.8 and 5.9, respectively. The Open Mapping Theorem (or equivalently, the Banach theorem on the continuity of the inverse operator), the Uniform Boundedness Principle (which is also called Banach-Steinhaus theorem) together with the Hahn- Banach Theorem are considered as the "three pillars of junctional analysis" . The first two of these are fairly easy consequence of the Baire Category Theorem which is indispensable. We discuss the Baire Category Theorem (see Theorem 5.96) in Section 5.10. Recall that an operator T E L(X, Y), 
222 Chapter 5: Linear Operators on Normed Spaces where X and Yare Banach spaces, is invertible if it maps X bijectively to Y = RT, the range space of T. We consider the Open Mapphlg The- orem in Section 5.11 which insures that the inverse of such an operator is automqtically continuous-this result is known as the Banach theorem on the continuity of the inverse, see Theorem 5.107. In Section 5.12, we es- tablish the Closed Graph Theorem via the Open Mapping Theorem. In Section 5.13, we state and prove the Uniform Boundedness Principle which was obtained by S. Banach and H. Steinhaus in 1927 and therefore, this re- sult is also known as the Banach-Steinhaus Theorem. Section 5.14 gives the classical proof of Hahn-Banach Theorem and derive a number of corollaries. 5.1 Finite Dimensional Normed Space Most engineering analysis deals with the solution of certain equations on finite dimensional space. It is therefore interesting to study the properties such as continuity and boundedness of certain linear transformations be- tween finite dimensional spaces. If X is a finite dimensional normed space over the field F and if B = {V1, V2, . . . , v n } is a basis for X, then for each x E X we have n X = E ajvj for some scalars aj E F. j=1 If we define n IIxll(a) - aElajl (a > 0) j=1 ( ) IIp IIxll p i ; laj I P , P > 1, IIx II 00 - max la.1 1 < .<n 3' _3_ then each of the above defines a norm on X. These examples show that there are infinitely many norms on the same vector space. We remark that any two open balls in a normed vector space V (finite or infinite) are homeomorphic. Indeed, {x E V : IIx-xoll < 6} = {xo+y E V : lIyll < 6} = {xo+6z E V : IIzll < I} so that B(xo; 6) = Xo + 6B(0; 1). Thus, x t--+ xo+6x is a homeomorphism from B(O; 1) onto B(xo; 8). Because of this reasoning, open unit balls and unit spheres play a significant role in the study of normed spaces as we see in many theorems in this chapter. 
5.1. Finite Dimensional Normed Space 223 First we discuss the notion of equivalent norms and important basic results associated with this topic. In particular, these results show that changing from one norm to an equivalent norm does not alter the convergence and completeness, see Corollaries 5.7 and 5.8. 5.1. Equivalent norms (see Definition 2.67). Two norms 11.111 and II · 112 on a normed space X are said to be equivalent iff the corresponding metrics are equivalent, Le. iff they generate the same topologies. Let B1 (a; f) and B 2 (a; f) denote respectively the balls of radius f and center at a in (X, II · Ih) and (X, II. 112)' 5.2. Lemma. Let II .Ih and 11.112 be two norms on a vector space X. Then B 1 (0; 1) C B 2 (0; 6) for some 6 > 0 iff IIxll2 < 611xlh for all x EX. Proof. Suppose that B 1 (0; 1) C B2 (0; 6) for some 6 > o. Assume the contrary that there exists a point Xo E X such that IIxoll2 _ > 611xolh for some d > O. Set y = xo/llxolh. Then, by the homogeneity condition of the normed space namely (N2), we have 1111112 = ::::::: = cb > 6 for some c > 1, and lIylclll = lie < 1, so that the point y/c is in B 1 (0; 1) but not in B 2 (0; 6), since Ily/cll2 = 6. This is a contradiction. However, the converse part is trivial. - Note that the two way implication '<==}' in Lemma 5.2 is not true, in general, if we replace norms by metrics. For example, we consider two metric spaces (JR, d) and (IR, p) where d and p are the discrete and the Euclidean metrics, respectively. Then for all x, y "E JR, we have {OJ = Bd(O; 1) C Bp(O; 6) = (-6,6) for each 6 > O. However, since d( x, y) < 1 for all x, y E JR, there exists no 6 > 0 such that the inequality d(x, y) < 6p(x, y) = 61x - yl holds for all x, y E III 5.3. Theorem. (see Proposition 2.70) Two norms 1I.lh and 11.112 on a normed space X are equivalent iff there exist two positive numbers c and C, independent of x EX, such that (5.4) cllxl11 < IIxll2 < Cllxlh for each x E X. (This is equivalent to the fact that the identity operator I : (X, II · 111)  (X, II · 112) is an homeomorphism.) 
224 Chapter 5: Linear Operators on Normed Spaces Proof. Suppose that (5.4) holds for some c, C > O. Let a E X and f > o. Then, by Lemma 5.2, it follows that IIxll2 < Cllxlh <==} B 1 (0; 1) C B 2 (0; C) <==} a + (C /f)B1 (0; f/C) C a + (C /f)B 2 (0; f) <==} B 1 (a;f/C) C B 2 (a;f) and similarly, we have B 2 (a; Cf) C B 1 (a; f) <==} IIxlh < c- 1 I1xIl2' Thus, the topologies induced by the norms 1I.lh and II .112 are the same. - The size of the constants of equivalence c, C in (5.4) is often useful in the study of fixed point theory. For example, in order to apply the contraction mapping theory to the map T : (X, II . II)  (X, II. II), we must have IITx - Tyll < allx - yll for some 0 < a < 1. On the other hand, if we know that II · 111 and II · 112 are two equivalent norms on X satisfying (5.4), then the contraction condition IITx - TyJh < allx - ylh implies that IITx - Tyll2 < CIITx - Tyll1 < Calix - ylh < Cc- 1 allx - y112' This observation shows that the map T may be a contraction with respect to II · Ih but need not be with respe.ct to II .112 unless Cc- 1 a < 1. Let us give a precise example to demonstrate the last fact. 5.5. Example. Consider T : }R2  ]R2 defined by ( X + Y x - Y ) T(x, y) = 2 ' 2 · It is easy to check that T is a contraction with a = 1/V2 and with respect to the Euclidean norm. However, if we consider T as a mapping from }R2 to }R2 with respect to I-norm, we find that T is not a contraction. Indeed, by the definition of the map T, we have T(2,0) - T(3, 0) = (1, 1) - (3/2,3/2) = -(1/2, 1/2) so that IIT(2,0) - T(3, O)lh = 1 and 11(2,0) - (3,0)111 :;: II( -1, O)lh = 1. . 
5.1. Finite Dimensional Normed Space 225 Obviously, if the inequalities (5.4) are satisfied, then the inequalities (5.6) C- 1 11x1l2 < IIxlh < c- 1 11x1l2 are also true for each x E X. It follows from the inequalities (5.4) and (5.6) that if X n  x in some norm, then so does in an equivalent norm. Thus, we have 5.7. Corollary. Two norms 11.111 and 11.112 on a normed space X are eqwvalent iff any sequence in X converges with respect to II · 111 converges with respect to II · 112 and conversely. We observe that if II · Ih is equivalent to II · 112 and II · 112 is equivalent to II · 113, then it is trivial to see that II · 111 is equivalent to II · 113. Thus, equivalent of norms is an equivalence relation. Also, we note that Theorem 5.3 and Corollary 5.7 do not extend to nonempty metric spaces with equiv- alent metrics since the property of Cauchy sequences is not invariant under equivalent metrics. In view of Theorem 5.3, we easily have the following corollary which requires no proof. 5.8. Corollary. Let the two norms 11.111 and 11.112 on a normed space X be equivalent. Then we have (a) (X, II. 111) is a Banach space iff (X, II. 112) is a Banach space. (b) A set is bounded in (X, II · Ih) iff it is bounded in (X,II'112). 5.9. Example. Let k(x) be a function such that 0 < c < k(x) < C for all x E [0,1], where c and C are some positive number independent of x. On C[O, 1], consider the L2- norm and the norm 11.11 defined by [ 1 ] 1/2 11/112 = 1 1/ (t W dt , [ 1 ] 1/2 11/11 = 1 k(t)l/(tW dt , respectively. Then, these two norms are equivalent. 5.10. Example. Now, we give an example of two norms in C[a, b] which are not equivalent. For f, 9 E C[a, b], we know that (5.11) III - gll1 = l b I/(t) - g(t)1 dt < (b - a)1I1 - glloo and (5.12) III - gll2 = (l b I/(t) - g(tW dt) 1/2 < Vb - a III - glloo. In particular, this observation shows that if a sequence of functions in C[a, b] converges with respect to the supnorm, then it converges with respect to 
226 Chapter 5: Linear Operators on Normed Spaces s 1 1 t o Figure 5.1: The graph of In (t) the L1 and L2_ norms. In terms of € - 6 concept, for every € and for ever y 10 E C[a, b], there exist 6 1 (with 6 1 = €/(b-a)) and 6 2 (with 6 2 = €/ v b - a) such that Boo (/0; 6 1 ) C B 1 (/o;€) and Boo (/0; 6 2 ) C B 2 (/0;€), respectively. Thus, every open sets in C[a, b] with the L1- norm and respec- tively with the L2- norm is open in C[a, b] with the supnorm. However, the converse part is not true. It suffices to show for a = 0 and b = 1. Define a sequence {In(t)} of functions in C[O, 1] (see Figure 5.1): in(t) = {  - nt if 0 < t < l/n if l/n < t < 1. Then, we compute 1 1 lIinlll = 2n  0 and II in 112 = V3Ti  0 as n  00 (Note that II/nlh is simply the area of the triangle with vertices (0,0), (l/n,O) and (0, 1)). On the other hand, for n = 1,2, . . ., we have II/nlloo = 1 so that the {In} does not converge to lo(t) = O. In terms of €-6 notation, we have In E B 1 [0; 1/2n] and In E B2[0; l/J3n] but In is neither in Boo [0; 1/2] nor in Boo [0; 1/v'3]. Thus, for each n = 1,2, . . ., it follows that B 1 (0; 1/2n) ct Boo (0; 1/2) and B 2 (0; 1/v'3?) rt. Boo(O; 1/V3). This means that there exists no 6 > 0 such that B 1 (0;6) C Boo (0; 1/2) and B 2 (0;6) C Boo(O;l/V3). Hence, the two norms 11.1100 and 1I.lh on C[a, b] are not equivalent. Similarly, II . 1100 and II · 112 on C[a, b] are not equivalent norms. - 
5.1. Finite Dimensional Normed Space 227 Finite dimensional real or complex vector spaces posses a simple struc- ture. In fact, if X is a finite dimensional vector space then . there exists a norm on X . all norms on X are equivalent . all norms on X are complete . every linear self map of X is continuous. We begi? with the following theorem which in particular yields the last consequence (see Proposition 5.16) as a corollary to this theorem. Moreover, this theorem shows that although one can define many different norms on finite dimensional linear spaces such as an and en, there is only one topology derived from these norms. However, we have already shown in Chapter 2 that for 1 < p < 00 the p-norms on ]in are all equivalent. 5.13. Theorem. Any two norms on a finite dimensional normed space X are equivalent. Equivalently, the inequalities of the type (5.4) in Theorem 5.3 holds for any two norms on X. Proof. Let X be a finite dimensional vector space, where dim X = n, B = {V1, V2, . . . , v n } is a basis for X, and that II · II is a norm on X. For each x EX, we have the unique representation n X = L ajvj for some scalars aj E IF, j=1 which gives an element x = (a1, a2, . . . , an) E r. Consequently, we have an one-to-one correspondence between the elements of X and r. Specifically, the map x = (a1, a2, . · · , an) E r I-t x = E j 1 ajvj is obviously linear and bijective. Now, ( ) 1/2 II x lb =  lajl2 and for each x = E j 1 ajVj EX, we can define ( ) 1/2 IIxll2 =  laj 1 2 which means that II x l12 = Ilxll2. Let us show that cllxl12 < Ilxll < Mllxll2. 
228 Chapter 5: Linear Operators on Normed Spaces Now, for each x E X n IIxll = L ajVj j=1 n < L lajlllvjll j=1 < (  IIVjIl2) 1/2 (  la jl2 ) 1/2 ( ) 1/2 = Mllxl12' where M = E j 111 v jl12 . Therefore, there exists a positive constant M such that (5.14) Ilxll < Mllxl12' for each x E X, which establishes the one half of the desired inequality. Now, we work in establishing the other inequality. Define s = { x = (a 1 , a2, . . . , an) E ]Ffl : II x 112 = I}, the 'unit sphere' for the Euclidean norm on r. Then S is closed and bounded, and hence is compact (by Heine-Borel teorem) with respect to the Euclidean norm. Define f : S  IR by n f( x) = f(a1, a2, · · · , an) = L ajvj = Ilxli. j=1 Since B is a linearly independent set and since x ES, Le. x :j:. 0, all aj cannot vanish simultaneously on S so that f( x ) > 0 on S. Moreover, If( x) - f( y )1 = IlIxll - lIylll < IIx - yll < Mllx - yll2 where the last inequality is a consequence of (5.14) and the fact that the norm is linear. This observation shows that f is continuous on S. Thus, f is a positive valued continuous function on the compact set S and therefore, f attains its minimum m at some point on the compact set S. Consequently, whenever X ES, we have f( x ) = IIxll > m. Note that m > O. Thus, for each 0 :j:. U = (C1, C2,. . . , cn) E r, (5.15) lIuli =  CjVj = lI u ll2f ( IIIJ > mll u ll2 = mllull2 which establishes the second half of the desired inequality. From (5.14) and (5.15), any given norm II · II is equivalent to the 2-norm II . 112. Since 
5.1. Finite Dimensional Normed Space 229 equivalence of norms is an equivalence relation, it follows that any two norms on X are equivalent. _ Theorem 5.13 does not hold for metric spaces. For example, the se- quence {l/n} in IR with usual metric converges to 0, but does not converge with respect to the discrete metric. So, IR with usual metric and IR with discrete metric are not equivalent. Also, as we have seen in Example 5.10, Theorem 5.13 does not hold for infinite dimensional normed spaces. 5.16. Proposition. Each linear map T from a finite dimensional normed space X into a normed space Y is continuous. Proof. Let X be finite dimensional normed space, and {V1, V2, . . . , v n } be a basis for X. Then for each x E X we have n X = Eajvj j=l n and Tx = E ajTvj, j=l for some scalars aj E IF, so that n n IITxll < E lajlllTvjl1 < M 1 E lajl, j=l j=l where M 1 = max1jn IITvjll. Since X is a finite dimensional normed space, all norms on X are equivalent. Since II · 111 defined by n IIxIl1:= Elajl j=l is a norm, there exists a constant M 2 > 0 such that IlxllI < M211xll. Thus, we have IITxll < M1M211xll from which the continuity of T follows. - 5.17. Corollary. Let X be a finite dimensional normed space and let r > O. Then, the closed unit ball B[O; r] = {x EX: IIxll < r} is compact. Proof. With the notation of the proof of Theorem 5.13 (see Equation (5.15)), we see that there exists a constant m > 0 such that m ( f; laj 1 2 ) 1/2 < f ; ajVj 
230 Chapter 5: Linear Operators on Normed Spaces for all x = 2:, ; 1 ajVj EX, where {V1, V2, . . . , v n } is a basis for X. In particular, x E B[O;r] => n '" a.v. L.J J J j=1 2 < r 2 n => m 2 2: lajl2 < r 2 j=1 => m 2 lajl2 < r 2 foreachj=1,2,...,n => lajl < r/m foreachj=1,2,...,n. Hence, B[O; r] is a closed subset of the compact set n S={x=2:ajVj: lajl < r/m, j=1,2,...,n} j=1 and this completes the proof. - 5.18. Theorem. Every finite dimensional normed space is Banach. In particular, each finite dimensional subspace of a normed space is closed. Proof. Suppose that Y is a finite dimensional subspace of a normed space X, where dim Y = m and m < n = dim X. Then Y is isomorphic with . Therefore, by Theorem 5.13, the given norm of X restricted to Y is equivalent to the Euclidean norm of . Thus, every Cauchy sequence for the given norm of X restricted to Y is a Cauchy sequence for the Euclidean norm and hence has a limit, since  is complete. In particular, Y is a complete metric space, and hence by Proposition 2.109(ii), Y is closed. _ Theorem 5.18 does not hold for an infinite dimensional subspace of a normed space. For example, if X = C[O, 1] and if Y c X is the set of all polynomials in IR with real coefficients then Y = span {1, t,..., t n ,...} and Y = C[O, 1] :j:. Y and hence Y is not closed. We have already seen in Chapter 2 that (C[a, b], II · 11(0) is separable while (100, doc) is not, see Examples 2.74. However, there are many infinite dimensional normed spaces which are separable. We first show the result for finite dimensional spaces. 5.19. Proposition. Each finite dimensional normed space over the field 1F is separable. In particular, lP(n) is separable for 1 < p < 00. 
5.2. Direct Sums and Complementary Subspaces 231 Proof. The statement is trivial if X = {OJ. Now, let dim X = n with n > 1, and {V1, V2, . . . , v n } be a basis for X. Then for each x E X we have n X = Lajvj, j=1 , for some scalars aj E IF. If IF = ]R, then the set XQ defined by XQ = { t aivj EX: ai E Q } J=1 is countable and is dense in X, see Example 2.74. Since Q is dense in ]R, given € > 0, we have laj - ail < nil:; II ' Thus, for each x E X and x' = E j 1 ajvj E XQ, we obtain n Ilx - x'il < L laj - ajlllvjll < €. j ":" 1 The case when IF = C follows in a similar fashion as discussed in Example 2.74. . 5.20. Corollary. The Banach space (eo, 11.11(0) is separable. We leave this corollary as an exercise. Is c a separable Banach space? Is eo a closed subspace of c? 5.2 Direct Sums and Complementary Subspaces We will introduce some notation that will be used in the sequel. Given two subspaces M and N of a vector space X, and x EX, we define x+N={x+n: nEN}, M+N={m+n: mEM,nEN}, and for each A E IF, AM = {Am: m EM}. The space X is said to be (algebraic) direct sum of M and N if X = M +N, and MnN = {OJ; or equivalently, we write X = M fBN meaning that each x E X is expressible uniquely (in the sense that there exist unique elements m E M and n E N) in the form x = m + n. In this case, the two subspaces M and N are called 
232 Chapter 5: Linear Operators on Normed Spaces (algebraic) complementary su.bspaces of (in) X. In other words, we say that M is called an algebraic complement (or just complement) of N. Given a single subspace M of X, we say that M is complemented in X if there exists a complementary subspace N, meaning that we can write X = M E9 N for some N. It is easy to see that the complementary subspaces need not be unique. For example, if M - N 1 - N 2 - { (x, 0) E IR 2 : x E }R} {(O,y) E]R2 : y E IR} { (y , y) E IR 2 : y E JR} then, we see that each (x, y) E IR can be written as (x,y) = (x,O) + (O,y) = (x - y,O) + (y,y) = (O,y - x) + (x,x) and hence, we have the decomposition }R2 = M E9 N 1 = M E9 N 2 = N 1 E9 N2. In fact, there are uncountably many subspaces M, N in }R2 such that IR 2 = M E9 N. Also, in this example, we observe that M n N 1 = M n N 2 = N 1 n N 2 = {OJ and M + N 2 = JR2 . Suppose that X = M E9 N, where M and N are two subspaces of X. Then we may define a map P : X  X, x I-t m, where x = m + n, so that Px = P(m + n) = m. Unique representation of x E X by x = m + n makes the map P well- defined. It is linear, since, for each a, 13 E IF and for all x = m + n E X and all y = m' + n' EX, P(ax + 13y) - P(a(m + n) + 13(m' + n')) - P(am + 13m' + an + 13n') - am + 13m' - aPx + 13Py. When necessary, we write PM (instead of P) to indicate the dependence of P on M and PM is called the projection onto the subspace M. 
5.2. Direct Sums and Complementary Subspaces 233 Clearly, the range of P is M. As Px EM, it follows that P(Px) = P(m) = m = Px, p2 = P, and therefore, the restriction PIM is then the identity map on M. Further, N is clearly the null space of P, and we have the following result. 5.21. Proposition. H X = M E9 N, then there is a linear map P on X with range M and kernel N and such that p2 = P. Thus, one can make the following definition. "Let X be a vector space over the field IF. A linear map P : X  X is called projection in (on) X if p2 = P, i.e P{Px) = Px for each x EX". Note that if P is a projection on X, then so is the map Q = 1 - P, since Q2 = (I - p)2 = 1 - 2P + p2 = 1 - 2P + P = 1 - P = Q. Now, we prove the converse of Proposition 5.21 . 5.22. Proposition. Given a projection P : X  X, we have X = M E9 N with M = Rp and N = Np. Proof. Let P be a projection on X and Q = 1 - P. Then, we have PQ = P(1 - P) = P - p2 = 0 = (1 - P)P = QP. Thus, Q(x) E N p for each x E X and x = Px + (1 - P)x for each x E X so that X = Rp + Np. Suppose y E Rp n Np. Since y E Rp, y = Px for some x E X. Now, y E Np implies that 0= Py = p2x = Px = Y E Np. Consequently, y = 0 and hence, X = Rp E9 N p. . 5.23. Example. Consider X = Coo c 1 00 , the vector space of all finitely supported sequences {zn} n 1. Let {An} n 1 be a sequence of scalars. Define T : X  X by the formula PZ = z' where Z == {Zn}n>1 and z' = {Z}n>1 with - - { o z' = n Zn + A(n+1)/2Zn+1 if n = 2k , kEN. if n :;: 2k - 1 Thus, P{ {zn}) = {Z1 + A1 Z 2, 0, Z3 + A2 Z 4, 0, Z5 + A3 Z a,. · .} 
234 Chapter 5: Linear Operators on Normed Spaces so that Rp = { { Zk} EX: Z2k = 0 for each kEN} and Np = {{Zk} EX:" Z2k-l + AkZ2k = 0 for each kEN}. We observe that P(PZ) = P(z') = z' = Pz, Le. p2 = P, and therefore, P is a projection. Finally, it is easy to show that P is bounded iff {An} is a bounded sequence. Moreover, if the sequence {An} of scalars is bounded then it can be easily verified that IIPII = 1 + II {An} 1100. . 5.3 Riesz Theorems Let Y be a proper subspace of the Euclidean space }Rn. Suppose x E }Rn is a unit vector orthogonal to Y, Le. IIxII2 = 1 and the dot product x.y = 0 for all 0 :j:. y E Y. Then the application of Pythagorean theorem to Y gives that the distance from x to the subs pace Y is 1, Le. dist (x, Y) = 1. This result holds for arbitrary finite dimensional spaces, see Theorem 5.33 (The use of the term 'orthogonal' will be discussed in Chapter 6). However, this property does not hold for general normed space, see Theorem 5.25. Now, we shall discuss the issue of nearest points in detail. If Y is a nonempty subset of a normed space X and Xo EX, then we say that the point y* E Y is nearest to Xo if lIy* - xoll = dist (xo, Y) = inf lIy - xoll. yEY In other words, the element y* is called the best approximation in Y to Xo. We observe that if Xo E Y, then Xo itself is the (unique) nearest point to Xo in Y. On the other hand, if Y is closed then this infimum exists and is positive which happens when Xo ft Y, but such an element (best approximation) does not exist, in general. Moreover, even if nearest points exist, they need not be unique, as the following Example 5.24 shows. 5.24. Example. Let X = }R2 with the supnorm lI(xl, x2)1I00 = max{lxll, I X 21} and Y = {(Xl,O) E }R2 : Xl E }R}. Observe that in the subspace Y, the subspace topology on Y is same as the Euclidean topology. Then Y is a closed subspace of X and the point a = (0,1) E X has infinitely many nearest points in Y; in fact, for y = (Xl,O) E Y with IXll < 1, lIa - ylloo = II( -Xl, 1)1100 = max{lxll, I} = 1 
5.3. Riesz Theorems 235 and hence, every point in Y with Xl E [-1,1] is nearest to (0,1). Similarly, every point (Xl, 0) E Y with Xl E [0,2] is nearest to (1,1). However, for example, if the maximum norm considered above is replaced by the Euclidean norm then the nearest point is unique (why?), see also Theorem 6.74. In fact, if a = (0,1) and y = (Xl, 0) then lIa - yll2 = 1I(-Xl,l)lb = VX + 1 so that with respect to the Euclidean norm dist (a, Y) = inf V x + 1 = 1 {::::::} Xl = O. xlER Hence, (0,0) is the unique best approximation in Y to a = (0, 1). Similarly, if b = (1, 1), and y = (Xl, 0) then, with respect to the Euclidean norm, we have dist (b, Y) = inf v' (1 - Xl)2 + 1 = 1 {::::::} Xl = 1 (Xt,O)EY so that (1,0) is the unique best approximation in Y to b = (1, 1). We also observe that one can construct examples of this type with X = an with respect to the maximum metric and the Euclidean metric, respectively. Consider the subspace Y = {(Xl, Xl) E }R2 : Xl E IR} C X = IR 2 with I-norm, and an element e = (1, -1) E X\Y. Then, for all y - (Xl, Xl) E Y, we have lie - ylh = 11(1 - Xl, -1 - xl)lh = 11 - xII + 11 + xII > 2 so that with respect to I-norm dist (c, Y) = inf {II - xII + 11 + xII} = 2 {::::::} IXll < 1 xlER and hence, every point (Xl, Xl) E Y, with Xl E [-1,1], is nearest to the point c = (1, -1). . It is therefore natural to ask: Under what conditions on Y will nearest points always exist? When is it unique? These are two important questions which have practical importance too. In fact, the following remarkable result due to Riesz is often helpful in providing proofs of several results in the study of normed spaces. The normalized element X A (unit vector) such that dist (x A , Y) = 1 in some sense is perpendicular to Y and is, therefore, called ortogonal to Y, and (The use of the term 'orthogonal' will be discussed in Chapter 6.) It is because of this reasoning, the following theorem due to Riesz is called "result about almost orthogonal element" or "almost perpendicularity." 
236 Chapter 5: Linear Operators on Normed Spaces 5.25. Theorem. (Riesz) Let Y be a proper closed subspace of a normed space X over IF. Then, for each A E (0,1), there exists a point x).. E X \Y (not necessarily unique) such that IIx)..1I = 1 and dist (x).., Y) = inf IIx).. - yll > A. yEY Proof. Choose an arbitrary element x E X \Y, and let d = inf yEY IIx- yll. Since Y is closed, it follows that d > 0; otherwise, x would be a limit point of Y so that x E Y, a contradiction. Since d > 0 and d/ A > d, by the definition of infimum, there exists a Yo E Y with (5.26) d < II x - Yo II < d / A. Define x-Yo x).. = . IIx-yoll Then IIx)..1I = 1 (Again, we observe that x)..  Y; otherwise, the point x = Yo + IIx - Yo II x).. would belong to Y, a contradiction). Further, for an arbitrary y E Y we have Ilx).. - yll - IIx - {Yo + yllx - yoll} II IIx-yoll IIx-y'li - IIx - Yo II , where y' = Yo + yllx - yoll E Y, d > IIx _ Yo II ' by the definition of d and (5.26), d > dj A = A, by (5.26), and the proof is complete. . 5.27. Corollary. (Compare with Corollary 5.17) Let X be a normed space such that the closed ball B[xo; r] = {x EX: IIx-xoll < r} is compact for some Xo E X and r > o. Then X is finite dimensional. Proof. Note that B[xo; r] = Xo +rB[O; 1] and, since the translation and homogeneity are norm continuous, therefore the compactness of B[xo; r] implis that the closed unit ball B[O; 1] = {x EX: IIxll < I} is compact. Thus, it suffices to prove the corollary for the closed unit ball. Suppose on the contrary that the normed space X is infinit,e dimen- sional. Let Xl E X be an arbitrary element with IIX111 = 1. Then we have a one dimensional subspace Y 1 = span {Xl} of X, which is proper and closed (see Theorem 5.18). By virtue of Theorem 5.25, there exists a point X2 E X\Y 1 such that IIX211 = 1 and IIX2 - yll > 2- 1 for all y E Y 1 . 
5.3. Riesz Theorems 237 In particular, IIX2 - xlii > 2- 1 . Then, Y2 = span {Xl, X2} is proper closed subspace of X so that there exist a point X3 E X\Y 2 such that IIx311 = 1 and IIX3 - yll > 2- 1 for all y E Y2. In particular, we have IIX3 - xlii > 2- 1 and IIX3 - x211 > 2- 1 . Continuing inductively, we see that there exists Xk E X, IIxkll = 1 with IIXk - yll > 2- 1 , for each y E Yk-1 = span {Xl, X2, · · · , Xk-1}' . In particular, II X k - X j II > 2 -1 , for each j = 1, 2, . . . , k - 1 which holds for each intger value of k as X is infinite dimensional. Thus, we obtain a sequence {xn} of elements on the unit sphere S(O; 1) which has no convergent subsequence (In fact, if there exists a subsequence X n "', then X n '" is Cauchy so that there exists an integer p such that IIx ni - x nj II < 3- 1 ' for all i, j > p. But, 2- 1 < IIx np - x np + 1 11 < 3- 1 contradicts the fact that B[O; 1] is com- pact). Hence, X must be finite dimensional. - .-- From Corollaries 5.17 and 5.27, we conclude that "a normed space X is finite dimensional iff the closed unit ball B[O; 1] = {x EX: IIxll < I} is com pact." 5.28. Corollary. A compact set K in an infinite dimensional space X is nowhere dense. Proof. Recall that if K is compact, then it is closed. IT it has an interior point, then it contains a closed ball B[xo; r] for some Xo E K and for some radius r > 0 which is compact (since a closed subset of a compact set is compact). By Corollary 5.27, X is a finite dimensional space which is a contradiction. - We can formulate the above discussion as follows: 5.29. Theorem. Let X be a normed space. Then the following statements are equivalent: 
238 Chapter 5: Linear Operators on Normed Spaces (a) X is finite dimensional. (b) The closed ball B[O; 1] = {x EX: Ilxll < I} is compact. ( c) Every bounded sequence has a convergent subsequence. (d) The unit sphere S(O; 1) = {x EX: Ilxll = I} is compact. Proof. (a)  (b) is Corollary 5.17. (b) => (c) is trivial, because if {xn} is a bounded sequence in X then there exists an M such that IIxnll < M for all n so that, by (b), the set {x EX: IIxll < M} is compact, since x I-t Mx is continuous and hence {xn} has a convergent subsequence. (c) => (d) If {x n } is a sequence in the unit sphere S(O; 1), then it is bounded and hence has a convergent subsequence in S(O; 1), say x n1e  a as k  00. Since a norm is a continuous function, we have IIxn1e II  lIall as k  00 which implies that lIall = 1, as IIxn1e II = 1 for each k. This means that a E S(O; 1) and hence, S(O; 1) is compact. (d) => (a) follows from the proof of Corollary 5.27. . It is important to note that Theorem 5.29 (known as local compactness theorem) has various applications. Further, from Corollary 5.27 we see that for any infinite dimensional normed space the closed unit ball is not compact. 5.30. Example. We know that (eo, 11.1100) is an infinite dimensional normed space (see Corollary 3.35). Then the closed unit ball B[O; 1] = {x E eo : IIxll < I} is not compact. On the other hand, we can directly prove that the unit sphere S(O; 1) = {x E eo : IIxll = I} is not compact which would then imply that eo is infinite dimensional. Let us show that unit sphere S(O; 1) is not compact. For this, we recall the definition of Kronecker 'delta' symbol defined on an indexed set A: { 0, 6o.{3 = 1, For each j > 1, we consider 0:/3 , 0:, /3 E A. 0:={3 ej := {6ij}i1 = {O,O,..., 1,0,...} where 1 appears in the j-th place. Clearly, ej E S(O; 1) for each j > 1 and, for each m  n, we have Ile m - enll oo = 1. Therefore, distinct terms of the sequence {en}nl are at a distance 1 from each other and hence, the sequence {en}nl has no convergent subsequence. 
5.3. Riesz Theorems 239 Similarly, from Tnorem 5.29, it can be easily checked that the closed unit ball B[O; 1] is not compact in C[a, b] and also in lP for 1 < p < 00. In this way, we can conclude that the spaces C[a, b] and lP (1 $ p < 00) are infinite dimensional. . The Riesz Theorem states that for any closed proper subspace Y of a normed space X, there exist points in the unit sphere S(O; 1) = {x EX: IIxll = I} whose distance from Y is as near as we please to 1 (but not 1). On the other hand, there exists a point x E X on the unit sphere whose distance from Y is exactly one if Y is finite dimensional (see Theorem 5.33). However, this is not necessarily be true, in general, for infinite dimensional subspaces; Le. there does not exist a point on the unit sphere whose distance from an infinite dimensional subspace Y is exctly equal to one, as the following example proves: 5.31. Example. Consider two proper closed subspaces X and Y of (C[O, 1], 11.11(0) defined by X - {f E (C[O, 1],11,11(0) : f(O) = OJ, Y - {J EX: L 1 f(t) dt = O}. Clearly, Y is a closed subspace of X. Indeed; if we let G{f) = L 1 f(t) dt then G is a linear functional on X and IG{f)1 < L 1 If(t)1 dt < IIflloo. Therefore, G is continuous and Y = {f EX: G(f) = O} is a closed subspace of X. Moreover, Y is proper, because, for example f(t) = t, t E [0, 1], belongs to X \ Y. Now, we show that there does not exist f E X \ Y with IIflloo = 1 and d(f, Y) = 1. - Suppose on the contrary, there exists such a funtion fl',E X with IIf11100 = 1 and (5.32) IIf1 - glloo > 1 for each 9 E Y. Now, for F E X \Y, we let 1 1 h(t) dt C - 0 - 1 L F(t) dt 
240 Chapter 5: Linear Operators on Normed Spaces Then 11 - cF E Y and, by (5.32), 1 < 11/1 - (11 - cF)lloo = IclllFlloo which, by substituting the value of c, gives 1 1 F(t) dt < 1 1 h (t) dt 1IFll00o We can allow 1 1 F(t) dt as close to 1 as we please, and still have I IF II 00 = 1 (For example, Fn(t) = t 1 / n as n  00 does the job!). Thus, 1 1 h(t) dt > 1. But, since 11 is continuous with 11/11100 = 1 (Le. If1 (t)1 < 1) and 11 (0) = 0, it can be geometrically seen that 1 1h (t)dt <1. This inequality is a contradiction and therefore, the assumption in (5.32) is not possible, and we complete the proof. . Consider the space X = (C[O, 1], 11.11(0) and, a subspace Y = Pn{lR), the set of all polynomials of degree not exceeding n in III Then Y is a closed subspace of X with dim Y = n + 1. Our aim is to look for a polynomial y* (t) = E=l ak t k E Y satisfying the following condition: For each x EX, IIx - y*Jloo = dist (x(t), Y) < sup Ix(t) - y(t)1 for all y E Y, tE[O,l] that is y* yields the best approximation to x(t) E C[O,I] among all poly- nomials of degree at most n. The existence of such a best approximaing element is guaranteed by the following result. 5.33. Theorem. Let Y be a finite dimensional, proper subspace of a normed space X over IF. Then we have the following facts: (i) there exists y* E Y such that IIx - y* II = dist (x, Y) (ii) there exists a point Xl E X such that IIx111 = 1 and dist (Xl, Y) = 1. Proof. Let Y be finite dimensional. Then, by Theorem 5.18, Y is closed. Choose an arbitrary element x E X \Y. Then d = dist (x, Y) > o. Suppose that y* is a best approximation'. Since 0 E Y, it follows that the best approximation point y* must satisfy the inequality Ilx - y*11 < IIxll = IIx - 011. 
5.3. Riesz Theorems 241 Therefore, it suffices to look for y* among the elements y E Y satisfying the inequality IIx - yll < Ilxli. Moreover, Ilx - yll < IIxll ==} lIyll = II (x - y) - xII < Ilx - yll + IIxll < 211xll. This observations helps us to consider slightly a larger set of vectors, namely the ball in Y, K = {y E Y : Ilyll < 211xll} and try to prove the existence of y* E Y. (i) Clearly, K is a closed ball in Y (and a bounded subset of the finite dimensional subspace Y) and therefore, Theorem 5.29 assures that K is a compact subset of Y. First, we note that the map cP : K --+ IRt , Y t-t IIx - yll is continuous on the compact set K: IcP(Y) - cP(y')1 = Ilix - yll -llx - y'lil < lIy - y'lI, and hence, in particular, cP(y) = Ilx - yll attains a minimum value at some point y* E K. Secondly, if y E Y \K, then lIyll > 211xll and therefore, again by triangle inequality, we have IIx - yll > Illyll - IIxlll > IIxll = IIx - 011 > Ilx - y* II = dist (x, K). Here strict inequality in the above inequalities indicates that dist (x, Y) cannot be attained for y E Y \K. So, for all y E Y \K, IIx - yll > IIx - y*1I = dist (x, K) and hence IIx - y*1I = dist (x, K) = dist (x, Y); Le. y* is, by definition, a best 24 approximation to x in Y. (ii) First we observe that IIx - y*1I < II (x - y*) - (y - y*)11 = IIx - yll so that dist (x, Y) = dist (x - y*, Y). Now, as in Theorem 5.25, define x-y* Xl = IIx _ y* II. 24Notice that we say "a best" rather than "the best", see Example 5.24. 
242 Chapter 5: Linear Operators on Normed Spaces Then Ilxlll = 1, d . ( Y) - dist (x - y., Y) _ Ilx - y*1I - 1 1St Xl, - IIx - Y.II - IIx _ y"l1 - and the proof is complete. . 5.34. Definition. A norm II . lion a vector space X is said to be strictly convex or rotund if (5.35) x,y E X, X:F y, Ilxll = Ilyll = 1 => X+y 2 < 1, or equivalently, x+y x,y E X, Ilxll = Ilyll = 2 => x = y. We often say that the normed space X is strictly convex, with the understanding that the corresponding norm is strictly convex in the sense of the above definition. We have the following result. 5.36. Proposition. A normed space X- is strictly convex iff for each x,y E X, X:F y, Ilxll = Ilyll = 1, we have IIAX + (1 - A)yll < 1 for A E (0,1). In other words, in a strictly convex space X, the open line segment between any pairs of points on the unit sphere in X lies entirely inside the unit ball. Proof. Let x, y E X and x :F y with Ilxll = lIyll = 1. Suppose that (5.37) IIAoX + (1 - Ao)yll < 1 holds for some AO E (0,1). If A E (0, AO), then set J.t = AI AO and Zo = AOX + (1 - AO)y so that J.t E (0,1), Ilzll < 1 and /-LZ + (1 - /-L)Y =  (AOX + (1 - AO)Y) + (1 - :J Y = AX + (1 - A)Y. Therefore, by the triangle inequality, we have II AX + (1- A)yll = IIJ.tz + (1- J.t)yll < J.tllzll + (1- J.t)lIyll < 1 for A E (0, AO). If A E (AO, 1), then set J.t = (1 - A)/(1 - AO) so that J.t E (0,1) and /-LZ + (1 - /-L)x = 11   (AOX + (1 - AO)Y) + (  = : ) X = AX + (1 - A)Y. 
5.3. Riesz Theorems 243 Again, IIAX + (1 - A)yll < 1 for A E (Ao, 1). Thus, if X is strictly convex then (5.37) holds for Ao = 1/2, and hence, IIAx + (1 - A)yll < 1 holds for all A E (0, 1). The reverse part is trivial. _ 5.38. Definition. A norm 11.11 on a vector space X is said to satisfy the parallelogram rule/identity if IIx + yl12 + IIx - yll2 = 2(llxlI2 + Ilyll2) for x, y EX. In Chapter 6, we shall see that this identity holds only in real normed spaces in which the norm is generated by an inner product. Each of the IP-space and the LP[a, b]-space fails to satisfy the parallelogram identity if p :j:. 2, see 6.33 and 6.38. 5.39. Example. Let X be the space of all sequences of complex numbers Z = {Zk}kl such that 00 E I Z 2k IP < 00 and k=l 00 E I Z 2k+11 q < 00, k=O where p and q are fixed positive real numbers with p > 1 and q > 1. For Z E X, define ( 00 ) l/p ( 00 ) 1/ q IIZII=  IZ2kIP +  IZ2k+1lq · Then it can be easily verified that this defines a norm on X. Moreover, we see that Z = {Zk}kl E X iff Z(p) = {Z2k}k1 E lP and Z(q) = {Z2k+1}k1 E lq. Because of this observation, it follows that {Zn} is Cauchy sequence in X iff the corresponding subsequences {Z!r)} and {Zq)} are Cauchy in IP and lq, respectively. Since IP is complete for each p > 1, it follows that X is a Banach space. However, the parallelogram identity is not satisfied in X. Indeed, if Z = {I, 0, 0, . . .} and W = {O, 1,0,0,. . .}, then Z + W = {I, 1,0,0,.. .}, Z - W = {I, -1,0,0,...} and hence, IIZ + WI1 2 + IIZ - WII 2 = 8  4 = 2(IIZ1I 2 + IIWI1 2 ). . 5.40. Proposition. A normed space X which satisfies the parallel- ogram identity is strictly convex. 
244 Chapter 5: Linear Operators on Normed Spaces Proof. If X,Y E X, x :F Y, Ilxll = 1 = IIYII, IIx - yll > € > 0, and if parallelogram rule holds, then IIx + yll2 = 2(lI x 1l 2 + lIyll2) - IIx - yll2 < 4 - £2 < 4, I.e. x ; y < 1. and therefore, the mid point (x + y)/2 lies strictly inside the unit ball. . Next, we observe that the set of best approximation possesses a reason- able geometric property. 5.41. Theorem. Let Y be a subspace of a normed space X, and let x EX. Then the set Y x consisting of all best approximation to x out of Y is a convex set. Proof. Let d = dist (x, Y). If Y x = 0 or Y x contains only one element, then there is nothing to prove. Therefore, we assume that Y x contains more than one element. Let Yl, Y2 E Y x . Then IIx -Ylil = IIX-Y211 = d. Next, given 0 < A < 1, let y* = AYl + (1 - A)Y2' We want to show that y* E Y x . Clearly, y* E Y and so, we find that IIx - y*11 > d and IIx - y*11 - IIx - (AYl + (1 - A)Y2)1I - IIA(X - Yl) + (1 - A)(X - Y2)11 < Allx - Ylil + (1 - A)lIx - Y211 - Ad + (1 - A)d = d. Hence, Ilx - y*11 = d so that y* E Y x . . If Y x consists of more than one point, then it must contain an entire line segment joining any two points in Y x . This observation shows that Y x is either empty, or contains exactly one point, or contains infinitely many points. Thus, we have 5.42. Theorem. H a normed space X contains no line segments on any sphere 8(0;6) = {x EX: IIxll = 6}, then each best approximation ( out of any subspace) is unique. For practical purposes, one would like uniqueness and not just the exis- tence of a best approximation. Here we have two simple geometric condi- tions that permit us to claim the uniqueness of best approximation on 
5.3. Riesz Theorems 245 convex sets, see Theorem 6.74 and Corollary 6.78. Indeed, in normed spaces, the strict form convexity condition guarantees the uniqueness of best approximation on convex sets. The following result is a consequence of Theorems 5.41 and 5.121 (below). However, for a clear understanding of this, we include the proof here. 5.43. Corollary. If X has a strictly convex norm, then, for each convex subset Y of X and for each x EX, there exists at most one best approximating element to x out of Y; i.e. the set Y x , consisting of all best approximation to x out ofY, is either empty or consists of a single element. Proof. Suppose that d = dist (x, Y), and that there exist two best approximations Yl, Y2 to x. Then IIx - Ylil = IIx - Y211 = d, Yl, Y2 E Y x . Since Y is convex, (Yl + Y2)/2 belongs to Y. Further, d - IIx - Ylil IIx - Y211 2 + 2 x - Yl X - Y2 2 + 2 I Yl + Y2 - I X - 2 > d > so that x - Yl X - Y2 Yl + Y2 2 + 2 = x- 2 . It is easy to see that the last equality implies Yl = Y2. Indeed, if Yl :j:. Y2 then x - Yl X - Y2 2 :j:. 2 so that, by Definition 5.34, it follows that _ Yl + Y2 _ X - Yl X - Y2 d x 2 - 2 + 2 < which is a contradiction to the last identity. This observation proves the uniqueness of the best approximating element. _ 5.44. Theorem. The following statements are equivalent: (i) (X, II · II) is strictly convex (ii) If x, Y E X and Ilx + yll = IIxll + lIylI, then x = AY for some A > O. 
246 Chapter 5: Linear Operators on Normed Spaces Proof. (i) => (ii): Assume that (X, II · II) is strictly convex. Suppose that x,y E X and (5.45) IIx + yll = IIxll + lIyll. Without loss of generality, we may assume that x # 0 and lIyll > IIxll. Then 2 > x y n;rr+TIYTI - ( 11: 11 + 11: 11 ) - ( 11: 11 - 11:11 ) > Ilx + yll ( IIII - IIII ) y IIxll IIxll + Ilyll Ilyll (llyll - Ilxll) by (5.45) and lIyll > IIxll, - Ilxll Ilxlillyll - 2 which shows that x' + y' 2 , x , y = 1, with x = IIxll ' y = TIYTI' Because of (5.35), it follows that x' = y'. Therefore, (ii) holds. (ii) => (i): Assume that (ii) holds. To show that X is strictly convex, we let x, y E X, x # y, IIxll = Ilyll = 1 and Ilx + yll = 2 = IIxll + lIyll, Le. x+y = Ilxll = lIyll = 1. 2 Then, by (ii), we get that x = AY for some A > O. It follows that A==1 lIyll which means that x = y. Therefore, X is strictly convex . 5.46. Example. (i) Consider the scalar field F, viewed as a normed space over F. Then it can be easily seen that this is strictly convex. (ii) Let X = Co or 1 00 , x = el + e2, Y = el - e2 where el = {I, 0, 0,. . .} and e2 = {O, 1,0, . . .}. Then IIxli oo = lIylloo = II (x + y) /21100 = 1 but x # AY. Thus, neither Co nor 1 00 is strictly convex. (iii) In}Rn, the norms II . 111 and II · 1100 are not strictly convex. . 
5.4. Approximation in Function Spaces 247 5.4 Approximation in Function Spaces Our goal in this section is to state and prove an important result due to Weierstrass which in a simple form states that the set of all polynomial functions is dense in C[a, b]. This result can be formulated in the following precise form. 5.47. Theorem. (Weierstrass Approximation Theorem, 1885) Let f E C[0,1]. Then, for every € > 0, there exists a polynomial p such that SUPtE[O,l] If(t) - p(t)1 < €. Before we present the proof of Theorem 5.47, it is essential to make few remarks and illustrations. A direct extension of Theorem 5.47 is "Every continuous function on a nonempty compact subset D of C can be approx- imated by complex polynomials in z and z". Note that the polynomial in this case will be of the form n p(z, z ) = E al,m zlzm , where n E N, al,m E C. l, m=O Since z = II z iff Izl = 1, from the above observation, it follows that "Every continuous function on a unit circle all = {z E C : Izi = 1} in C can be uniformly approximated by the functions in the space {P: pz) = k )n ak zk , where n E NU {O}, a- n ,... ,an E C} ". Various other extensions of Theorem 5.47 is available in the literature which can be found in advanced texts on this topic. For the proof of the Weierstrass approximation theorem we need to show that for each f E C[O, 1] there exists a sequence of polynomials Pn such that lim n -+ oo Ilf - Pn II 00 = O. Also, we remark that the underlying interval [0, 1] in Theorem 5.47 is of no consequence here. The intervals [0,1] and [-1,1] are popular choices, but it hardly matters which interval we choose. In fact, given F E C[a, b] (-00 < a < b < 00) the function f defined by f(t) = F((b - a)t + a), t E [0,1], is an element of C[O,I]. By Theorem 5.47, given € > 0 there exists a polynomial p in [0, 1] such that sup If(t) - p(t)1 < € tE[O,l] which is equivalent to sup IF(t) - P(t)1 < €, where P(t) = p((t - a)/(b - a)). tE[a,b] 
248 Chapter 5: Linear Operators on Normed Spaces Note that the space P[a, b] of all real polynomials, P(x) = ao + alt + . . · + ant n , n = 0, 1,2, . . . , on [a, b] is invariant under translation and is an infinite dimensional sub- space of C[a, b]. Thus, by Theorem 5.47, we can approximate a real valued continuous function on [a, b] arbitrarily closely in modulus by a real valued polynomial in [a, b]. Since classical approximation theorems are often for- mulated in terms of dense sets in metric spaces, we can extend the above idea to a general metric space defined in terms of closure. Thus, one of the easy ways to explore the structure of a metric space is to look for sets whose closure is the whole space. Using Theorem 5.47 we see that the set Y = PQ[a, b] of all polynomials with rational coefficients is dense in C[a, b] with uniform metric. In fact, for each ak E IR and for every € > 0, there exists a point Tk E Q such that lak - Tk I < €. Form q(t) = "0 + Tl t + · · · + Tnt n , n = 0, 1,2, . . . which is in PQ[a, b]. By Theorem 5.47, for f E C[a, b] and for each € > 0 there is a polynomial p with real coefficients ak such that doo(f,p) = sup If(t) - p(t)1 < €. tE[a,b] Therefore, doo(f, q) < doo(f,p) + doo(P, q) < f + t laic - Tic I ( max It l ) Ic < Cf tE[a,b] k=O for some constant c. This shows that, Y is dense in C[a, b]. But, Y is countable, since Q and the union of a countable number of countable sets is again countable. Consequently, we have 5.48. Corollary. The space C[a, b] is separable. Also, it follows from Theorem 5.47 that each f E C[O, 1] has a distance 0 from P[O, 1]. Further, since not every member of C[O, 1] is a polynomial, we cannot expect a best approximating polynomial to exist for each f E C[O,l]. For example, the function f defined by { t sin(ljt) f(t) = o if t :F 0 ift=O is in C[O, 1] but cannot possibly agree with any polynomial in [0,1] because f has infinite number of zeros in [0, 1]. 
5.4. Approximation in Function Spaces 249 5.49. Example. Consider the function f : [a, b]  IR defined by f(x) = Ix - cl, c E [a, b]. We provide a direct method of getting a polynomial approximation for f on [a, b]. Without loss of generality we can assume a = 0 and b = 1 so that C E [0, 1]. First we let C E (0, 1/2] and write Ix  ci = {c 2 - [c 2 - (x - c)2]} 1/2 = c(l- y)1/2, with y = 1- ((x - c)/c)2 so that the resulting series expansion for Ix - cl is given by the series  (-1/2,k) k c ( 1 k ) y. k=O ' Here (a,O) = 1 for a :j:. 0 and (a, k) is the ascending factorial notation defined by (a, k) = a( a + 1) · · · (a + k - 1). Therefore, we can rewrite the series expansion as c [ I - f: Ckyk ] , k=1 where C1 = 1/2 and for k > 2 (-1/2,k) 1 1 3 2k-3 -Ck = (1, k) = - 2 · 4 · 6 ' 2k · Note that Ck > 0 for all k > 1 and Ck = ak-1 - ak where ao = 1 and 1 · 3 . 5 · · · (2k - 1) ak = 2.4.6...(2k) so that n L Ck = ao - an = 1 - an < 1. k=1 Thus, E  1 Ck < 00 and, therefore, the series 00 1- Lckyk k=1 converges absolutely and uniformly for Iyl < 1. Equivalently, we say that the series C[I-  Cdl-((X-C)/C)2)k], 
250 Chapter 5: Linear Operators on Normed Spaces converges uniformly to Ix - cl whenever 1-( x:c r <1 and hence, for Ix - cl < c, or equivalently, x E [0,2c]. Thus, the polynomial of partial sums converges uniformly to Ix-cion the interval [0, 2c], a fortiori in [0, 1]. The conclusion for c E (1/2,1) is similar if we replace c 2 by 1- c 2 . . The proof of Theorem 5.47 that we present here is due to the Rus- sian mathematician S.N.Bernstein in 1912, who constructed, for every I E C[O, 1], an explicit formula for a sequence of polynomials converging to I. These are called the Bernstein polynomials. 5.50. Definition. Let I be a function defined on the closed interval [0,1]. The polynomial (Bn(/))(x) defined by (Bn(f))(X):=Bn(X)=/(  ) ()xk(1_X)n-k, XE[O,1] is called the Bernstein polynomial (associated to I) of degree at most n. Here () denotes the usual Binomial coefficient defined by ( n ) n! k - k!(n - k)!. We remark that Bn(1 + g) = Bn(/) + Bn(g) and Bn(A/) = ABn(/) (A E IR). Moreover, Bn(/) > 0 whenever I > 0 and therefore, the map I I-t Bn(/) is linear, and positive. Now we are in a position to prove Theorem 5.47. Proof. We need some preparation. Let us start with the well-known Binomial formula (x + y)n = t (  ) xkyn-k. k=O Define 10 = 1, 11 = x, 12 = x 2 . We first show that (5.51) Bn/o = 10, Bnl1 = 11, Bnh - h = 11 - h = x(1 - x) . n n 
5.4. Approximation in Function Spaces 251 The Binomial formula for y = 1 - x gives (5.52) t (  ) Xk(l- x)n-k = 1 k=O so that Bnfo = fOe Differentiation of the Binomial formula with respect to x and then multiplication of the resulting equation by x gives i k()xkyn-k =nx(x+y)n-l and a similar operation on this equation yields t k 2 (  ) xkyn-k = n[(n - l)x(x + y)n-2 + (x + y)n-l ]x. k=O Substitution of y = 1 - x in the last two identities and the division by n and n 2 respectively give (5.53)  [  ] ()Xk(l- x)n-k = x, i.e. Bnlt = It, and, (5.54)  [  f ()Xk(l- x)n-k = (1-  ) x 2 +  x so that Bn h = (1 -  ) h +  It, i.e. Bn h - h =  (It - h). Thus, (5.51) follows. Finally, by (5.52)-(5.54), it follows that i [  -xf ()Xk(l-xt-k = (1-  ) x 2 +  X-2x2+X2 = x(\:-x) so that (5.55)  [  - xf ()Xk(l- x)n-k < 4 because x(l - x) has the maximum value 1/4 in [0,1]. Then, because of (5.52), we have IBn(f) - l(x)1 -  [I (  ) - I(X)] ()Xk(l- x)n-k <  1 (  ) - I(x) ()Xk(l- x)n-k. 
252 Chapter 5: Linear Operators on Normed Spaces Let f E C[O,l], M = SUPtE[O,l] If(t)l, and let € > O. By the uniform continuity of f on [0, 1], there exists a 6 > 0 such that If(y) - f(x)1 < € whenever y,x E [0,1] and Iy - xl < 6. Next, we observe that (5.56) 2M k 2 ( k ) 2M k 2 -€ - - - - x < f - - f( x ) < € + - - - X 6 2 n - n - 6 2 n holds for all x, kin E [0,1]. Indeed, if I(kln) - xl < 6 then (5.56) trivially follows from -€ < f(  ) -f(x) < f. On the other hand, if I(kln)-xl > 6 then (5.56) follows from the inequalities 2M k 2 ( k ) 2M k 2 -- - - x < -2M < f - - f( x ) < 2M < - - - x . 6 2 n - - n - -6 2 n Thus, for a given € > 0 and given x, we can split the sum into two parts: (i) those k's in the set K = {O, 1, 2, . . . , n} for which I(kln) - xl < 6 and name the subset of K which satisfies the last inequality as Kl (ii) those k's in K for which I(kln) - xl > 6 and name this subset as K 2 . Now, K = Kl U K 2 and obtain IBn(f)(x) - f(x)1 <  f (  ) - f(x) ()xA:(1- x)n-k < € L ()xk(1- x)n-k kEKl + 2:: L [  - xf ()XA:(1 - x)n-A: kE K 2 < €.1 + 2:: ( 4 )' by (5.52) and (5.55), M M - € + 2n6 2 < 2€, whenever n > 2€6 2 ' 
5.5. Schauder Basis 253 Thus, for f E C[O, 1], sup IBnf(t) - f(t)1 < 2€ tE[O,1] must hold for all sufficiently large n. Hence, lim Bnf(t) = f(t) n-+oo uniformly in [0, 1], and the proof is complete. . This theorem, in particular, implies that Bernstein polynomials are dense in C[O,l]. There are several other proofs and extension (in various forms) of this theorem in the literature. However, an interesting gener- alization of this theorem is the following classical result due to Bohman- Korovkin whose proof is essentially identical to the above proof due to Bernstein, see [Za, Chapter 2]. 5.57. Theorem. (Bohman-Korovkin) Let {Tn} be a sequence of positive linear operators from C[O,l] into itself. H lim n -+ oo Tn(f) = f for all f in the test set 8 =: {I, x, x2}, then lim n -+ oo Tn(f) = f holds for every f E C[O, 1]. 5.5 Schauder Basis Let {tPn}n1 be a nonzero sequence of elements in.'an infinite dimensional normed space X over the field IF. The sequence {tPn} is called a Schauder basis for X if each element x E X admits a unique sequence of scalars {an} in IF such that 00 x = L antPn n=1 with the series converging in norm to x; Le. II EZ=1 aktPk - xII  0 as n  00 (IT X is a finite dimensional normed space and dim X = n, then a Schauder basis in X is just any (vector space) basis and the above represen:" tation is to be understood as x = EZ=1 aktPk ). Clearly, a Schauder basis for X is linearly independent. Moreover, any Schauder basis has dense linear span, Le. the subspace span {tPk : kEN} = { t ak rPk : al,"., an E F, n E N } , k=1 containing a countable dense subset of X, namely the set { takrPk: k=1,2,...,n, nEN } , k=1 
254 Chapter 5: Linear Operators on Normed Spaces where the coefficients ak must be chosen in the following way: If IF = C, then a k = 0: k + i {3 k, 0: k , {3 k E Q, and if 1F = JR., then we assume {3k = O. Thus, we have 5.58. Theorem. A normed space X that is equipped with a Schauder basis is separable. For example, 1 00 cannot have a Schauder basis since 1 00 is not sepa- rable. During Banach's time, Schauder bases were constructed for all the familiar separable Banach spaces such as LP[a, b], lP (1 < p < 00), and C[a, b]. On the other hand, the converse statement of Theorem .5.58 is not true. In fact, in 1927, Schauder asked the following well-known "problem of a Schauder basis": Does a separable Banach space necessarily have a Schauder basis? The answer is no and in 1972, Per Enflo settled this long- standing open question by proving that there exist separable Banach spaces without Schauder bases. We now give some examples of Schauder basis. 5.59. Example. The sequence {en}n>l helps us to write each ele- ment Z = {Zn}nl in lP (1 < p < 00) as the sum Z = E  I Znen. To prove this fact, we set n Sn = LZkek = {ZI,Z2)... ,zn,O,.. .}. k=l We show that the sequence {sn} converges to z, where Z = {Zn}nl' In fact, n IIz - snllp = Z - L Zkek k=l ( 00 ) IIp P - k  1 IZkl P  0 as n  00; Consequently, 00 Z = lim sn =  Znen n-+ 00 L....J n=l in lP norm. Hence, {en}nl is a Schauder basis for lP (Note that {en} is not a Hamel basis because every Z = {zn} E lP cannot be written as a linear combination of a finite number of ej's). A similar argument can be applied to (eo, II · 11(0) and to conclude that { en} n  I is a Schauder basis for eo. We have already seen that 1 00 cannot have a Schauder basis. How- ever, the fact that {en}nl cannot be a Schauder basis for 1 00 is seen very concretely, because otherwise n Z - LZkek k=l  0 asnoo ==> sup IZkl  0 kn+l 00 
5.6. Bounded Linear Operators 255 which is not true for all Z E loo (as the elen1ent {I, 1, 1,. ..} shows!). We note that the limit transition is not possible for loo in the last one way implication. . 5.6 Bounded Linear Operators We shall study linear mappings T : X  Y, where X and Yare two normed spaces over the same field]F. For the sake of simplicity, we use the same symbol 11.11 for both the norms in X and Y especially when we study linear operators from X into Y. Then the operator norm or uniform norm of a linear transformation T in L(X, Y) is defined by IITII = sup{IITxll : Ilxll = I} (if it exists). We note that, we have written IITxl1 and IIxll in place of IITxlly and IIxllx, respectively. Recall that, L(X) := L(X, X). The linear operator T E L(X, Y) is called bounded if there exists an M > 0 such that ( 5.60) IITxll < Mllxll for all x E X. The set of all bounded linear transformations from the normed space X into the normed space Y will be denoted by B(X Y). If X = Y, we often write B(X) := B(X, X). The identity mapping, denoted by I, belongs to B(X). In (5.60), if IIxll = 1 then IITxll < M, and IITII is then the infimum of all such M satisfying (5.60). In other words, T is called a bounded operator if IITII is finite; otherwise called unbounded operator. Now, we state a simple criterion for the evaluation of the operator norm IITII for a bounded linear operator T. If there exists a 0 :j:. Xo E X such that equality holds in the inequality (5.60), then we have IITxoll = Mllxoll < IIT1l1lxo11 for some 0 :j:. Xo E X so that M < IITII. But, by the definition, IITII < M. Combining the last two inequalities, we obtain that M = IITII. As an immediate consequence of (5.60), we have 5.61. Proposition. T E B(X, Y) iffT maps Cauchy sequences into Cauchy sequences. Proof. The direct implication part follows trivially from (5.60). For the proof of reverse part, let T map Cauchy sequences into Cauchy sequences. Suppose to the contrary that T is unbounded. Then, there exists a sequence { x n } such that IITxnll > n 2 11x n ll for eah n and therefore, for those X n :j:. 0, we can form a subsequence {Yn}, where x n Yn = nllxnU' 
256 Chapter 5: Linear Operators on Normed Spaces Then llYn II = 1/n  0 so that {Yn} converges and hence, is a Cauchy sequence. On the other hand, liT II - IITxnll n 2 l1x n ll_ Yn - > - n nllxnll nllxnll so that {TYn} is unbounded and hence, {TYn} is not a Cauchy sequence, which is a contradiction. Hence, T must be bounded. - Later, in Theorem 5.66, we present several equivalent characterization for boundedness of the linear operator T. If Y = F, then the operators in B(X, IF) are called bounded linear func- tionals. Also, note that a bounded functional f on X satisfies the condition If(x)1 < Mllxll for all x EX, and for f E B(X,IF), II f II = sup { If ( x ) I : II x II = I}. Further, from (5.60), we see that a bounded linear operator maps bounded sets in X onto bounded sets in Y. This suggests the term "bounded op- erator", see also the equivalence in Theorem 5.66. Thus, the present use of the word "bounded" is different from the notion of boundedness for an ordinary complex function f on a set X in which f is. bounded would mean If(x)1 < M as x runs over the whole space X. Since f(AX) = Af(x), for all scalar values of A, we observe that no nonzero linear operator can satisfy the last condition. It is a simple exercise to show that only the zero-transformation has this extreme form of boundedness. We mention here that the linear functional is the special case of the so called convex functional: the mappings f from a normed space X into IR such that f(AX + (1 - A)Y) < Af(x) + (1 - A)f(y), x, Y E X, A E (0,1). If T : X  Y is linear, then, for any x :F 0, Y = x/llxli satisfies IlylI = 1, and therefore II Tx II ( X ) IIxll = T jj;jj = IITyll < IITII, which shows that (5.62) IITxll < IITII Uxll for all x EX. For x = 0, the last inequality is clear, since T(O) = T(O + 0) = T(O) + T(O), Le. T(O)=O. 
5.6. Bounded Linear Operators 257 Thus, if T is a bounded linear operator then (5.60) holds for any choice of the real number M with M > IITII. In particular, it follows from the last inequality that 11TII = sup{IITxll : IIxll < I}. Finally, assume that T is bounded, i.e. T satisfies (5.62). Let IITxll M =  IIxll · If we choose IIxll = 1, then from the definition we must have IITII < M. Further, if we let x  0, then by (5.62) IITxll < IITII IIxll - which gives M < IITII. Therefore, M = IITII. Hence, there is a number of alternate expressions for IITII = IITIIB(x,y) in the classical setting and in conclusion, we have 5.63. Theorem. If T : X  Y is a bounded linear transformation between two normed spaces and X :j:. {OJ, then liT II - inf{M: IITxll < Mllxll for all x E X} - sup IITxl1 IIx 11=1 - sup IITxll IIxll1 II Tx II - : IIxll · In the degenerate case X = {OJ, two out of these expressions in Theorem 5.63 are meaningless. However, the last equality is meaningful which gives IITII = O. Unless otherwise stated specifically, we always assume that the bounded linear operators are equipped with the operator norm defined in Theorem 5.63. We shall write IITII = 00 if T E L(X, Y) is not bounded. 5.64. Example. Consider X = (C[O, 1], II · 11(0) and define a func- tional T q, on X by T,p : X  Ii, f(t) t-t 1 1 rjJ(t)f(t) dt, where cp is a given function in X such that 4>(t)  0 on [0, 1]. We call Tq, a multiplication operator. Obviously, Tq, is linear and IT,pfl < Mllflloo. where M = 1 1 IrjJ(t) I dt, i.e. IIT,p1l < M, 
258 Chapter 5: Linear Operators on Normed Spaces so that Tq, is a continuous linear functional for each 4> EX. On the other hand, for each positive integer n, we have M = {1 IcjJ(t) I dt = {1 ( 14>(t)1 + n l 4>(t)1 2 ) dt 10 10 1 + nl4>(t)1 {1 IcjJ(t) I d T: ( n4>(t) ) - 10 1 + nlcjJ(t) I t + ., 1 + nlcjJ(t) I (I dt n4>(t) < 10 ;- + IIT.,II 1 + nlcjJ(t) I 00 < .!. + IIT.,II. n Letting n  00, we obtain the reverse inequality M = 1 1 IcjJ(t) I dt < IIT.,II. Hence, we must have IIT.,II = 1 1 IcjJ(t) I dt which shows that the minimum of all the values of M satisfying (5.60) is attained and the minimum is given by 1 1 IcjJ(t)1 dt. In particular, choose 4>(t) = 1 for all t E [0,1]. Then it follows that IT<t>4>1 = 1 since 114>1100 = 1. Hence, in this special case, we obtain IITq,11 = 1. In the equivalent characterizations in Theorem 5.63, we cannot expect that the ex- istence of such a minimum for every linear operator between normed spaces. This reasoning indicates why infimum is used rather than a minimum in the first equality in Theorem 5.63. . From the second equality in Theorem 5.63, we note that IITII is the least upper bound of the range of the map T ( 5.65) x  IITxlI on the unit sphere. In this connection, recall that, in the finite dimensional space X, the unit sphere is compact. Therefore, according to Theorem 5.33 and the Weierstrass theorem, namely "the continuous image of a compact set is compact" , there exists Xl E X with IIXlllx = 1 and IITxllly = liT II , 
5.6. Bounded Linear Operators 259 Le. the maximum is attained. On the other hand, in an infinite dimensional space, the sphere is not compact and hence, the map in (5.65), in general, does not attain its maximum on this set. Our next theorem shows that any linear transformation T between two normed spaces is either uniformly continuous or else everywhere discontin- uous. It is important to know that the linearity of T forces the continuity of T at one point to imply the continuity of T everywhere in the domain of T. Further, we also see in the next theorem that the bounded opera- tors are precisely the continuous linear operators and hence the people who work in this context refer to "continuous" by the name "bounded" for lin- ear operators. Again, we remark that the interchangeable use of the words "continuous" and "bounded" is justified from the fact that T : X  Y is continuous (in the topological sense) iff it is bounded in the sense of the definition (5.60). 5.66. Theorem. Let T : X  Y be a linear transformation be- tween the two normed spaces X and Y. Then the following statements are equivalent: (i) T is continuous at 0 (or at any point of X) (ii) T is continuous on X (iii) T is uniformly continuous on X (iv) T is bounded (Equivalently, T(Bx[O; 1]) is a bounded subset of Y, where Bx[O; 1] = {x EX: IIxll < I}). Proof. Without loss of generality we may assume that T :j:. o. The implications "(Hi) ==> (ii) ==> (i)" are clear. Therefore, it suffices to prove the implications "(i) ==> (iv)" and "(iv) ==> (iii) " . Assume (i) holds. Therefore the continuity of T at 0 implies that there exists a 6 such that IITxll < 1 whenever x E X and IIxll < 6. Therefore, for x = 6x' with IIx'll < 1, the last implication becomes IIT(6x'll < 1 whenever x' E X with IIx'll < 1, or equivalently, IITx'll < 1/6, Le. IITII < 1/6 and (iv) holds. Assume (iv) holds, Le. IITII is finite. Let x, y E X and € > O. Set 6 = €/IITII so that for x, y E X we have IIT(x - y)1I = IITx - Tyll < IITllllx - yll < € whenever IIx - yll < 6, arid therefore, (iii) holds. . We note that the linearity of T in the hypothesis of Theorem 5.66 cannot be dropped, in general, to get the same conclusion as the example T : IR  
260 Chapter 5: Linear Operators on Normed Spaces JR, x I-t x 2 , clarifies (with the Euclidean norm on IR). Here T is continuous on IR but not uniformly continuous on III As a consequence of the continuity, the nulls pace NT of T : X  Y forms a closed subspace of X. 5.67. Corollary. If a linear operator T between two normed spaces X and Y is continuous, then the nullspace NT is a closed subspace of X. Moreover, for linear functionals, we have the following facts: (i) A linear functional I on a normed space X is continuous iff the nullspace N/ is a closed subspace of X. (ii) H a nonzero linear functional I on a Banach space X is discontinuous, then the nullspace N/ is dense in X. Proof. Let T : X -+ Y be continuous. By Proposition 2.59, the inverse image of any closed set is closed. Since {OJ is closed and since T-I({O}) = {x EX: Tx = O} = NT, it follows that NT is closed. Alternatively, assume that T : X  Y is continuous and x E NT is arbitrary. Then there exists a sequence {x n } in NT such that X n  x. Continuity of T gives TX n -+ Tx. But TX n = 0 for each n so that Tx = 0 which implies that x E NT so that NT is closed. (i) We need only to prove that if the nulls pace N/ is closed, then I is continuous. Suppose that N / is closed. There is nothing to prove if I is a zero functional. IT I is not a zero functional, then there is a point a E X such that I(a) = 1. Then, a must belong to the open set X\I-l({O}). Thus, there exists an r > 0 such that the ball B(a; r) is" contained in X \1- 1 ({O}). We show that If(x)1 < IIxll for l x EX. r Suppose on the contrary that there exists an 0  Xl E X such that If(Xl)1 > IIXlll . r Define y = -xl/I(XI). Then Ilyll = Ilxlll I I/(Xl)1 < r, so that a + y E B(a; r). Therefore, a + y  1-1 ({O}), Le. I(a + y)  O. However, because of the linearity of I, Le. Y E B(O; r), f(a + y) = f(a) + f(y) = 1 + f ( - fl) ) = 1 - 1 = O. 
5.6. Bounded Linear Operators 261 This contradiction shows that If(x)1 < Ilxli/r for all x E X. Consequently, f is continuous. (ii) We leave it as an exercise. _ Later from Corollary 5.139, we obtain the following fact. "Suppose that we are given a linear functional f defined on a normed space and that N f is not dense. Then, Corollary 5.139 guarantees the existence of a continuous linear functional on X" . 5.68. Example. Note that if T is a bounded linear operator, then the range space RT need not be closed. For example, let T : 1 00  1 00 be defined by z = {Zn}n>l I-t {zn/n }n>l. - - Then, it is easy to see that IITII < 1 and Tz = 0 implies that z = o. Therefore, T E B(IOO, 1(0) and it is injective. Further, for each kEN (k = {I,2,...,k,k,...} E 1 00 and T (k = {I, 1, . . . , 1, k / (k + 1), k / (k + 2), . . .} E RT. We observe that as k  00 T(k  {I,I,I,...,} E 1 00 . But, {I, 1, 1, . . .}  RT, because {I,I,I,...}=Tz ==> z={1,2,3,...}100. Therefore, RT is not closed. Alternatively, as an observation, we may note that for large k, there is no c > 0 such that lI(klloo < cIlT(klloo. Hence, by Theorem 5.109 below, we conclude that the range space RT is not closed. Of course, another example is also given in Example 5.110. . The set B(X, Y) is closed with respect to vector space operations. In- deed, if T and S are in B(X, Y) then, because of the continuity of the operations of addition and scalar multiplication in Y, aT + bS is also linear and continuous. Hence, aT + bS E B(X, Y). 5.69. Convergence of sequences of bounded linear operators. A sequence {Tn} in B(X, Y) from a normed space X into another normed space Y is said to converge uniformly in the norm of B(X, Y) if there exists aTE B(X, Y) such that IITn - TII  0 as n  00; Le. given € > 0 there exists an integer N > 0 such that sup IITnx - Txll < € for all n > N. IIx"l 
262 Chapter 5: Linear Operators on Normed Spaces The sequence {Tn} of B(X, Y) is said to converge strongly if there exists a T E B(X, Y) such that lim n -+ oo IITnx - Txll = 0 for all x E X. From the inequality IITn x - Txll < IITn - Tllllxll, it follows that uniform convergence implies strong convergence. However, the converse is not true. We have already seen that L(X, Y), the set of all linear transformations from X  Y, is itself a vector space when X and Y are vector spaces.. We have not ye proved that the operator norm IITII, T E B(X, Y), actually is a norm. It is time to remedy this oversight. We also address the completeness property of the space B ( X, Y). 5.70. Theorem. Let X, Y and Z be normed spaces over the same field IF. Then we have the following facts: (i) B(X, Y) is a normed space over 1F under the operator norm. (ii) IfY is complete, then B(X, Y) is complete. (iii) H Z is a normed space over IF and if T E B(X, Y) and S E B(Y, Z), then the composition SoT =: ST belongs to B(X, Z) and the operator norm is submultiplicative, i.e. liS 0 TII < IISIlIiTIi. (iv) If X is finite dimensional, then any linear mapping is bounded, i.e. L(X, Y) = B(X, Y). Proof. (i) From Section 1.7, it follows that B(X, Y) is a vector space. Now, we show that B(X, Y) is a normed space with respect to the operator norm. For T, S E B(X, Y), we have the following: (N1) Clearly IITII > 0, since it is defined as the supremum of a set of nonnegative real numbers. Note that IITII = sup IITxll = 0 <=> x#O IIxll <=> <=> <=> IITxll = 0 for all x =I- 0 II xII IITxll = 0 for all x  0 Tx = 0 for all x EX, since TO = 0, T =0. (N2) Next, IIATII = sUPllxll=l II (AT)xll = sUPllxll=l IAllITxll = IAIIiTIl for all scalar A E IF. In particular, T E B(X, Y) => AT E B(X, Y) for A EOIF. 
5.6. Bounded Linear Operators 263 (N3) Since sUPllxll=l II (T + S)xll < sUPllxll=l IITxll + sUPllxll=l II Sx II , it fol- lows that ... liT + SII < IITII + IISII. In particular, T, S E B(X, Y) => T + S E B(X, Y). Therefore, B(X, Y) is a normed space. (ii) Suppose that {Tn} is a Cauchy sequence in B(X, Y), where Y is a Banach space. Then IITn - Tmll  0 as n,m  00. Since for each x E X, IITn x - T mxll = II (Tn - T m)xll < IITn - T m IIlIxll  0 as m, n  00, it follows that {Tnx} is a Cauchy sequence in Y for each x EX. Thus, by the completeness of Y, it has a limit in Y which we denote it by Tx:= lim Tnx. n-+oo Clearly, the operator T : X  Y is linear. In fact, for x, x' E X and A E IF, T(AX + x') lim Tn(AX + x') n-+oo - lim [ATn(X) + Tn(x')] n-+oo A lim Tn(x) + lim Tn(x') n-+oo n-+oo - ATx + Tx' so that T is a linear operator. Is it bounded? By virtue of the estimate IIITnll-IITmlll < IITn - Tmll, the real sequence {IITnll} is Cauchy, hence bounded; Le. there exists a constant K with IITnl1 < K for all n > 1 and so IITnxll < Kllxll for all x EX, n > 1. Since for all x E X we have Tnx  Tx, passing to the limit in the last inequality, one has IITxll < Kllxll for all x EX, and so T E B(X, Y). It remains to check that Tn  T in the operator norm of B(X, Y). Now, IITn - TII - sup IITnx - Txll, by definition, IIxlll - sup lim IITnx - T mxll IIxlll m-+oo < sup lim IITn - Tmllllxll IIxlll m-+oo < lim IITn - Tmll. m-+oo 
264 Chapter 5: Linear Operators on Normed Spaces Since {Tn} is Cauchy, it follows that IITn - TII  0 as n  00, so that Tn  T in the operator norm of B(X, Y). (iii) Clearly, ST = SoT is linear. Let x EX. Then, we have II (ST) (x ) II = IIS(Tx) II < IISlIlITxll < IISlllITllllxll which shows that if T E B(X, Y) and S E B(Y, Z), then ST E B(X, Z) and liS 0 TII < IISIlIiTIi. Thus, the operator norm is submultiplicative: IISTIIB(X,Z) < IISIIB(Y,z) IITIIB(X,y). Note that equality does not hold in general. Indeed, if we let x = y = Z = JR2, T(x,y) = (x,O), and S(x,y) = (O,y), then SoT = 0 and IISII = IITJI = 1. (iv) follows from Theorem 5.16. . The space B(X, 1F) has a unique place in functional analysis and is called the conjugate space (or dual space or norm dua of X and is usually denoted by a special notation X. instead of B (X, 1F). IT we need to specify the nature of IF, then we speak of real linear functionals (if IF = JR) and complex duals (if IF = C). By Theorem 5.70(i), B(X, Y) is a normed space and, in particular, X. is a normed space. It is, therefore, natur to ask when X. is complete. A rather surprising answer follows from the fact that IF (C or JR) is complete. 5.71. Corollary. The dual space X. is a Banach space (whether or not X is). This follows from Theorem 5.70(ii). 5.72. The algebra of continuous operators. A vector space V over IF is called an algebra iff there exists a map V x V  V, (x, y) I-t xy, which satisfies the following conditions for all x, y, z E V and A, JJ E IF: (i) x(yz) = (xy)z [Associative with respect to multiplication] (ii) A(J.tY) = (AJ.t)X (iii) x(y + z) = xy + xz and (x + y)z = xz + yz. [Multiplication is distributive with respect to addition] 
5.6. Bounded Linear Operators 265 An algebra V is said to be commutative iff xy = yx for all x, y E V. An algebra is said to have an identity element if there exists an e E V (called identity/unit element) such that ex = xe = x for all x E V. For example, since "T, S E L(X) implies that T 0 S E L(X)", L(X) is in a natural way an algebra over the field IF with unit element Ix, the identity operator. Moreover, II Ix II = 1. In many situations it makes sense to multiply elements of a normed space together. Not all algebras posses identity elements, as Remark 5.74 below shows. A normed algebra X is a normed space (X, 11.11) which is also an algebra (over the same field) together with a submultiplicativity of the norm: Ilxyll < IIxlillyll for all x, y EX. A complete normed algebra is called a Banach algebra. If a Banach algebra B has an unit element, namely the identity element of B written as e (or simply by 1 or I), then we also require that Ilell = 1. For example, the set C[O,I] of all continuous functions on [0,1] with the supnorm II · 1100 forms a Banach algebra with the multiplication given by (fg)(x) = f(x)g(x). A constant function with value 1 is the multiplicative unit element. A Banach algebra which is commutative is said to be commutative Banach algebra . Theorem 5.70(iii), in particular, shows that B(X) := B(X,X) with the operator norm is a normed algebra with unity (namely Ix = x), i.e. it has an additional "multiplication operation" which makes it a noncommutative algebra (see the example below). For example, with X = }Rn and taking a basis for }Rn , we may identify B (}Rn) with the space of n x n real matrices which is known to be a noncommutative Banach algebra for n > 2. Further, if T E B(X) then we define the power Tn of T by Tl = T and Tn+l = To Tn := T(Tn) for each n E N. 5.73. Example. Let X = (C[O, 1], II . 11(0)' Consider T, S E B(X) defined by (Tf)(y) = y 1 1 xf(x) dx and (Sf)(y) = yf(y), respectively. Then T(Sf)(y) = T(yf(y)) = Y 1 1 x 2 f(x) dx and S(Tf)(y) = yTf(y) = y211 xf(x) dx 
266 Chapter 5: Linear Operators on Normed Spaces which give TS :j:. ST, Le. B(X) is noncommutative algebra. . 5.74. Remark. In many parts of mathematics, an 'algebra' is un- derstood to have an identity element in it. However, this is not the case in functional analysis as we see below: (i) The space C with usual norm IIzll2 = Izl is the simplest Banach algebra which has unit element in it. (ii) The spaces of continuou.s functions which vanishes at infinity such as Co(lR) with pointwise multiplication and group algebras such as £1 (JR) with convolution as multiplication, are standard examples of an algebra. On can show that £1 (IR) does not have multiplication identi ty. (Hi) The space Loo, Co, C, 1 00 and Hoo (the bounded analytic functions on the unit disc  with supremum norm) are example of algebras (Here the multiplication is defined in the pointwise manner). (iv) Let X = en. Then B(X) can be identified with the algebra Mnxn (C) of all n x n matrices with entries from C. Under this identification, the operator multiplication corresponds to matrix multiplication. . Suppose that X is a Banach space. Then, by Theorem 5. 70(ii), it follows that the norm T I-t IITII actually makes B(X) a Banach algebra. It is usual to refer to B(X) as the algebra of bounded operators in X. Another property of B(X) is that it has a unique identity junit element I such that TI = IT = T for all T E B(X). In fact, IIIIIB(x) = sup IIIxll = sup IIxll = 1. IIx 11=1 IIx 11=1 5.75. Definition. (Closed operaor) Let X, Y be normed spaces and D T e x, a linear subspace. Then the operator T : D T  Y is closed if X n E DT for all n and X n  x E X, and TX n  Y E Y, then following hold x E DT and y = Tx. A motivation of the word "closed" in describing such operators follow from Proposition 5.114 where we proved that an operator T between two normed spaces is closed iff its graph is closed. 5.76. Examples of bounded linear operators. Boundedness of a linear transformation is often easy to check as we see in the following ex- amples: 
5.6. Bounded Linear Operators 267 (1) Let X = C[O, 1] with the supnorm and T : X  X, I(x) I-t xk I(x), where k is a fixed positive integer. Then T is linear and liT 11100 < 11/1100 for each I EX, which shows that IITII < 1. Thus, T is bounded. In fact, by cho.osing I(x) = x, we can obtain that IITII = 1 (since IIxll oo = 1 and liT 11100 = Ilx k + 1 1100 = 1 for x E [0, 1]). Does the same conclusion hold if X = Cc[O, 1] with I(x) I-t xk I(x)? Using the LP-norm with p > 1, we have IITfll = 1 1 !X k1P If(x)IP dx < IIfll so that IITlllp < II/lIp for all I E (C[O, 1], 1I.lIp)' What is IITII? (2) Let X = (C1[0, 1], II . 11(0)' the linear subspace of C[O, 1] consisting of all real valued functions on [0,1] that have continuous derivatives with the supnorm, and let Y = C[O, 1] with supnorm. Note that Y is a Banach space whereas X is an incomplete normed space. Indeed, consider the sequence of functions {/n(x)} in C 1 [0, 1] defined by In(x) = V (2x - 1)2 + (l/n 2 ). Then {/n(x)} is Cauchy and converges pointwise to the function I(x) = 12x - 11 which is in C[O,I]. On the other hand, II/n - 11100 = sup ( V (2x - 1)2 + (l/n2) - 1 2 x - 11) =.!:.  0 xe[O,1] n which shows that the convergence is uniform. But I is not dif- ferentiable at t = 1/2 and therefore, I cannot be in X. Thus, (C1[0, 1], II · 11(0) is not closed in (C[O, 1], II · 11(0) and by Proposi- tion 2.109(ii), (C1 [0, 1], II · 11(0) is not complete. Define T : X  Y, I I-t I'. Then T is clearly linear but not continuous. In fact, if In(x) = cos(n1rx) then I(x) = -n1rsin(n1rx) so that for each n > 1, we have II/nlloo = sup I/n(x)1 = 1, liT In 1100 = sup 1/(x)1 = n1r xe[O,1] xe[O,1] and liT In 1100 II/nlloo = n1r. For a similar example, we can consider the polynomial functions on [0, 1] with the supnorm: for example, let gn(x) = x n . Then, for each n E N, we have IIYniloo = 1 and IITgnlloo = sup {nxn-1} = n, xe[O,1] 
268 Chapter 5: Linear Operators on Normed Spaces so that IITgnlloo = n. IIgnlloo Since Il/nlloo = IIgnlloo = 1, both liT Inlloo and IITgnlloo increase in- definitely for n  00, there is no constant M such that liT 11100 < Mll/lioo for all I EX. Finally, for each I E X, consider the sequence {In (x)} defined by e- nx In (x) - I(x) = n so that I(x) - I'(x) = _e- nx . Then with respect to the supnorm In  I uniformly in X but I(x) + I'(x) in supnorm. Indeed, I(O) + 1'(0) - 1 so that T is not continuous. Therefore, in all the three illustrating examples, T is unbounded and hence not continuous. This operator T used in this example is often referred as differential operator. Also, note that T is closed because the dimension of the nullspace of T is one. Indeed, if In  I and Tin = I  9 then, since the derived sequence is uniformly convergent, I' exists and I  I'. Therefore, we have I' = 9 and hence I E X with T I = g, which shows that T is closed. But T is neither one-to-one (since the dim NT = 1) nor continuous. From this example we see that unbounded linear operators do occur in the applications. However, for each A E C, the operator T : X  Y, I(x)  e-'xx J: I(t)e'xt dt, is continuous. (3) Define a shift operator T : 1 2  1 2 , Z = {Zl, Z2, . . . , Zk, . . .} ...-t w = {O, Zl, Z2, · · · , ZIe, · · .}. Clearly, T is an isometry, since IITzII2 = IIzI12 for all Z E 1 2 . Therefore, T is a continuous operator with IITII = 1. Suppose, we consider a simple backward shift operator S : 1 2  1 2 , Z = {z 1 , Z2, . . . Z k, . . .} ...-t w = {Z2, Z 3, · · · , Z Ie, · · · } · Then, S is clearly not an isometry. Observe that 00 00 IISzlI = E I Z kl 2 < E I Z kl 2 = IIzlI k=2 k=l 
5.6. Bounded Linear Operators 269 and equality holds in this inequality for Z = {O, Z2, Z3, . . .}. Thus, the operator S is continuous with IISII = 1. For Z = {Zl, Z2,..., Zk,.. .}, we have STz = S({0,Zl,Z2,'" ,Zk,.. .}) = {Zl,Z2,... ,Zk,"'} = Z and T S Z = T( {Z2, Z3, . . . , Zk, . . .}) = {O, Z2, Z3 . . . , Zk, . . .}  z. Thus, ST = I and TS :j:. I, Le. neither T nor S is. invertible. On the other hand, T has a left inverse (and is one-to-one), while S has a right inverse (and is onto). (4) The linear operator T : 1 00  1 00 , {zn}  {zn/nQ} (where a > 1 is fixed), is continuous. (5) Let PF denote the set of all polynomials with coefficients in the field IF (IR or C). Define T : PF  PF by p(z) I-t np(z) with respect to the supnorm, where n is the degree of the polynomial. Then T is neither linear nor bounded: IITplioo = nllplloo, where Ilplloo = sup Ip(t)l. tE[O,l] (6) Define T : IR n  ]Rn, X = (Xl, X2,. . . , X n ) I-t (X2, X3, . . . , x n , 0), with the Euclidean norm. Then for each 0 :j:. x E IRn we have IITxlI = IIxlI - I X ll2 < 1 Le. IITxII2 < IIx1I2' Ilxll Ilxll -, and therefore, IITII < 1. Since IIT(O,X2,X3,...,x n )1I2 = 1, 11(0, X2, X3,..., X n )1I2 we get IITII > 1. Hence, we must have IITII = 1. . 5.77. Remark. Let X and Y be normed space and T E L(X, Y). Then, with respect to the metric induced by the norm, we have T is an isometry {:::::} d(Tx,Ty) = d(x,y) for each x,y E X {:::::} II-T(x - y)1I = IITx - Tyll = IIx - yll {:::::} IITxl1 = IIxli for each x EX. This observation shows that if X :j:. {OJ and if T is an isometry, then IITII = 1. However, the converse is not true; Le. IITII = 1 does not necessarily imply 
270 Chapter 5: Linear Operators on Normed Spaces that T is an isometry. Indeed, for I E X = (C[O, 1], 11.11(0) and T : X  }R defined by T(f) == 1 1 f(t) dt, we have IITII = 1 but for I(t) = t, IT(/)I = 1/2 :F 11/1100' This example shows that IITII = 1 does not guarantee that T is an isometry. . 5.78. Examples of bounded linear functionals. We gather to- gether a set of examples concerning the bounded linear functionals on normed spaces. (1) Let X = C[O,l] with the supnorm and Y = }R with the Euclidean norm. Define T : X  Y, I I-t 1(0). Then T is linear. Moreover, since ITII = 1/(0)1 < 11/1100 for each I E X, T is continuous. (2) Consider X = C[O,l] with L 2 -norm and Y = }R with the Euclidean norm. Then T : X  Y, I I-t 1(0), is linear and discontinuous. Consider In(t) = (1- t)n from C[O, 1] so that Tin = In(O) = 1 for all 1fl, E N and Tin  1 as n  00. Then [ 1 ] 1/2 1 IIfnll2 = 1 (1 - t)2n dt - (2n + 1)1/2  0 as n  00 and therefore, In  I(t) = 0 as n  00. But TI = 1(0) = O:F 1 and hence, the operator T in this case is neither bounded nor closed. Similarly, for the sequence 9n(t) = n1/8e-nt2, we have IIgnll2 = [1 1 (n1/8e-nt2)2 dt] 1/2 = n1/8 [1 1 e-2nt2 dt] 1/2  0 as n  00. But T maps 9n into 9n(0) = n 1 / 8 which is a divergent sequence of numbers. (3) The operator T : }R2  }R2 , i.e. T : C  C, defined by anyone of the following way (a) (x, y) I-t (-y, -x), Le. z t-+ -i z , (b) (x,y) t-+ (x - y,x + y), Le. z I-t (1 + i)z, (c) (x, y) I-t (x + y, 0), Le. z I-t [(1 - i)z + (1 + i) z] /2, is continuous. We note that the possible difference between the two operators, T : }R2  }R2 and T : C  C, is that the former is real linear while the later is complex linear. (4) For each j = 1,2,..., n, consider t4e projection mapping defined by 1rj : }Rn  IR, X = (Xl, X2, . . . , x n ) I-t Xj, with the Euclidean norm. Then 1r j is a linear functional on }Rn such that l1rj(x)1 = IXjl < IIxII2 
5.6. Bounded Linear Operators 271 and therefore, for each j, 1rj is continuous at 0 E }Rn and hence everywhere in }Rn. Indeed, one could directly see that l1rj(x) -1rj{Y)1 = IXj - Yjl < IIx - y1I2, Y = (Y1,Y2,... ,Yn) E }Rn, and the continuity of the projection mapping 1rj follows. (5) Consider a linear operator T : £1[0, 1]  IR, I t--+ f01 tl(t) dt. Then IT II = 1 1 tl(t) dt < 1 1 Itl(t)1 dt < 111111 so that IT II < 11/111 for all I E £1 [0,1], and therefore, IITII < 1. Now we show that IITII = 1. For this, we define a sequence {In} of functions in £1 [0, 1]: In(t) = {  for 0 < t < 1 - 1/ n for 1 - 1/ n < t < 1. Then IITII = 1 follows from the observations IT In I = {1 tln(t)dt = n {1 tdt = 1- 2 1  1 as n  00, 10 11-1/n n and IIInll1 = (1 In(t) dt = n (1 dt = 1. 10 1 1 - 1 /n (6) Let X denote the space of all continuous functions I on any closed interval I C IR such that I vanishes outside a finite interval [a, b] (depending on I) with the supnorm 11/1100 = sup I/(t)1 = sup I/(t)1 < 00. tER tE[atb] These functions can be integrated. Define T : X  }R by I(t)  i: I(t) dt = lb I(t) dt. Then T is unbounded. Indeed, consider a sequence of functions {In} in X (see Figure 5.2): In(t) = o l+t 1 n+l-t for t E (- 00, -1) U (n + 1, 00 ) for t E [-1, 0] for t E [0, n] for t E [n, n + 1]. 
272 Chapter 5: Linear Operators on Normed Spaces -1 0 n n+I Figure 5.2: The graph of fn(t) Clearly, In(t) E X, II/nlloo = SUPtER I/n(t)1 = 1 and 1 1 0 i n I n+l Tln= I(t)dt= (I+t)dt+ dt+ (n+I-t)dt=n+l. R -IOn Therefore, T is unbounded in the unit. ball of X. It is important to note that if the interval [a, b] is fixed, then the operator T becomes bounded. (7) Let a = (ai, a2, . . . , an) E 1 2 (n) be a fixed nonzero element. Define n I: l2(n)  IF, Z = (Zl,Z2,...,Zn) t--+ LZk a k. k=l Then I is clearly linear. By the Schwarz inequality, we have n n I/(z)1 = LZk a k < L IZk a kl < IIzII211all2 k=l k=l and therefore, I is a bounded linear functional. Indeed, for Z  0, we find that I/(z)1 . IIzll2 < lIall2. I.e. 11/11 < lIall2. The choice Z = a gives I/(a)1 = EZ=11ak12 = lIall so that 11/11 = Ila112' Note that if a is a zero vector in 12(n), then I becomes a zero functional and therefore, 11/11 = lIall2 is trivial. Similarly, for a fixed nonzero vector a = {an}nl E 1 2 , I : 1 2  IF define by I(z) = E  1 Zk a k for Z = {zn}, is a linear functional with 11/11 = lIall2. (8) If I : 11  IF is defined by Z = {Zk}kl t--+ E  1 Zk, then we have 00 I/(z)1 < L IZkl = IIzlh k=l so that 11/11 = 1. Hence, I is a bounded linear functional on 11. 
5.6. Bounded Linear Operators 273 (9) Let X = Coo c 1 00 , the vector space of all finitely supported sequences {Zn}n1. Define T : X  ]F by the formula 00 T( {zn}) = L Zn. n=1 For each n E N, let n Zn = Lej. j=1 Then, we have that Zn E X for each n E N. Moreover, for each n EN, IIZnlloo = 1 Le. Zn E Bx[O; 1] and T(Zn) = n. It follows that T(Bx[O; 1]) is not bounded and therefore, T is not a bounded operator. (10) Suppose that 9 E (CF[a, b], II . 112) is fixed. Define b T : (CF[a, b], 11.112)  IF, I t-t 1 I(t) g(t) dt. Then, as in the previous example, T is linear and the inequality IT II = l b I(t) g(t) dt < l b I/(t) g(t) 1 dt < 11/1I211g112 (see Holder inequality) shows that T is bounded with IITII < IIg112. Further, Tg = IlglI and therefore, we obtain that IITII = IIg112. . 5.79. Example. If X = C[O,I] is equipped with the supnorm and if k(x, y) is continuous on the unit square [0,1] x [0,1], then the operator T defined by (5.80) T: X  X, I t-t TI, i.e. (Tf)(x) = 1 1 k(x,y)/(y)dy is bounded (The function k(x, y) is usually referred to as the kernel of the operator T). In particular for k(x, y) = x, the functional T : X  X, (Tf)(x) = x 1 1 I(y) dy, is bounded with IITII = 1. Here, the linearity ofT is easy to verify. Further, if M = sUPOx,y1 Ik(x, y)1 then for each f E C[O, 1], we find that IITfiloo = sup IT(f(x))1 0x1 - sup (1 k(x, y)/(y) dy 0x1 10 < M sup If(y)1 = Mllflloo. 0y1 
274 Chapter 5: Linear Operators on Normed Spaces Therefore, T is a bounded (continuous) linear operator. In this example, continuity of T may be verified directly. Indeed, if In  I in C[O,I] then (since the convergence in C[O, 1] is uniform, the limit operation can be taken inside the integral), lim (Tfn)(x) = (l k(x,y){ lim fn(y)}dy = {l k(x,y)f(y)dy = (Tf)(x) n-+oo J 0 n-+oo J 0 and the continuity of T follows. Also, we note that the integral expression in (5.80) may be used to define various integral operators T : X  X with X = L 2 [0, 1], etc. . 5.81. Example. We give an example showing that a bounded set in (C[a, b], 11.11(0) is bounded in (C[a, b], 11.111) but not conversely (see also Exercise 5.159). Suppose that A C C[a, b] is bounded with respect to the supnorm. Then for each I E A, we have 11/111 < (b - a)lI/lIoo, and therefore, the first part is clear. For the converse part, we assume a = 0 and b = 1 for the sake of convenience, and consider { I - nt fn(t) = 0 if 0 < t < I/n , n > 1, if I/n < t < 1 and the subset A = {In E C[O, 1] : n > I}. Then II/nlloo = max I/n(t)1 = n tE[O,l] and 1 1 1 1/n nP-1 Il/nll = I/n(t)I P dt = n P (1 - nt)P dt = l ' o 0 p+ Therefore, A is unbounded with respect to the supnorm as well as with respect to the LP-norm for p > 1. But II/nlll = 1/2. . 5.7 Inverse Operators Recall that if I : X  Y is a map with I(X) C Y, then a necessary and sufficient condition for the existence of an inverse map, 1-1 : I(X)  X, is that I is injective. Thus, if T : X  Y is linear operator then the condition that the nullspace NT = {OJ is necessary and sufficient for the existence of the inverse map T-1 : RT = T(X)  X. Clearly, T-1 is linear. 
5. 7. Inverse Operators 275 Indeed, let Y1, Y2 E RT and, Xi = T-1Yi (i = 1,2) be their preimages. Since T is linear, we have T(>"X'l + J.t X 2) = >"TX1 + J.tTX2 = >"Y1 + J.tY2 for >.., J.t E F, which implies that there exists an element >"X1 + J.tX2 such that T-1(>"Y1 + J.tY2) = >"X1 + J.tX2 = >..T-1Y1 + J.tT- 1 Y2' Thus, T-1 is linear. Now, we introduce the formal definition: Let T : X  Y be a given linear operator. We say that T has an (algebraic) inverse or is (algebraically) invertible if there exists a linear operator S : RT  X defined on RT and assuming values in the domain of T with the property that ST = Ix, Le. STx = X for every x E Domain (T) and TS = IRT' Le. TSy = Y for every Y E RT. Here Ix and IRT are the identityjselfmappings on X and RT, respectively. The operator S is said to be the inverse of T and is denoted by T-1. The operator T and T-1 are termed mutual inverses and from this definition it follows that (T-1)-1 =T. In the space L(X), the operator T and T-1 E L(X) are characterized by T-1T = I = TT- 1 . Further, if T, S E L(X) then we can easily see that (TS)-l = S-lT- 1 . 5.82. Example. Consider the differential equation given by (5.83) au" (t) + bu' (t) + cu(t) = v(t) with some initial conditions u(O) = 0 = u'(O). Using standard results from ordinary differential equations, it is easy to solve this differential equation. Indeed, the auxiliary equation is a>..2 + b>" + c = O. If the discriminant b 2 - 4ac is positive, then the above quadratic equation has two distinct real roots >"1, >"2 .( say). Thus the two linear independent solutions of homogeneous differential equation Su = au" + bu' + cu = 0 
276 Chapter 5: Linear Operators on Normed Spaces is {e-Xl t , e-X 2 t}. U sing standard method, one can show that the solution of the nonhomogeneous differential equation Su = v is i t e-Xl t - e-X2 t (5.84) u(t) = 0 k(t - s)v(s) ds, k(t) = a(Al _ A2) · This problem may be translated to the framework of linear operators. H v E X, X = C([O,oo)), then so does u and in this case, (5.84) can be written in the form Tv = u, where T : X  X, (Tv)(t) = l t k(t - s)v(s) ds. This operator is clearly linear and continuous. We note that (5.83) is equiv- alent to Su = v, where S is again a linear operator. Moreover, S is not defined on all of X, but is only defined on the dense linear subspace of twice differentiable functions. Furthermore, S is not continuous. However, T = S-1 is bounded when restricted to any compact interval [0, e] for any e > 0, since IITvll = sup i t k(t - s)v(s) ds tE [O,c] 0 < sup i t Ik(t - s)llv(s)1 ds tE[O,c] 0 < Mllvll oo alA1 - A21 for some constant M > O. We know that a necessary and sufficient condition for S to be invertible is that S is bijective. So, in this example, S : Y  X is (algebraically) invertible if Y = C2[0, c] and X = C[O, e]. Thus, the solution for the operator equation Su = v exists. . Finally, since the inverse of the linear transformation is again a linear transformation, Theorem 5.66 is particularly useful to have the following characterization of linear homeomorphism. 5.85. Theorem. Let T : X  Y be a linear operator between two normed spaces and T(X) = Y. Then T is a homeomorphism iff there exist c > 0 and C > 0 such that (5.86) ellxll < IITxll < Cllxll for each x E X. Proof. Suppose that T is a homeomorphism, Le. T is both bijective and bicontinuous. By Theorem 5.66, the continuity of T implies that there exists a constant C > 0 such that IITxll < Cllxll for each x EX. 
5.8. Completion of Normed Spaces 277 Similarly, continuity of T-l implies that there exists a c > 0 such that IIT-lyll < c- 1 l1yll for each y E T(X) = Y. Since for each y E Y, y = Tx for some x E X so that IIT- 1 (Tx) II < c- 1 I1Txll, Le. cllxll < IITxll, for each x EX. Conversely, suppose that the inequality (5.86) holds. Then, by Theorem 5.66, T is continuous. If we put Tx = 0 in cllxll < II Tx II , then it follows that NT = {OJ so that T is one-to-one. But, by hypothesis, T is onto. Therefore, T is bijective. In particular, T- 1 exists on Y = T(X) and is linear. To show that T-l is continuous on T(X), we let y E Y = T(X) and write x = T-ly. Then Tx = y so that cllxll < IITxl1 => IIT-lyll < c- 1 I1yll, for each y E T(X). By Theorem 5.66, T- 1 is continuous on T(X) = Y. . It is interesting to draw the following criterion for the inverse operator to be bounded. 5.87. Corollary. Let T : X  Y be a linear operator between two normed spaces X, Y, and let T be onto. Then T-l exists and is continuous iff there exists c > 0 such that cllxll < IITxll for each x EX. For example, let X = {z = {Zk}kl} equipped with the norm IIzll = IIzll2 + IIzlloo (lIzll < 00), where II · 112 and II · 1100 are 12-norm and loo-norm, respectively. Then it is trivial to see that IIzll2 < IIzll < 211z112 and therefore, X is homeomorphic to 1 2 . 5.8 Completion of Normed Spaces The completion of a metric space has been discussed in Section 2.8. In this section we address the following question: Given an incomplete normed space X, is it possible to enlarge X so that the new space is Banach? As with the metric spaces if a normed space X is not complete, then it is possible to expand X by adding new elements and suitably redefining the norm to cope with the new elements so that the resulting space is complete. 
278 Chapter 5: Linear Operators on Normed Spaces Again, the completion of the set of rational numbers Q serves as a model for the general completion process. What do we mean by a completion of a normed space X? This is simply a Banach space X. which contains a dense subspace which is isometric to X. First we prove the following lemma which is essential in the proof of our next theorem. 5.88. Lemma. Let Xl and X 2 be two normed spaces and X2 be a Banach space. Suppose that Y is a dense subspace of Xl. 1fT E B(Y, X 2 ), then there exists a unique extension ofT to S E B(X 1 , X2), with Sly = T, i.e. T(y) = S(y) for all y E Y. Proof. Let x E Xl' Since Y is dense in Xl, for each x E Xl, there exists a sequence {Yn} in Y converging to x. The boundedness of T shows that IITYn - TYmll < IITllllYn - Ymll for each m, n E N. Now, {Yn} is a Cauchy sequence in Y (since every convergent sequence is Cauchy) and therefore, the last inequality implies that {TYn} is a Cauchy sequence in X 2 . Since X 2 is complete, the sequence {TYn} converges to some element x' E X 2: lim TYn = x'. n-+oo This limit depends only on x and not on the sequence {Yn}. To prove this claim, let us assume that both {Yn} and {y} converge to x. Then, {Ty} must converge to some element z' E X2. We need to show that x' = z'. Now we have IITYn - Ty1I < IITllllYn - y1I < IITII {llYn - xII + lIy - xII}. and the desired claim follows as n  00. Alternatively, this can be seen in the following way. Combine the two sequences as follows: " { Yn Yk = Y if k = 2n - 1, n > 1, if k = 2n, n > 1, so that {Y}kl = {Y1, Y, Y2, y,.. .}. The mixed sequence {y} converges to x and so, the sequence {Ty} is convergent in X 2 . But, since {Ty} has one subsequence converging to x' and another subsequence converging to z', it follows that x' = z' 
5.8. Completion of Normed Spaces 279 as claimed. We then define an operator S : Xl -+ X 2 as follows: Sx = lim TYn (:= x'). n-.+oo It is easy to check that S is a linear extension. Indeed, if x E Y then we can take Yn = x for all n so that Sx = Tx for all x E Y. Moreover, the continuity of the norm II. II and the boundedness of T show that IISxll = lim IITYnl1 < IITII lim IIYnll = IITllllxl1 n-+oo n-.+oo and therefore, IISII < IITII. Also, 1f311 = sup IISxll > sup IITxll = IITII XEX1,X:;CO Ilxll xEY,x,eO Ilxll and therefore, IISII = IITII. The final step of the proof is to show that S is unique. If S' were another extension of T, then for each x E Xl there exists {Yn} such that Yn -+ x and, by the continuity of S', S'x = lim S'Yn = lim TYn = Sx n-+oo n-+oo which proves the uniqueness. . 5.89. Theorem. For every normed space X = (X, 11,11), there exists a Banach space X. = (X., II . II.) such that (i) X C X. in the sense that X can be identified via an isomorphism with a subset of X. . (ii) Ilxli = Ilxll. for all x E X (iii) X. contains a dense subspace that is isometric with X ( The space X. is called a completion of the normed space X ). The com- pletion is unique up to an isometry. Proof. The proof of this theorem is divided into several steps. Let X = (X, II · IIx) be a given normed space. Let S denote the set of all Cauchy sequences in X. Two Cauchy sequences {x n } and {Yn} in S are said to be equivalent, written {x n } f"oJ {Yn}, iff IIxn - Ynllx -+ 0 as n -+ 00. For simplicity, we shall write II · II rather than II. Ilx. 
280 Chapter 5: Linear Operators on Normed Spaces Step 1: The relation "-J is an equivalence relation on S. As in the proof of Theorem 2.112, it is easy to check that the relation "-J is reflexive, symmetric and transitive. These properties are easy to derive as every normed space is a metric space with respect to the metric d(x, y) = Ilx - yll. Therefore, the set of all Cauchy sequences of X is decomposed into equivalence classes, where two Cauchy sequences belong to the same equivalence class x* iff they are equivalent. Let X. = S / "-J be the collection of all these equivalence classes x*. If {xn} E x*, we will say that the Cauchy sequence {xn} is a representa- tive/ element of x*. Our aim is to show that X* is just the required Banach space, and for this we need to complete the following tasks: . define a vector addition and scalar multiplication on X* . show that X. is a vector space . define a norm on X* and make X* a normed space . show that X is isomorphic to a subspace Xo of X* . show that X. is complete . show that Xo is dense in X*. Step 2: Vector space structure on X*. Consider two elements/classes x*,y* E X. and choose two representatives {xn} and {Yn} of x* and y*, respectively. Our aim is to make X* into a vector space. Now, for all n,m > 1, we have lI(xn + Yn) - (xm + Ym)11 < IIxn - xmll + llYn - Ymll. Since {xn} and {Yn} are Cauchy sequences, the sequence {x n +Yn} is also a Cauchy sequence and therefore, it belongs to some equivalence class, which may be symbolically denoted by x. + y*. Thus, we define the sum of two elements x* and y* of X* with representatives {xn} and {Yn} respectively, to be the class of all Cauchy sequences equivalent to {x n + Yn}' Moreover, for each a E IF, we have lIaxn - aX m II = lalllxn - X m II so that {axn} is also a Cauchy sequence. Therefore, {axn} belongs to some equivalence class which may be symbolically denoted by ax*. Thus, for each a E F, we define the scalar multiplication of an element x* E X* with the representative {xn} to be the class of all Cauchy sequences equivalent to {ax n }. We prove that the above definition is indeed well defined by showing that the operations of addition and scalar multiplication in X* do not depend on the choice of the representatives representing the classes x* and y*. Indeed, suppose {X n }, {x} E x. and {Yn}, {y} E y*. 
5.8. Completion of Normed Spaces 281 Then, in view of the definition of the classes x* and y*, we have {Xn}  {x} and {Yn}  {y}, that is, Ilxn - x1I  0 and llYn - y11  0 as n  00. But then, by the triangle inequality, we have II(xn + Yn) - (x + y)11 < Ilxn - x1I + llYn - y1I  0 as n  00; thus {x n + Yn}  {x + Y} and therefore, {x + Y} E x* + y* so that the addition in X* is well defined. Let a E F. As in the case of vector adition, it is easy to see that (ax)* := ax* is independent of the choice of a representative from x*. Indeed, if {x n }, {x} E x* then {Xn}  {x}, Le. IIx n - x11  0 so that Ilaxn - ax11  0, Le. {axn}  {ax}. Thus, {ax} E (ax)* which shows that (ax)* does not depend on the choice of a sequence from x*. To say that X* is in fact a vector space, we need to verify all the axioms of the definition .of a vector space. For instance, if x*, y*, z* E X* and {x n }, {Yn}, {zn} are the representatives of x*, y*, z*, respectively, then (x* +y*) +z* is the equivalence class containing {(x n + Yn) + zn} while x* + (y* + z*) is the equivalence class containing {x n + (Yn + zn) }. Since X is a vector space, we have (x n + Yn) + Zn = X n + (Yn + zn) and therefore, it follows that (x* + y*) + z* = x* + (y* + z*), for each x*, y*, z* E X*. The remaining axioms may be verified similarly. Let us find a class that plays the role of 'additive identity'-zero 0* in X*. This element is deter- mined by the condition x* + 0* = x* and so, we conclude that the class 0* E X* is the equivalence class of sequences converging to 0 EX, and 0* contains the constant sequence {OJ := {O, 0, 0, . . .} as one representative of 0*. Step 3: Normed space structure on X*. Let x* E X* and {x n } E x*. Then Illxnll -lIxmlll < Ilx n - xmll  0 as n, m  00 
282 Chapter 5: Linear O.perators on Normed Spaces and thus, {llxnll} is a Cauchy sequence in the Banach space III This obser- vation shows that the limit lim n -+ oo IIxn II exists. Hence, we may introduce the function II · IIx. on X. as follows: ( 5.90) IIx. II. = lim IIxn II n-+oo where {xn} is a Cauchy sequence in the class x.. Again, for simplicity purpose, we shall write II · II. rather than II . IIx.. For the limit (5.90) to make sense, it is crucial to show that this limit is independent of the chosen representatives representing x.. This can be checked as follows: Suppose {Xn} and {x} are two representatives of the same equivalence class X.. Then, by the triangle inequality, IlIxnll-llx1I1 < IIxn - x1I  0 as n  00 which implies that the limit in (5.90) is indeed independent of the choice of a representative from x.. It is now straightforward to verify that the function II · II. satisfies all the axioms (Nl)-(N3) of Definition 3.2 on the space X.. (i) Obviously Ilx.1I > 0, since IIxnll > o. Assume IIx.1I = o. By (5.90), o = lim IIxnll = lim IIx n - 011 n-+oo n-+oo so that {x n } f"oJ {O}:= {O,O,O,...}. In particular, the class x. con- tains the constant sequence {O, 0, 0, . . .}, Le. x. = O. and the positiv- ity condition (Nl) holds. (H) The homogeneity condition, namely, lIax. II = lallix. II, follows from the fact that lIax n II = lalllxn II. (Hi) Finally, since IIx n + Ynll < IIxnll + llYn II, we have IIx. + y. II. lim IIxn + Ynll n-+oo < lim [llx n II + llYn II] n-+oo - lim IIx n II + lim IIYnti n-+oo n-+oo - IIx. II. + IIY. II. so that the triangle inequality (N3) follows. Consider a mapping T : X  Xo defined by T (x) = x. for x EX, where x. is the equivalence class which contains the stationary sequence { x }. For x :j:. Y, Tx :j:. Ty and this map is well defined. Indeed, if x E X then 
5.8. Completion of Normed Spaces 283 there is at most one equivalence class consisting the stationary sequence {x} and hence the map is well defined. Clearly, this map is onto since, for each x* E Xo, there exists a unique element x E X such that the stationary sequence { x } E x. with Tx = x*. Finally, if x, Y E X and { x }, { y } are the corresponding constant sequences, then IITx - Tyli. = IIx. - y.lI. = Ilx - yll showing that T is a distance preserving surjection from X into Xo. This proves the existence of an isometry between X and Xo C X.. Step 4: T(X) = Xo is a dense subspace of X.. Let x* E X* and € > 0 be given. We need to show that the ball Bx. (x.; €) contains at least one element of Xo other than x.. For this, we choose a representative {x n } of x*. Since {x n } is Cauchy in X, there exists an N E N such that Ilx n - X m II < €/2 whenever m, n > N. Since T : X  Xo is an isometry, with XN = u E X, we consider the element T(u) E Xo and note that the stationary sequence { u } E u*(= T(u)). This observation shows that Ilx* - T(u)ll. - lim Ilx n - xN11 n-.+oo < €/2, which means that T(XN) E Bx. (x*; f). As Xo is the range ofT, we conclude that x. is in the closure of Xo. Thus, Xo is dense in X.. Step 5: X. is a Banach space: Indeed the fact that the space X. is complete follows from Step 7 of Theorem 2.112. However, we provide a slightly different proof. Also, the following proof may also be used to prove Step 7 of Theorem 2.112 Let {x.(k)} k>l .be a Cauchy sequence in X.. For each positive integer k, let {xk)}n>l-be a Cauchy sequence in X belonging to the equivalence class that repesents the element x.(k) E X*. Since {xk) }nl is a Cauchy sequence in X, there exists an N k E N with N k > k and 1 Il x(k) - x(k) Ilx < - whenever m n > N k . n m - 2 k ' - Clearly, this is possible because any representative is a Cauchy sequence. Set Yk = x. We claim that {Yk} is a Cauchy sequence in X. Then for all n, IIYk - Yllix < IIYk - xk) IIx + Ilxk) - x) Ilx + IIx) - Yllix. Letting n  00, it follows that 1 1 IIYk - Yllix < 2 k + IIx.(k) - x.(l) II. + 2 ' 
284 Chapter 5: Linear Operators on Normed Spaces since, by (5.90), lim Ilxk) - x) Ilx = Ilx.(k) - x.(l) II.. n-.+oo Now, the assumption that {x.(k)} is Cauchy implies that {Yk} is indeed a Cauchy sequence in X and therefore, it defines an equivalence class y. E X. which is represented by the Cauchy sequence {Yk}. Furthermore, Ilx.(k) - y. II. - lim Ilx(k) - Ynllx n-.+oo n < lim {lIxk) - Ykllx + IIYk - Ynllx} n-+oo lim IIxk) - Yk Ilx + lim IIYk - Ynllx n-.+oo n-+oo < 2 1 k + Hrn IIYk - Ynllx n-+oo  0 as k  00, since {Yk} was already shown to be a Cauchy sequence in X. It follows from this that x.(k)  y. in X. as k  00, and therefore, each Cauchy sequence {x.(k)} in X. converges. We conclude that X. is complete. We have shown that for any normed space (X, 11.11) there exists a Banach space (X., II. II.) containing a subspace (X o , II .11.) such that . (X o , II . II.) is isometric to (X, II · II) . Xo is dense in X.. Now, we aim to show that there is only one such completion X. up to an isometry (in other words, any other such completion is isometric to X.). For the proof of the uniqueness, we use Lemma 5.88. Let (Xi, II · IIi) (i = 1,2) be two completions of X and T i : X  RTi be the corresponding isometries such that Ti(X) = RTi is dense 'in Xi. Clearly, Tl 0 T 2 - 1 is an isometry from RT2 onto RT 1 , and by Lemma 5.88 it extends to isometry from X; onto Xi. . 5.91. Example. It can be shown that the same space C[a, b] of all continuous functions on [a, b] with the supnorm is the completion of the following incomplete normed spaces: (i) The space of all piecewise linear functions 1 (i.e. Iht""t"'+l] is a linear function, where to = a < tl < t 2 < ... < t n = b) defined on [a, b] with the supnorm 11/1100 = sup I/(t)1 tE[a,b] (H) The space of all polynomials (Hi) The space of all infinitely differentiable functions on [a, b]. 
5.9. Quotient Spaces 285 Similarly, for p E [1, 00 ), LP [a, b] is the completion of the space of all [ b ] lip polynomials with p-norm II/lip = fa 1/(t)IP dt and also of the space of all piecewise linear functions with the same norm. We leave the proof of the above items as exercise problems. 5.9 Quotient Spaces Before we proceed to state the precise definition of a quotient space, let us start with a simple example: consider a 3-dimensional Euclidean space IR 3 = {x = (x I , X 2 , X 3) : x I , X2 , X3 E IR}. Clearly, Yo = {(Xl, X2, X3) E IR 3 : x3 = O} = span {el, e2} i a subspace of IR3 which is in fact the Xlx2-plane. Then all the planes parallel to Yo may be denoted by Yo:, where Yo: = {x = (Xl, X2 , X3) E IR 3 : X3 = a}, a E III Suppose that the planes are added and multiplied by a scalar in a natural way: Yo: + Y,a = Yo:+,a, AYo: = YAo: (A E IR). These operations give rise to a vector space whose elements are planes parallel to Yo, where Yo acts as the zero element. The vector space so obtained is called the quotient space ofIR3 with respect to Yo, and is denoted by IR3/Y O . We note that there is a one-to-one correspondence "a  Yo:" between IR and IR3/Y O preserving the linear structure: a + {3  Yo: + Y,a, Aa  YAo:. Now, if X is a vector space and Y a subspace of X then for X E X we may define the coset [x] relative to Y by [x] := {x' EX: x' - x E Y} = {x + z : Z E Y} = x + Y. Define x' "-J x iff x' - x E Y. Then '''-J' defines an equivalence relation on X. Indeed, (i) since 0 E Y, it is clear that x "-J x for x - x = 0 E Y (ii) if x' "-J x,. then x' - x E Y and, since Y is a subspace, x - x' -(x' - x) E Y, we have x "-J x' (iii) if x' "-J x and x "-J x", then x' - x E Y, x - x" E Y and, since Y is a subspace, x' - x" = (x' - x) + (x - x") E Y so that x' "-J x". 
286 Chapter 5: Linear Operators on Normed Spaces This proves the equivalence relation on X. Further, if x' f".J x, this means that x' E x + Y and any other vector x" is equivalent to x' under f".J must also belong to x + Y. Hence any two equivalence classes [x] = x + Y and [y] = y + Y is either identical or disjoint: Le. either [x] = [y] or [x] n [y] = 0. Let us denote the set of all cosets, or equivalence classes by X/Y: X/Y := {[x] : x E X} = {x + Y : x EX}. We read X/Y as 'X modulo Y' or, more simply, X mod Y and is called the quotient space or factor space of X with respect to Y. Let us now introduce the operations of addition and scalar multiplication on the elements of X/Y so as to make X/Y a vector space. First, we note that [x] = [y] whenever x - y E Y so that we can identify the elements of X/Yo For A E IF, [x], [y] E X/Y, the operations of addition and scalar multiplication can be introduced in X/Y in a natural way: [x] + [y] A [x] - [x+y], Le. - [Ax], Le. (x + Y) + (y + Y) = (x + y) + Y A(x + Y) = AX + Y. One can easily see from the linearity of the subspace Y that this definition of linear operations is well defined. First we note that x+Y=Y<==}xEY so that two cosets Xl + Y and X2 + Yare equal, as sets Xl + Y = X2 + Y, which is true iff Xl - X2 E Y, Le. Xl - X2 = Yl for some Yl E Y. Thus, with these operations, it is easy to verify that the space X/Y becomes a vector space over the field F. First, we consider [x], [x'], [z], [z'] E X / Y such that [x] = [x'] and [z] = [z'] which is equivalent to x - x' , z - z' E Y and so , , £ E Y x = x + Yl, Z = z + Y2 or some Yl, Y2 · We want to show that [x] + [z] = [x'] + [z']. To do this, we note that [x] + [z] = (x + z) + Y = x' + Yl + z' + Y2 + Y = x' + z' + Y because Yl + Y2 + Y = Y. Hence, the operation of addition is well defined. Similarly, we see that the scalar multiplication is well defined. The map q : X  X/Y, x I-t [x] := x + Y, 
5.9. Quotient Spaces 287 is called the quotient map of X into X/Yo This map is clearly linear and q(x) = q(x - y) for each y E Y. It is also called natural homomorphism or canonical homomorphism of X onto the quotient space X / Y. 5.92. Example. Let X = }R3 and Y = span {(I, 1,0)}, a closed subspace of }R3. Then X/Y is a two dimensional real vector space, (1,0,1) + Y and (0,0,1) + Y form one among the many pairs of elements that generate X/Yo . 5.93. Example. Let X = II and Y = {{Zk}kl E II : Zl = Z2 = . · · = Zn = OJ. Then, Y is a closed subspace of II and that X/Y is isomorphic to }Rn. . Notice that in the first two examples, Y is a closed subspace. Thus, our real interest on quotient spaces lies in the case where X is a normed space. In this case, the question remains to see is how one can define a norm on X/Y in a natural way. First, we note that such a norm should make the quotient map q continuous and hence, the norm on X/Y should satisfy a condition like IIq(x)llx/y := Il[x]lIx/y < Ilxlix. Since [x] = [x - y] for each y E Y, the norm condition on X/Y must also satisfy II [x] II < Ilx - yll for each y E Y. This idea leads us to define a norm on X/Y Il[x]1I = inf Ilx - yll = inf Ilx + yll = dist (x, Y), yEY yEY that is, Il[x]llx/y = inf 1I(llx. . <Eq-l([x]) It is easy to see that this definition is well defined and is a semi-norm. Note that Il[x]llx/y = O.<==} x E Y . This observation shows that the above semi-norm on X/Y becomes a norm precisely when Y is a closed subspace of X. Hereafter we shall refer to the normed space X/Y with the above norm as the quotient space. We now verify if the above definition satisfies the axioms of a normed space. I (1) 0 E Y gives Il[x]llx/y < Ilxllx. 
288 Chapter 5: Linear Operators on Normed Spaces (2) Clearly, lI[x]llx/y > o. Suppose that lI[x]llx/y = O. Then, in view of the definition, this means that there exists a sequence {Yn} E Y with llYn - xlix  0, Le. Yn  X as n  00. Since Y is closed, x E Y. Since x E Y, we get x+Y=Y. (3) For 0 # A E F, II[AX]lIx/y = inf IIAX - yllx = inf IIA(X - Y)lIx = IAlll[x]lIx/y. yEY yEY Clearly, II[AX]llx/y = IAlll[x]lIx/y for A = 0 as 0 E Y, and therefore, inf yEY Ilyllx = O. (4) Finally, given x,x' E X, and € > 0, choose Y,Y' E Y such that Ilx - yllx < Il[x]llx/y + € and IIx' - y'llx < Il[x']llx/y + €. Then we have Il[x] + [x']lIx/y - Il[x + x']lIx/y < lI(x+x') - (y+y')lIx, y+y' E Y, - II (x - y) + (x' - y')lIx < IIx - yllx + IIx' - y'lIx, by triangle inequality, < Il[x]llx/y + Il[x'Jllx/y + 2€. Since € > 0 is arbitrary, the triangle inequality (N3) Il[x] + [x']lIx/y < Il[x]lIx/y + Il[x']llx/y follows. (5) Since . IIq(x)lIx/y IIq(x)lIx/y := II[x]lIx/y < IIxllx, I.e. sup II II < 1, x#O x x the quotient map q is continuous and it maps the unit ball Bx (0; 1) into the unit ball Bx/y(O; 1). In fact, the quotient map q is onto so that q becomes an open map. . If T : X  Y is a linear map between two vector spaces X and Y, then X jker T is one of the most commonly occurring quotient space, so that the induced map T' : XjkerT  Y, T'[x] = Tx, is one-to-one on XjkerT. We have now shown that XjY is a normed space. Let us now addrs the question of when the quotient space XjY is Banach. 
5.9. Quotient Spaces 289 5.94. Theorem. If X is Banach space, and if Y is a closed (hence complete) subspace of X, then X/Y is also a Banach space. Proof. It suffices to show that X/Y is complete. To prove this, we make use of Proposition 3.24. Let {xn} be a sequence in X for which E  1 q(xn) is an absolutely convergent series in X/Y, Le. E  lllq(xn)llx/y < 00. For each n, choose Yn E Y such that IIxn + Ynllx < IIq(xn)lIx/y + 2- n . Then, E  1 IIxn + Ynllx < 00 and therefore, the series E  1 (xn + Yn) converges to some x in X, since X is complete. Finally, since the quotient map is continuous, E  1 (xn + Yn) = x implies that 00 00 L q(xn + Yn) = L q(xn) = q(x) E X/Y n=l n=l and hence, X/Y is complete. . The converse of Theorem 5.94 is true as we see in Exercise 5.174. In general, the sum of two closed subspaces of a Banach space need not be closed, unless one of the subspaces. is finite dimensional (see Exercise 5.176). For a counterexample in 1 2 , let Xl be the vector space of all real se- quences {Yn}  1 for which Yn = 0 if n is odd, and X 2 be the sequences {zn}  1 for whicb Z2n = nZ2n-l, n = 1,2,.... Then, the spaces Y l and Y 2 defined by Y l = , 2 n Xl and Y 2 = , 2 n X 2 are closed subspaces of 1 2 . Clearly, every sequence {Xn}nl in 1 2 can be written uniquely as a sum of elements of Xl and X 2 . Indeed, if we write {Xl, X2, · · .} - {O, Y2, 0, Y4, 0, . . .} + {Zl, Zl, Z3, 2z 3 , ZS, 3z s , · · .} - {Zl, Y2 + Zl, Z3, Y4 + 2z 3 , ZS, Ya + 3z s ,.. .}, then we have Zl = Xl, Y2 = X2 - Xl, Z3 = X3, Y4 = X4 - 2X3, and so on, which implies the unique representation {Xl, X2, . · .} = {O, X2 - Xl, 0, X4 - 2X3, 0, xa - 3xs, . . .} + {Xl, Xl, X3, 2X3, xs, 3xs, . . .}. If a sequence has all but finitely many terms zero, so do the two summands. Thus, all such sequences belong to Y l + Y 2 , showing that Y 1 + Y 2 is dense in 
290 Chapter 5: Linear Operators on Normed Spaces l2, Le. Y l + Y 2 = l2. If we consider the sequence {I, 0, 1/2,0, 1/3, . . .} E l2, then we note that its only representation as elements of Xl and X 2 is {I, 0,1/2,0,1/3,0,...} = {O, -1,0, -1,0, -I,...} + {I, 1, 1/2, 1, 1/3, 1,.. .}, and so it does not belong to Y l + Y 2 . (We note that {O, -1,0, -I,...}  Y l and {I, 1, 1/2, 1, 1/3, . . .}  Y 2 .) Thus, Y l + Y 2 is not closed in 1 2 . 5.10 Baire Category Theorem Let us recall some basic and standard terminology. Most of the results and examples which we assemble here for the understanding of the Baire Category Theorem are really for metric spaces since every normed space is a metric space. Let Y be a subset of a topological space X. 'rhen the subset Y is said to be nowhere dense if the interior of the closure of Y is empty, Le. int ( Y ) = 0. Clearly, this is equivalent of saying that Y contains no nonempty open set. Thus, Y is nowhere dense iff for every x E Y , and for every € > 0, B(x; €) n (X \ Y) # 0. The subset Y is said to be of first category if Y can be expressed as a countable union of nowhere dense sets; otherwise it is said to be of second category. These are called the Baire Categories. Clearly the empty set 0 is of first category. In particular, this implies that any second category set is nonempty. This observation forms a basis for many existence proofs of second category sets. In a complete metric space, a first category set is called meager while the complement of a first category set is called residual set. Clearly, every nowhere dense subset is of first category. The following simple examples are useful: (i) In IR with usual metric, the subset y = { 1,  ' .. . ,  ' .. . } is nowhere dense since Y = Y U {OJ and int Y = 0. (H) Recall that every nonempty subset Y in a discrete metric space X is both open and closed. Therefore, Y = Y and int Y = Y so that int Y # 0. This observation implies that every non empty subset in a discrete metric space is not nowhere dense. (Hi) Every finite subset Y in IR with usual metric is nowhere dense because Y = Y and int Y = 0 as any open subset in IR is an interval which cannot be contained in Y as Y is a finite set. (iv) The union AU B of two first category sets A and B in a metric space X is also of first category because A = U An, B = U Bn implies AU B = U An U Bn, nEN nEN nEN 
5.10. Baire Category Theorem 291 where An's and Bn's are nowhere dense subsets (and hence An U Bn is nowhere dense for each n). (v) A countable union of first category sets in a metric space X is also of first category. . 5.95. emma. (Cantor Intersection Theorem) Suppose that X be nonempty complete metric space and that {X n } is a decreasing nested sequence of nonempty closed sets in X, i.e Xn+1 C X n , n E N. If the sequence of diameters diam (X n ) converges to zero, then there exists exactly one point x in the intersection nnEN Xn. Proof. For each n E N, let X n be a point of Xn. Our strategy is to show that the sequence {xn} is Cauchy. Since X n is decreasing, the points x n , X n +1, .. . all will be in X n ; Le. for m > n, X m E Xn. By the definition of the diameter, d(xn, x m ) < diam (X n ) for all m > n from which it follows that d(xn, x m ) < diam (X n )  0 as n  00 and so {xm} is a Cauchy sequence. As (X, d) is complete, X m  x E X. We claim that the limit point x is the point that appears in the statement of the theorem, and that x is unique. For each m, X m E X n for m > nand, X n being closed, we see that the limit point x E X n for each n. It follows that x EnnEN Xn. The uniqueness of such x is clear. Indeed, if x, yare points in the intersection then, by the definition of the diameter, we have d(x, y) < diam (X n ) for all n which, as n  00, implies that d(x, y) = 0 and so x = y. . 5.96. Theorem. (Baire Category Theorem) A nonempty com- plete metric space cannot be written as a countable union of nowhere dense sets. Proof. Suppose the statement were false. Then we could write X = U X n ( = U Xn ) nEN nEN where each X n is nowhere dense (Le none of X n contains any open set). Fix an open ball B in X. Then, X 1 cannot contain the ball B. So there exists a point Xl in B \ X 1. Clearly, this set is open and hence, it contains 
292 Chapter 5: Linear Operators on Normed Spaces an open ball B(Xl; €l) of radius €l < 1 such that its closure B(Xl; €l) is contained in B \ X 1 : B(Xl;€1)n X 1 =0. Similarly, there exists a point X2 ft X 2 and a ball B(X2; €2) of radius €2 < €1/2 such that B(X2; €2) C B(Xl; €l), B(X2; €2) n X2 = 0. Clearly, B(X2; €2) n X l = 0. Applying this a rgument i nductively we get a nested decreasing sequence of closed balls B(x n ; €n) disjoint from Xk (1 < k < n), since B(x n ; €n) n X k = 0 for each k = 1,2,. . . , n, and where the radius of the n-th ball B(x n ; €n) is at most 2- n , since €n-l €n-2 €l 1 €n < 2 < 22 < . . · < 2 n - 1 < 2 n - 1 · For m > n we have d(xm,xn) < d(Xn,Xn+l) +... + d(Xm-l,Xm) < €n + €n+l + · · . + €m-l 1 1 1 < 2 n - 1 + 2 n + · · · + 2 m - 2 1 ( 1 + ! + . . . + 1 ) 2 n - 1 2 2 m - n - 1 < 2n1_1 ( 1-\/2 ) = 2 n1 _ 2 ' and therefore, diam (B(x n ; €n))  O. So, by the Cantor Int ersection The- orem, there exists a unique x E X such that x EnnEN B(x n ; €n). As B(x n ; €n) n X n = 0, we have x ft X n for each n. But this would mean x ft Un EN X n = X, a contradiction. - In the language of metric spaces, a reformulation of the Baire Category Theorem is that "A nonempty complete metric space is of second categorY'; Le. if X is a complete metric space and X = U X n , then int ( X n ) :j:. 0 for some n. This equivalent statement implies that the sets of first category in a complete metric space is small in some "topological sense" than that of second category sets. 5.97. Corollary. No closed interval I = [a, b], a < b, in IR is of first category. Proof. This result follows from the Baire Category Theorem as I = [a, b] is a complete metric space. We provide a direct proof of this corollary. 
5.10. Baire Category Theorem 293 If the statement were false, then we could write I = UnEN In where each In is nowhere dense. Let J 1 = [a1, b 1 ] be a subinterval of I with 1 1 n J 1 = 0 and b 1 - a1 < !. In this way, by induction, we get a nested decreasing sequence of closed intervals such that the n-th interval I n = [an, b n ] has a length at most 2- n and is disjoint from 1k for each k. Hence, by the Cantor Intersection Theorem, the set nJ n contains a unique point common to each of the intervals I n . This point can't belong to any 1k which is a contradiction, since every point has to be in 1k for some k. . The following result is a consequence of Corollary 5.97. 5.98. Corollary. Every open interval (a, b) in IR is of second cate- gory. Proof. If (a, b) were of first category, then [a, b], being the union of ( a, b), {a} and {b}, would have been of first category. This is a contradiction to Corollary 5.97. . 5.99. Example. We know that IR with usual metric is a complete metric space and therefore, Theorem 5.96 implies that "the set of real numbers IR is of second category. Further, each single element set {q} in IR has the property that {q}= {q} , int{q}=0 intQ=0. Since Q can be written as a countable union- of singleton sets, Q = UqEQ { q}, the above observation implies that Q is of first category (and so is the set of real algebraic numbers, while the set of real transcendental numbers is of the second category). As a simple consequence of Baire Category Theorem, it follows that ((f is nonempty and therefore, the set of irrational numbers is of second category. . 5.100. Example. Consider the subset Z (of integers) oflR with usual metric. Then, Z is of first category. Indeed, Z has no limit point since for each n E Z the neighbourhood Ix-nl < 1 contains no points of Z other than n itself which shows that Z = Z , since int Z = 0 = int ( Z ), which implies that Z is nowhere dense in IR. (Alternately, since ZC = U  _oo(n,n + 1), the countable union of open intervals, ZC is open in IR; Le. Z is closed as well as Z is nowhere dense in IR). . 5.101. Remark. A set not being nowhere dense does not mean that the set is dense. For example, consider the subset I = (1,2) (or (a, b) with a < b) of the real line IR with usual metric. Then, the closure of I is I = [1, 2] and int I :j:. 0. But the closure of I is clearly not all of IR. 
294 Chapter 5: Linear Operators on Normed Spaces We also observe that, if we have X = A U B where A is known to be of first category and X to be of second category then it can be easily seen that B must be of second category. In particular, since IR = QU]I , it follows that IR \ Q = ]I, the set f all irrationals, is of second category. . More generally, we have 5.102. Corollary. Nonempty open sets of complete metric spaces are always of second category. In particular, the complete metric space }Rn is of second category. This corollary is known before Baire itself. Finally, we state and prove an interesting application of the Baire Category Theorem. 5.103. Theorem. A Banach space cannot have a countably infinite Hamel basis. Proof. Suppose that {cPn} is a countably infinite Hamel basis for a Ba- nach space X. Then each of the finite dimensional spaces Y n = span {cPk : 1 < k < n} is closed. Clearly X = U  1 Y n . But if X is infinite dimen- sional, then each finite dimensional space must have empty interior; that is, finite dimensional subspaces are nowhere dense since Y n = Y n :j:. X for each n. Indeed, since Y n :j:. X for each n, Proposition 3.14 shows that Y n is nowhere dense. This observations proves that X is a countable union of nowhere dense sets, a contradiction to the Baire Category Theorem. Thus, either {cPn} is finite (Le. X is finite dimensional) or {cPn} must be uncountable. _ 5.11 Open Mapping Theorem Our main objective in this section is to establish the Open Mapping The- orem which follows from the Baire Category Theorem and the following lemma. 5.104. Lemma. Let T : X  Y be a boun ded linear operator be- tween Banach spaces. If By(O; €) c T(Bx(O; 1)) for some € > 0, then By(O; €) C T(Bx(O; 2)). Proof. Recall that if Bx(O; €) denotes the open ball centered at the origin in a normed space X and having radius €, then Bx(O; €) = €Bx(O; 1) so that, by the linearity of T, we must have T(Bx(O;€)) = T(€Bx(O; 1)) = €T(Bx(O; 1)). We define U 1 / E = T(Bx(O;€)) so that U 1 = T(Bx(O; 1)). Let y E Y be an arbitrary point in By (0; €) for some € > O. By hypothesis, By (0; €) C U 1 
5.11. Open Mapping Theorem 295 and so there exists YI E U I such that lIy - YIII < €/2, Le. Y - YI E By(O; €/2). Further, YI E U 1 gives that YI = TXI for some Xl with IIXIII < 1. We may recall that By(O;€) C U I <==> By(O;€/2)cT(Bx(O;I/2)= U 2, since by hypothesis By(O; €) C T(Bx(O; 1)). Now, since Y - YI E By(O;€/2) C U 2 , there exists Y2 E U2 such that Ily - YI - Y211 < €/4. Further, Y2 E U2 implies that Y2 = TX2 for some X2 E Bx(O; 1/2). There- fore, by induction, we obtain a sequence {xn} in X such that X n E Bx(O; 1/2n-I), TX n = Yn and IIY-(Yl+".+Yn)lI= Y-  TXk - Y-T(  Xk) € <- 2 n so that n E Yk  Y as n  00. k=l Now, for each n, we define Sn = L:=l Xk, and note that n+p Ilsn+p - snll < E IIXkl1 k=n+l n+p 1 < E 2 k - 1 k=n+l -  ( 1 + ! + · · · +  ) 2 n 2 2p-1 _ 2-n+I(1 - 2- P ) which shows that {sn} is a Cauchy sequence in the complete space X and hence converges in X. Let X = lim n -+ oo Sn = L:  I x n in X. Note that 00 00 1 IIxll < E IIxnll < E 2 n - 1 = 2 n=l n=l 
296 Chapter 5: Linear Operators on Normed Spaces and therefore, x E Bx(O; 2). By the continuity and linearity of T, we see that n n Tx = lim T(sn) = lim  TXk = lim  Yk = Y, noo noo noo k=1 k=1 which gives that y E T(Bx(O; 2)) and hence, By(O; €) C T(Bx{O; 2)). . We observe that the same method of proof of Lemma 5.104 works to prove the statement with 2 replaced by any number greater than 1. How- ever, the statement in Lemma 5.104 above suffices for the proof of the open mapping theorem. 5.105. Lemma. Let T : X  Y be a surjective bounded linear operator between Banach spaces X and Y. If zero is an interior point of a subset S of X, then zero is also an interior point ofT(S). Proof. If zero is an interior point of a subset S, then there exists an € > 0 such that €Bx{O; 1) C S. Then, by the linearity of T, we must have T(€Bx{O; 1)) = €T{Bx(O; 1)) c T{S). Therefore, to establish the result, it suffices to show that zero is an interior point of T(Bx(O; 1)). We note that for each x E X we can choose n E N such that x E B x (O; n), since 00 00 X = U Bx{O; n) = U nBx(O; 1). n=1 n=1 Further, since T is onto, {T(Bx(O;n))}nEN covers Y so that 00 00 Y = U T(Bx(O;n)) = U nT(B}((O; 1)). n=1 n=1 Since Y is complete, by Baire's Category Theorem, the closure of one of them (say T(Bx(O; no))) has a nonempty interior and hence must contain an open ball in Y. Again T(Bx(O;no)) = noT{Bx(O; 1)) and note that the map y I-t noY being a homeomorphism of Y, T (B x (0; 1)) must contain an open ball in Y. Therefore, there exists Yo E T(Bx{O; 1)) and € > 0 such that By{yo; €) = Yo + €By{O; 1) c T{Bx{O; 1)). Since T(Bx(O; 1)) and By(O; €) are symmetric, we have -Yo + €By(O; 1) c T(Bx(O; 1)) 
5.11. Open Mapping Theorem 297 (we recall that a set A is symmetr ic iff x E A implies that -x E A). IT Y E By(O; f), then the convexity of T(Bx(O ; 1)) gives th at Y = Yo : Y + -Y0 2 + Y E T(Bx(Oj 1». Consequently, By(O; €) C T(Bx(O; 1)). By Lemma 5.104, we have By(O; €) c T(Bx(O; 2)), Le. By(O; €/2) C T(Bx(O; 1)), which shows that zero is an interior point of T(Bx(O; 1)). - 5.106. Theorem. (Open Mapping Theorem) A surjective bounded linear operator T : X  Y between two Banach spaces is an open mapping. Proof. Let G be open subset of Y. If G = 0, then T(G) = T(0) = 0 so that T(G) is open in this trivial case. Next, we assume that G # 0. Since G is open, for each Xo E G, we have Xo + €Bx(O; 1) = Bx(xo; €) C G for some € > 0, Le. Bx(O; €) C G - Xo so that 0 is an interior point of G - Xo. Suppose that y E T(G). Then, there exists an x E G such that y = Tx, and note that T(G) - y = T(G) - Tx = T(G - x) By Lemma 5.105, zero is also an interior point of T(G - x) = T(G) - y which is same as to say that y is interior point of T(G). _ An immediate consequence is the following 5.107. Theorem. (Banach Isomorphism Theorem) Let T be one-to-one bounded linear operator T from a Banach space X onto a Banach space Y. Then it has a bounded linear inverse. In particular, it is a homeomorphism (so is an isomorphism). Proof. We already know that the inverse map T-l is linear. Since T is one-to-one with RT = Y (since T is onto), it remains to show that T-l : Y  X is continuous. Suppose G is an open set in X. Then, as T is bijective, the inverse image of Gunder T- 1 is (T-l)-l(G) = T(G). But by the Open mapping theorem, it follows that T(G) is open and hence T-l is continuous. _ This theorem is also called bounded inverse theorem. The following example illustrates the necessity of the underlying space to be complete. 
298 Chapter 5: Linear Operators on Normed Spaces 5.108. Example. Let X = (Coo, 11.11(0). We have already seen that X is not complete. Define two operators T and S as follows: T : X  X, {Zn}n>l t--+ { Zn } - n nl and S : X  X, {Zn}nl t--+ {nZn}nl . Clearly, both T and S are linear and inverse of each other. Moreover, T is bounded with IITI! < 1 because IIT( {zn} )1100 = sup I Zn I < sup IZnl = lI{zn}lIoo. nEN n nEN Define Zn = {I, 1, . . . , 1,0,0, . . .} where 1 appears in the first n places of Zn. Note that IIZnlloo = 1 for each Zn E X, n E N. But S(Zn) = {I, 2, 3, . . ., n, 0, 0,. ..} and IIS(Zn)lIoo = n showing that S = T- 1 is unbounded. . As an example of one of the many applications of the Banach theorem we prove the following useful characterization of closed embedding of Banach spaces. 5.109. Theorem. .Let X and Y be Banach spaces and T E B(X, Y). Then T is injective and R T is closed in Y iff there exists a G > 0 such that IIxll < GIITxll for all x EX. Proof. (=»: Since closed subspace of a complete space is complete, RT is a Banach space. Suppose that T is one-to-one with closed range. Consider the map T- 1 : R T  X. It is the inverse of a bounded isomorphism between X and RT, by Theorem 5.107. Therefore, there is a positive real number G > 0 such that IIT-lyll < Gllyl! for all y E RT; Le. T-l : RT  X is bounded. Replacing y by Tx we obtain the required inequali ty. ( {=): If the inequality holds, then T is clearly one-to-one, and if {Txn} is a Cauchy sequence in R T then {x n } is Cauchy by hypothesis. Hence, {xn} converges to some x E X and because T is continuous, {Tx n } converges to Tx. Therefore, RT is complete and hence, RT is closed. - Our next example illustrates the application of Theorem 5.109 (see also Example 5.68). 
5.12. Closed Graph Theorem 299 5.110. Example. For X = (C[O, 1], 11.1100)' consider the linear op- erator T E L(X) defined by f(t) I-t J f(s) ds (t E [0,1]). Clearly, T is bounded. If we write 9 = Tf, then g(O) = 0, g'(t) = f(t) and, Tf = 0 implies that f(t) = 0 in [0,1]. Therefore, we conclude that T is injective and, the range space is given by RT = {g E C 1 [0, 1] : g(O) = OJ. Now, by Theorem 5.109, RT is not closed. Indeed, the sequence {fn} defined by fn(t) = nt n - 1 shows that fn E X, Ilflloo = nand IITfnlloo = 1 for each n E N. Thus, there exists no C > 0 such that Ilfnlloo < CllTfnlloo for large value n. Hence, by Theorem 5.109, RT is not closed. . 5.12 Closed Graph Theorem Before we establish the Closed Graph Theorem with the help of the Open Mapping Theorem, we need some preparations. Let T : D C X  Y be an operator between the two sets X and Y. Then the subset GT={(x,Tx): XED}cXxY is called the graph of T. Logically a graph of T is really the same as the mapping T itself, regarded as a set of ordered pairs. Suppose that X and Yare two normed spaces and T : D C X  Y is a linear operator. We recall that X x Y is a vector space with the following operations (Xl, Yl) + (X2, Y2) = (Xl + X2, Y1 + Y2) A(xl, Yl) = (AXl, AYl). Further, we see that it is a normed space with the norm II (x, y)lIxxY = max{lIxllx, Ilylly}. Because of the linearity of T it follows that its graph G T is a linear subset of X x Y. The linear operator T : D C X  Y is called a closed operator if G T is a closed subspace of X x Y: Le. GT is closed iff X n E D c X (domain of T) for all nand X n  x E X, and TX n  Y E Y => xED and y = Tx. If T is a bounded operator, then it is easy to verify that G T is a closed subspace of X x Y. Even an unbounded linear operator may have closed graph, see Example 5.119. 5111. Proposition. Suppose that X and Y are Banach spaces. Then X x Y with respect to the supnorm II (x, y)llxxY = max{llxllx, lIylly} is also a Banach space. 
300 Chapter 5: Linear Operators on Normed Spaces Proof. If {( X n , Yn)} is a Cauchy. sequence in X x Y, then the equation max{lIxm - xnll, IIYm - YnJ!} = II(xm,Ym) - (xn,Yn)1I shows that {xn} and {Yn} are Cauchy sequences in X and Y, respectively. As X and Yare Banach spaces, the sequences {xn} and {Yn} must converge to some points x E X and Y E Y, respectively. Again, the equation lI(xn, Yn) - (x, y)1I = max{lIxm - xII, llYn - ylI} implies that the sequence {( X n , Yn)} converges to (x, y) E X x Y as n  00. Thus, X x Y is complete. _ 5.112. Proposition. Suppose that X and Y are Banach spaces. Then X x Y is also a Banach space with respect to norm lI(x,y)llxxY = vl llxlI + lIyJI}. Proof. Observe the inequalities IIxm - xnllx < II (xm, Ym} - (xn, Yn}IIXxY = Vllxm - xnlli + IIYm - Ynll, IIYm - Ynliy < lI(xm,Ym) - (xn,Yn}IIXxy = V llx m - xnlli + IIYm - Ynll and the equation lI(xn, Yn) - (x, y}1I = lI(xn - x, Yn - y}lIxxY = V llx n - xIIi + llYn - ylI. Use the idea of the last proposition to obtain the desired conclusion. _ More generally we have following result which can be readily checked. 5.113. Proposition. H X and Y are Banach spaces, then also is X x Y with respect to norm lI(x,y)lIxxY = (lIxll + Ilyll)l/p for each p > 1. We remark that all these norms in the last three propositions are equiva- lent. Now, we state and prove the following simple test for closed operators. 5.114. Proposition. Let X and Y be normed spaces and T : D C X  Y be linear. Then T is closed iff the following condition holds: X n E D, X n  x E X, and TX n  Y E Y => xED and Y = Tx. Proof. =>: Let T be closed and let GT = {(x,Tx) : x E X} be its graph. To show that G T is closed, we must show that a limit of GT is 
5.12. Closed Graph Theorem 301 actually a member of G T . Let (x, y) be a limit point of G T . Then there exists a sequence of points, (x n , Tx n ) of GT, where X n ED, such that lI(x n - x, TX n - y)lIxxY = lI(x n , Tx n ) - (x, y)llxxY  0 as n  00. Taking the norm on X x Y (for instance, p = 1 of Proposition 5.113), we have IIxn - xlix + IITx n - ylly  0 as n  00 which implies that X n  x and TX n  y as n  00. Since T is closed, this gives xED and y = Tx, and so (x,y) = (x,Tx) E G T . Thus, G T is closed. {=:: Conversely, let G T be closed. To show that T is closed, we let X n E D for all n, X n  x and TX n  y as n  00. We must show xED and y = Tx. Our assumption implies that every neighbourhood of (x,y) contains a point of G T so that (x n , Tx n )  (x, y) E G T . Since GT is closed, G T = G T , and so we obtain that (x, y) E G T . By the definition of G T , this gives xED and y = Tx. . The next result is important. It will be used often to prove the continuity of a linear operator. 5.115. Theorem. (Closed Graph Theorem) A linear operator between two Banach spaces is bounded iff its graph is a closed subset of X x Y. Proof. (=»: Let T : X  Y be a linear operator. Suppose that T is continuous. Then X n -+ x implies that TX n  Tx showing that T is closed and therefore, its graph is a closed subset of X x Y. ({=:): Let G T = {(x,Tx) : x E X} C X x Y denote the graph of T. Since G T is a closed subspace of the Banach space X x Y (being a product of two Banach spaces with respect to the above supnorm), G T itself is a Banach space. Consider the operators defined by IIx : GT  X, (x, Tx) I-t x, and IIy : G T  Y, (x, Tx) I-t Tx. 
302 Chapter 5: Linear Operators on Normed Spaces Clearly, IIx is a linear bijection. It is easy to see that both IIx and IIy are bounded linear operators. Indeed, the boundedness follows from the fact that IIIIx(x, Tx)llx = Ilxllx < max{llxllx, IITxlly} = lI(x, Tx)lIxxY and similarly, IIIIy(x, Tx)lly = IITxlly < max{llxllx, IITxlly} = lI(x, Tx)lIxxy. Further, the inverse maps IIx l is bounded by the Open Mapping Theo- rem. Note that To IIx = IIy and the composition X  G T C X X Y  Y is simply T = IIy 0 II:x 1 , which is a composition of two continuous oper- ators, so T is continuous. Alternately, the boundedness of II:x 1 and the inequality IITxlly < II(x,Tx)lIxxY = IIII:x 1 (x)lIxxy < IIII:x 1 1111 x llx can be used to obtain that T is bounded. . 5.116. Example.. Let {Ak}k1 be a sequence of scalars such that 00 E IAkZkl < 00 for all Z = {Zk}k1 Ell. k=1 Consider the mapping T : II  II via the form T(Z) = T( {Zk}kl) = {AkZk}kl. J Clearly, T is linear. Is it bounded? For the boundedness of T, by the Closed Graph Theorem, it suffices to show that G T is closed. Let (Z, W) E GT , where W = {Wk}kl. Then there exists a sequence {Zn}nl, Zn = {zn(k) }kl, such that Zn  Z and T(Zn)  W in ll. Hence, for each kEN, we have zn(k)  Zk and AkZn(k)  Wk as n  00. Consequently, for each kEN, Wk = lim AkZn(k) = AkZk n-+oo which implies that T(Z) = W so that (Z, W) E GT. Therefore, GT is closed and hence, T is continuous on ll. . 
5.12. Closed Graph Theorem 303 In order to prove that T is bounded, the Closed Graph Theorem is often used in the following form: 5.117. Corollary. Let T : X  Y be a linear operator between the Banach spaces X and Y such that x n  0 and TX n  y as n  00 :::} y = o. Then T is bounded. Proof. Since X n  x and TX n  y implies that Tx = y, the graph of T is closed. Hence, T must be bounded by the Closed Graph Theorem. _ 5.118. Example. From 5.76, it follows that a closed linear operator T may fail to be continuous unless both X and Yare Banach spaces. Further, in the statement of the Closed Graph Theorem, it is also essential that the operator T whose graph is considered is linear. For example, let X = Y = IR with absolute function I · I as norm. Consider the map I : IR  IR defined by { l/x f(x) = 0 if x :F 0 if x = o. Clearly, I is not continuous on IR. However, the graph of I is given by Gj = {(x, I(x)) : x E IR} = {(O,O)} U {(x, l/x) : x E IR\ {a} } which is a closed subset of IR 2 = IR x IR with respect to the I-norm 1I(.,.)IIRxR = I.IR + I.IR. . 5.119. Example. From Example 5.76(2), we see that there exists an unbounded linear operator having a closed graph. Indeed, for D - (C 1 [0,1], 11.11(0), X = Y = (C[O, 1],11.11(0)' the operator T : D c X  X, I t--+ I' is clearly linear but not continuous. To show that the graph of T is closed, let In E D for all n E Nand In  I E X and T In  9 in X. Note that the convergence with respect to the supnorm is uniform so that I = Tin converges to 9 uniformly in X. Since In converges I, we have that I is differentiable and I' = g. Thus, the graph of T is closed. Does this example contradict the closed graph theorem? . 
304 Chapter 5: Linear Operators on Normed Spaces Our main interest in the remaining part of this section lies in Banach spaces and continuous projections. In particular, if X is a normed space and M is a subspace of X, does there always exist a continuous linear projection map such that M = Rp? Of course, the following simple arguments confirm that the range of continuous linear projection must be closed: if PX n  x, then, by the continuity of P, PXn = p2xn = P(Pxn)  Px and hence, we must have Px = x, Le. x E Rp so that M = Rp is closed. Moreover, N = Np is closed if P is continuous and since, for each x, x = x - Px + Px where x - Px E N p and Px E Rp, we have X = M fB N. Now, we discuss the converse part. 5.120. Proposition. H X is a Banach space, M c X is closed and X = M fB N for some closed subspace N of X, then the corresponding projection P with N = N p is necessarily continuous. Proof. Let X = M fB N. Then every x E X has a unique representation x = m + n, m E M, n E N. Define P : X  X by Px = m, x EX. We have already proved that such a map is linear, p2 = P, Rp = M and N p = N. It remains to show that P is continuous which is, in fact, an easy consequence of the Closed Graph Theorem. To apply the Closed Graph Theorem, it suffices to show that P is closed. To do this, we suppose {x j} eX, Xj = mj + nj E M fB N (j = 1,2,.. .), where Xj  x and PXj  y as j  00. Then for each j, PXj E M and, because M is closed, we must have y E M. We need to show that y = Px, and for this we observe that Xj = mj + nj  x = m + n E M fB N as j  00, which implies that nj = Xj - mj = Xj - PXj  x - yEN as j  00, 
5.13. Uniform Boundedness Principle 305 since N is closed. Therefore, x=y+(x-y), YEM, x-yEN and the uniqueness of the representation x = m + n yields that y = m = Px E M. Thus, Rp is closed in X x X which, by the Closed Graph Theorem, implies that P is continuous. - From Propositions 5.22 and 5.120, we conclude the following basic char- acterization theorem which gives the one-to-one correspondence between projections and the complementary subspaces. 5.121. Theorem. Let X be a Banach space, and M c X be a closed subspace. Then, M is complemented in X in the sense that there is another closed subspace N c X such that X = M E9 N. Equivalently, M is complemented in X iff M is the range of a continuous linear projection P on X. Note that, a projection P on a Banach space X is -always a linear con- tinuous operator on X. 5.13 Uniform Boundedness Principle The Uniform Boundedness Principle (or the Banach-Steinhaus Theorem) also follows from the Baire Category Theorem. 5.122. Theorem. (Uniform Boundedness Principle) Let X, Y be two Banach spaces. Suppose that S = {Ta}aEA c B(X, Y) is a family of bounded linear operators from X into Y. If, for each x EX, the set {Tax} is a bounded subset ofY (i.e. SUPTES IIT(x)lIy < 00 for all x EX), then the set {IiTall} is bounded (i.e sUPTES IITII < 00 ). Proof. Let B = {x EX: II Tax II < 1 for a: E A}. Then B is closed, since each Ta is continuous. By hypothesis, for each x EX, there exists n such that liT axil < n for all a: E A so that n-1x E B. Thus, for n E N, if we let X n = {x EX: IITaxll < n for a: E A}, 
306 Chapter 5: Linear Operators on Normed Spaces then it is easy to see that x = U X n = U nB. nEN nEN In fact, if this were not true then there would exist a point x E X such that x  X n for each n so that IITax11 > n for all n and for some a = a(n). This would then mean that the sequence {Tax} would not be bounded, a contradiction. As X is complete, the Baire Category Theorem shows that one of the sets, say X N, contains an interior point. We show that XN is closed. Let {Xk} be a sequence in XN such that Xk  x as k  00. Then, for each a E A, we have IITa(Xk)11 < N for k = 1,2, . . . and, by the continuity of the norm on Y, we get lim IITa(Xk) = IITax11 < N k-+oo showing that x E XN. Thus, XN is closed. Since X N = XN and int (XN)  0, XN must contain an open ball B(a; c5) for a suitable a E X and c5 > o. We claim that 2N liT a II < T for a E A. Now, for any y E X with lIyll < 1, we have . x - a = c5y with x E B(a;c5) and therefore, IITa(a + c5y)1I = IITaa + c5Tayll < N which implies that c5I1 T ayll < N + IITaali. We note that a E B(a; c5) and so, IITaall < N. This observation shows that 2N IITal1 < T for a E A where c5 and N are fixed constants. This shows that IITall is uniformly bounded by M = 2N / c5, where M is independent of n. - In a simple language, the Uniform Boundedness Principle implies that "a set of linear operators defined on a Banach space which is bounded point- wise is unifonnly bounded." An interesting and surprising application of the Uniform Boundedness Principle is given in Section 6.10. 
5.14. Extension of Continuous Functionals 307 5.14 Extension of Continuous Functionals One of the fundamental principles in Banach space theory for dealing with dual spaces of normed spaces is the Hahn-Banach Theorem. 25 Without the Hahn-Banach theorem, functional analysis would be very different from the structure that exists today. The Hahn-Banach theorem, which is a favourite of almost every analyst, assures us the existence of the extension of a continuous linear functional defined on a subspace of normed space X, and hence implies that the dual space of a nontrivial normed space has sufficiently rich structure. Remember that the use of the term "extension" is standard, meaning that a function F defined on a super set X of M is an extension of f defined on M if FIM = f, Le. f(y) = F(y) for all y E M. We first note that the existence of nontrivial continuous linear functionals on a subspace of a normed space X is easy to obtain. In fact, for 0 :j:. Xo EX, let M = span { xo} and consider f on M by f(y) = a:llxoll, y = a:xo E M. It is easy to see that f is a linear functional on M and If(y)1 = Ilyll so that IIfll = 1. However, the extension of f to the whole space is not so easy. The Hahn-Banach theorem which we shall soon discuss will give rise to a number of applications which includes a procedure for extending a linear functional defined on a subspace M of a given normed space X. Let us first state the precise definition of a linear extension. There are many ways for extending a given linear map. Let X and Y be two vector spaces, M a subspace of X and f : M  Y a linear map. Then we say that a linear map F : X  Y is an extension of f to X if FIM = f. Here our particular emphasize will be when X is a normed space, Y = 1F and M, a subspace of X. Thus the problem of extension for our setting may be formulated as follows: "A (continuous) linear functional F on a normed space X is called ( a continuous extension) an extension of a given (continuous) linear func- tional f defined on a subspace M of X if FIM = f." Note that the exten- sion of a continuous linear functional never decreases its nrms. Indeed, IIFII - sup{IF(x)l: x E X, Ilxll = I} > sup{IF(x)l: x E M, Ilxll = I} - Ilfll. One of our objectives in this section is to look for extensions of functionals with the same norm, Le. IIFII = Ilfll. However, we obtain such extensions as special cases. The construction of an extension of a continuous linear 25Hahn (1879-1934) was a pioneer in set theory and functional analysis. Hahn is best remembered for the Hahn-Banach theorem although he has made important contribu- tions to the theory of calculus of variations. 
308 Chapter 5: Linear Operators on Normed Spaces functional defined on dense subsets is somewhat easier to prove (see Theo- rem 5.123). However, we shall consider a more complicated case with the help of the Hahn-Banach theorem (see Theorem 5.124). In complex analysis, the concept of analytic continuation concerns with analytic extensions of a given function defined on D C C. At this place it is important to point out that the numerical implementation of analytic continuation has applications to a large number of problems including the one which concerns the computation of transonic flow past aerofoils. Note that this has not much to do with linear functionals although it is a kind of extension. At this point, it is important and natural to ask the following ques- tion: Are there "enough" elements in X* = B{X,]F)? In fact, an interest- ing application is that there are enough continuous linear functionals, for example, to separate points of X; meaning that for every pair of points x, Y E X, x # Y, there exists a continuous linear functional F on X such that F(x) # F(y) (see Corollary 5.145). This separation property (as well as many others) is essential in a variety of applications in the subject of modern analysis. We start with the following result which is considered to be the simple version of the Hahn-Banach theorem but is usually not called the Hahn- Banach theorem. 5.123. Theorem. Let X be a normed space, M a dense subset of X, and f E M* = B(M,]F). Then f can be extended to a unique F E X* such that IIfll = IIFII. Proof. By hypothesis, M = X. Therefore, for every x EX, there exists a sequence {Yn} in M such that Yn  x as n  00, Le. llYn - xlix  0 as n  00. Define F(x) = lim f(Yn). n-+oo Then it is easy to check that this definition is well defined. Indeed, since every convergent sequence is Cauchy, from the inequality If{Yn) - f(Ym)1 < IIflillYn - Ymll, it follows that {f(Yn)} is a Cauchy sequence in IF and hence converges. Moreover, if {y} is another sequence in M with Y  x then the inequality If(Yn) - f{y)1 < IIflillYn - y1I < IIfll [llYn - xII + lIy - xiI] implies that Urn f{Yn) = lim f(y). n-+oo n-+oo 
5.14. Extension of Continuous Functionals 309 These two observations show that the definition of F(x) is well defined. For each Y EM, let c(y) denote the corresponding stationary/constant sequence, all of whose terms are Y, Le. c(y) = {Yn} with Yn = Y for all n. Clearly, this sequences is convergent with the limit y. In particular, F(y) = I(Y) for all y E M, Le. FIM = I. The linearity of F follows from the fact that I is linear and the addition and the scalar multiplication are continuous on a normed space . Thus, to complete the proof, we need to show that IIFII = 11/11 and that F is unique. Allowing n  00 in the inequality I/(Yn)1 < 11/111IYnll, we obtain that IF(x)1 < 1I/IIIIxii which holds for each x EX, and therefore IIFII < 11/11. We know that the extension of a functional never decreases its norm so that 11/11 < IIFII. Combining the last two inequalities, we find that IIFII = 11/11. It remains to show that F is unique. Suppose that F and G are two continuous linear extensions of I, and {Yn} is a sequence in M with Yn  x, where x E X. By virtue of the continuity of the functionals, we see that F(x) = lim F(Yn) = lim I(Yn) = lim G(Yn) = G(x) n-+oo n-+oo n-+oo and therefore, F = G. . Let X be a vector space over 1F (1F = IR or C) and p, a real valued function defined on X such that (i) p(x + y) < p(x) + p(y) for all x, Y E X. (ii) p(a:x) = Ia:lp(x) for all x E X and a: E F. Here p is called a seminorm and the pair (X,p) is called a seminormed space. We have p(x) = p(x - Y + y) < p(x - y) + p(y), Le. p(x) - p(y) < p(x - y) and, similarly p(y) - p(x) < p(y - x) = p( -(x - y)) = p(x - y) 
310 Chapter 5: Linear Operators on Normed Spaces so that Ip(x) - p(y)1 < p(x - y) for all x,y E X. Putting y = 0 in the last inequality and noting that p(O) = 0 (take a = 0 in (ii) above), we obtain p( x) > 0 for all x EX. When p is a seminorm such that p(x) > 0 for each x # 0, we call it a norm and the corresponding pair (X,p) is called a normed space. This notion is a generalization of the normed space which we have studied under a special case p(x) = IIxU. There are two main versions of the Hahn-Banach theorem. Of course the first version of the Hahn-Banach theorem follows from the second version. For reasons of convenience, we do not always specify which of the two we have in mind when we refer to the Hahn-Banach theorem. Also, there are many equally elegant proofs available. Now we proceed to prove the Hahn-Banach theorem. 5.124. Theorem. (Main versions) Let X be a vector space over 1F (1F = IR or C) and p, a seminorm defined on X as above. Let M be a subspace of X, and let f be a linear functional on M with If(z)1 < p(z) for all z E M. Then f can be extended to a linear functional F defined on all of X such that IF(x)1 < p(x) for all x E X. First, we remark that Theorem 5.124 has two main parts, one for real vector spaces (called the first version of the Hahn-Banach Theorem) and the other for complex vector spaces (called the second version of the Hahn- Banach Theorem). By a real linear functional on a complex vector space X, we mean the following: If we restrict ourselves to the multiplication by real numbers only on X, then we regard X as a real vector space and may be denoted by XR. Thus, we say that 9 is a real linear functional on the complex vector space X if g(ax) = ag(x), for a E IR and for any x E X. Moreover, if 9 = U + iV is a complex valued linear functional on X then both U = Re 9 and V = 1m 9 are real linear . Indeed, for a E IR, U(ax + y) + iV(ax + y) - g(ax + y) - ag(x) + g(y) (since 9 is linear ), - a[U(x) + iV(x)] + [U(y) + iV(y)] - [aU(x) + U(y)] + i[aV(x) + V(y)] 
5.14. Extension of Continuous Functionals 311 and the linearity of U and V follows if we equate the real and the imaginary parts on both sides. It is interesting to observe that the Hahn-Banach theorem for complex vector spaces was proved only after eight years of the appearance of the version for real vector spaces. Later in this section we state a number of results which are consequences of Theorem 5.124. The first result (see Theorem 5.134) tells us that any continuous linear functional defined on a subspace can be extended to continuous linear functional on the whole space with the preservation of its original norm. Proof. The idea of the proof is to use Zorn's lemma (see Lemma 1.4). To do this, we let E be the collection of all ordered pairs (Yo:, 90:) such that (i) Yo: is a subspace of X containing M (ii) 90: is a linear functional on Yo: satisfying . 90:(Z) = I(z) for all z E M · lyo:(y)1 < p(y) for all y E Yo:. Obviously the collection E is nonempty, since (I, M) E E. We can make E into a partially ordered set by defining an order relation  on E as follows: (Yo:, 90:)  (Y,a, y,a) <==} 9,a is an extension of 90: Le Yo: c Y,a and 9,a Iyo = 90:. A straightforward verification shows that the above defined relation  is a partial order. Now, we consider the totally ordered set (chain) J = {(Y A ,9A) : A E A} of E, Le. for every pair of elements a, (3 E A either (Yo:, 90:)  (Y,a,9,a) or (Y,a,9,a)  (Yo:, 90:)' Define Y = U Y A . AEA It follows that Y is a subspace of X (in general, of course, the union of a family of subspaces need not be a subspace, but these are nested). Next, we define 9 on Y by letting 9(x) . 9A(X) if x E Y A . Clearly, 9 is a linear functional. Moreover, 9(x) = I(x) for all x E M and ly(x)1 < p(x) for each x E Y. 
312 Chapter 5: Linear Operators on Normed Spaces Thus, we have (Y,g) E E. In addition, (Y,g) is an upper bound for J. Hence, every chain of E has an upper bound in E and so, by Zorn's lemma (see Lemma 1.4), it has a maximal element, say (X o , G). The theorem is proved once we show that Xo = X (so, in that case, we may take F to be G). Suppose that Xo :j:. X. Then there exists an x E X which is not in Xo. Let Mo be the subspace spanned by x and Xo: Mo = span {x} +Xo. We claim that there exists a linear functional H on Mo such that . H(x) = G(x) on Xo . IH(y)1 < p(y) on Mo. But then, (Mo, H) E E and (X o , G)  (Mo, H) with Xo :j:. Mo, thereby contradicting the maximality of (Xo, G). Let us give a proof of this claim. To do this, since each element y E Mo may be expressed uniquely in the form y = ax + z, we define a functional H on Mo by (5.125) H(ax + z) = ac + G(z), z E Xo, where c = H(x) is a real number that is to be chosen such that (5.126) IH(ax + z)1 < p(ax + z), z E Xo, (Observe that if y E Xo, then a = 0 so that H = G on X o ). Clearly, it is easy to verify that H is linear on Mo for any constant c. However, the issue is to choose a proper constant c = H(x) which can guarantee the validity of inequality (5.126). In order to choose such a constant, let us assume that X is a vector space over III For the proof of (5.126), it suffices to show that (5.127) H(ax + z) < p(ax + z), z E Xo, because (5.127) implies that -H(ax + z) = H( -ax - z) < p( -ax - z) = p(ax + z). We shall soon verify that, for the proof of (5.127), it suffices to restrict a = :i:1; that is, (5.128) H(x + z) < p(x + z), for z E Xo and (5.129) H( -x + z) < p( -x + z) = p(x - z), for z E Xo. 
5.14. Extension of Continuous Functionals 313 By (5.125), the last two inequalities are equivalent to (5.130) c = H(x + z) - G(z) < p(x + z) - G(z), for each z E Xo and (5.131) c = G(z) - H(z - x) > G(z) - p(x - z), for each z E Xo. Thus, we need to find a value c = H(x) E IR such that G(z) - p(x - z) < c < p(x + z) - G(z) for all z E Xo. To prove this, we use the following inequality which holds for any u, v E Xo: (5.132) - -G(v) - p(x + v) < p(x + u) - G(u), u, v E Xo. Indeed, the fact that G is linear on Xo and IG(x)1 < p(x) on Xo yield G(u) - G(v) - G(u - v) < p(u-v) - p(u + x + (-v - x)) < p(u+x)+p(-v-x) - p(u + x) + p(v + x) which means (5.132) holds. Now, in (5.132), suppose that u is fixed while v is allowed to vary through all of Xo. Then the set of all real numbers {-G(v) -p(x+v): v E Xo} has an upper bound and therefore, A = sup{ -G(v) - p(x + v) : v E Xo} exists. Similarly, fixing v and allowing u to vary over Xo, we find that B = inf{p(x + u) - G(u) : u E Xo} exists, so that A < B. Hence, there exists a real number c such that A < c < B. In the event A = B, c is the common value. In fact, by the construction of A and B, for any number c in the in- terval [A, B], both (5.130) and (5.131) (and hence (5.128) and (5.129) ) are satisfied. Finally, using (5.128) and (5.129) we verify (5.127) for each a E IR by considering three cases: Case (i) : Let a > o. Then, by (5.128), H(ax + z) = aH(x + zla) < ap(x + zla) = p(ax + z). 
314 Chapter 5: Linear Operators on Normed Spaces Case (ii) : Let a: < O. Then, by (5.129), , H(a:x + z) = -a:H( -x - z/a:) < -a:p( -x - z/a:) = p(a:x + z). Case (iii) : If a: = 0, we have H(z) = G(z) < p(z) on Xo. Thus, (5.126) holds for all a: E IR and z E Xo. That is, we have extended G defined on Xo to H defined on Mo strictly containing Xo such that IH(x)1 < p(x) for x E Mo. Suppose that X is a vector space over ]F = C. It is worthwhile to remember the trick that is used here. As f is C-valued linear functional, we split f into its real and imaginary parts f(z) = u(z) + iv(z) where u = Re f and v = 1m fare IR-valued linear functions on M, considered as a subspace of XR the real vector space. Because of the complex linearity of f, v(z) = -Re (if(z)) = -Ref(iz) = -u(iz) so that we have ( 5.133) f(z) = u(z) - iu(iz). Moreover, lu(z)1 = IRef(z)1 < If(z)1 < p(z) for all z E M. From what we have proved in the case of the real vector space over IR, there exists an IR-valued linear functional U defined on XR extending u such that IU(z)1 < p(z) on XR. Guided by (5.133), we define F : X  C by F(z) = U(z) - iU(iz) where U is IR-linear on XR. We show that this is the desired eension. It is easy to check the following: . F is C-linear (because F is IR-linear with F(iz) = iF(z)) Indeed, since U(x + y) = U(x) + U(y), F has the additivity property. Now, for a: = a + ib E C, and x EX, we have F(a:x) - F(ax + i bx) - U(ax + i bx) - iU(i(ax + i bx)) - U(ax) + U(i bx) - i[U(i ax) + U( -bx)] - [U(ax) - iU(i ax)] + [U(i bx) - iU( -bx)] - a[U(x) - iU(i x)] + b[U(i x) + i U(x)] - (a + i b) [U(x) - i U(i x)] - a:F(x), which shows that F is C-linear. 
5.14. Extension of Continuous Functionals 315 . The functional F extends I. Since U is extension of u for x EM, F(x) = U(x) - iU(ix) = u(x) - iu(ix) = I(x). Finally, for x EX, let us write F(x) = IF(x)le i9 , where () = argF(x). By the complex linearity of F, it follows that IF(x)1 - e- i9 F(x) _ F(e- i9 x) _ ReF(e- i9 x) _ U(e- i9 x) < p(e- i9 x) = le- i9 Ip(x) = p(x) and we complete the proof. _ 5.134. Theorem. Let X be a normed space, M a subspace of X and I E M* = B(M,]F). Then I can be extended to some F E X. = B(X,]F) such that 11/11 = IIFII. Proof. Define p on X by p(x) = 1I/IIIIxil for all x E X, where 11/11 = sup{l/(x)1 : x E M, IIxll < 1}. Then, we have (i) p(x) > 0 on X (ii) For all x,y EX, p(x + y) = 1I/IIIIx + yll < 11/11 (lIxll + IlylD = p(x) + p(y) (iii) p(ax) = 1I/IIIIaxii = 1I/1I{lalllxll} = lalp(x), and therefore, p is a seminorm on X. Also I/(x)1 < 1I/IIIIxil = p(x), for x E M showing that p satisfies the required conditions of Theorem 5.124. Thus, by Theorem 5.124, there exists a linear functional F on X such that F = I on M which implies that IIFII < 11/11. 
316 Chapter 5: Linear Operators on Normed Spaces Since F is the extension of I, as noted in the beginning, we have 11/11 < IIFII. Hence, IIFII = 11/11 and F is the desired extension of I to all of X. . 5.135. Theorem. Let M be a subspace of a normed space X and let Xo E X be such that d = d(xo, M) = inf Ilxo - mil > o. mEM Then there exists a continuous linear functional F on X so that IIFII = 1, F(M) = {OJ and F(xo) = d. Proof. Since d > 0, Xo ft M. Let Mo be the subspace spanned by M and Xo. Since Xo ft M, every y E M may be represented uniquely in the form y = m + axo, with m E M, a E 1F: Mo:= M + span {xo} = {y = m+axo: mE M, a E 1F}. Define I on Mo by (5.136) I(y) = ad, y = m + axo E Mo. (i) Clearly, I is a linear functional on Mo. Indeed if I(Yl) = ald and I(Y2) = a2 d , where Yl = ml + alXO, Y2 = m2 + a2 X O, then for each AE1F Yl + AY2 = (ml + Am2) + (al + A(2)XO so that for each A E 1F I(Yl + AY2) = (al + A(2)d = ald + A(a2 d ) = I(Yl) + A/(Y2)' (ii) If Y = m + axo E Mo and a # 0, then lIylI - lallixo - (-m/a)1I > laid (since -m/a E M) - I/(Y)I, by (5.136), so that Ilyll > I/(Y)I, for Y = m + axo. Therefore, IIIII < 1 (note that the last inequality trivially holds if a = 0, since Ilyll = IImll > 0 = If(y)l). On the other hand to prove IIIII > 1, we choose a sequence {m n } in M such that Ilxo - mnll  d, then d = I/(xo - mn)1 < 1I/IIIIxo - mnll  1I/IId so that IIIII > 1. Thus, IIIII = 1. 
5.14. Extension of Continuous Functionals 317 (iii) If y = Xo (m = 0, a = 1), then I(xo) = 1. d = d. If y = m E M (a = 0), then I(y) = O. d = 0 for each y E M and therefore, I vanishes on M. Finally, the theorem now follows by extending I on Mo to F on X such that IIFII = 11/11 = 1. . Theorem 5.135 is equivalent to 5.137. Theorem. Assume the hypotheses of Theorem 5.135. Then there exists a continuous linear functional F on X such that IIFI! = lid, F(M) = {OJ and F(xo) = 1. Proof. In the proof of Theorem 5.135, consider I(y) = a (instead of I(y) = ad) and proceed exactly as in the proof of Theorem 5.135. . Usually, there exist situations in which we shall have the occasion to apply the above results in the following form. 5.138. 'Corollary. Let M be a closed subspace of a normed space X and let Xo ft M. Then the conclusions of Theorems 5.135 and 5.137 hold. Proof. By hypothesis d= d(xo, M) > 0 and the hypotheses of Theorem 5.135 are met; and hence the conclusions follow. . 5.139. Corollary. Let M be a closed subspace of a normed space X and that M eX. Then there exists a nontrivial continuous linear ..... functional F such that F(M) = O. Proof. By hypothesis, M = M and M :j:. X. Therefore, there exists a point Xo E X\ M such that d(xo,M) > d(xo, M ) > o. The desired conclusion follows from Theorem 5.135. . For instance, let VI and V2 be two linearly independent vectors in a normed space X. Choose M = span {VI} and Xo = V2 in Corollary 5.138. Then we have the following simple consequence: there exists a continuous linear functional I such that I(VI) = 0 and I(V2) = 1. More generally, we have 5.140. Corollary. Let {VI, V2,.'., v n } be a set of linearly indepen- dent vectors in a normed space X. Then there exists a set {II, 12,.'.' In} of continuous linear functionals on X such that Ii (Vj) = 6 ij . In particular, each V E span {Vj : 1 < j < n} has the form n V = L Ij(v)vj. j=l 
318 Chapter 5: Linear Operators on Normed Spaces Proof. For each fixed i (i = 1,2,. . . , n), let M i = span {Vj': 1 < j < n, j  i}. Then, M i is finite dimensional with Vi ft M i and dim M i = n - 1 and therefore, M i is a closed subspace of X. By Corollary 5.138, there exists fi E X. such that fi(M i ) = 0, fi(Vi) = 1. and the first assertion follows as this is true for each such fixed i. Since each V in span {Vj : 1 < j < n} can be written as n V = E CtjVj for some Ctj E IF, j=1 it follows that n fi(V) = ECtjfi(Vj) = Cti j=1 which proves the second assertion. . 5.141. Corollary. If X is a finite dimensional normed space over F, then dim X = dim X. . Proof. Suppose that dim X = n and B = {V1, V2, . . . , v n } is a basis for X. Then, X = span{vj : 1 < j < n} and therefore, by Corollary 5.140, we have the following: (i) there exists a set B' = {f1, f2'...' fn} in X. such that fi(Vj) = 6 ij (ii) for each v E X, v = E j 1 fj(v)Vj. We claim that B' is a basis for the dual space X.. Now n n E!3jfj = 0 => E!3jfj(v) = O(v) = 0 for each v E X j=1 j=1 n => E/3jfj(Vi) = 0 for each Vi, 1 < i < n, j=1 n => E /3j6ji = 0 for each i, 1 < i < n (by (ii) ), j=1 => /3i = 0 for each i, 1 < i < n, so that B' is linearly independent set. Next, we show that B' spans X.. Let f E X. be arbitrary element. By (i), we have f(v) = f (  h(v)Vj ) 
5.14. Extension of Continuous Functionals 319 n - E/j(v)/(vj), since I is linear, j=l - (  f(Vj)/j) (v) which is true for each v E X and hence, it follows that n I = E/3jlj with /3j = I(vj). j=l Thus, each I E X. can be expressed as a linear combination of the elements of B'. Hence, B' is a basis for X.. . Corollary 5.141 apparently implies that every linear functional on the finite dimensional normed space X is bounded (see also Proposition 5.16). We recall that a closed subspace of a normed space need not be com- plemented (unless it is finite dimensional, see Corollary 5.142 ). However, it will be shown (see Corollary 6.82 and Theorem 6.95) that every closed subs pace M of a Hilbert space X is complemented by M.L: X = M EB M.L . 5.142. Corollary. Every finite dimensional subspace M of a normed space X is complemented in X. Proof. Suppose that dim M = nand B = {VI, V2, . . . , v n } is a basis for M. By Corollary 5.141, we have the following: (i) there exists a set {II, 12, . . . , In} in X. such that li( Vj) = 6 ij (ii) for each V E M, v = E j I/j(v)vj. Define N = n  IN!" where N/i = {x EX: Ii (X) = OJ. Then N is a closed subspace of X. We claim that . M n N = {OJ . X = M + N. Suppose that Z E M n N, i.e Z E M and zEN. Then zEN implies that z E N/, for each 1 < i < n, so that li(Z) = 0 for each 1 < i < n and z E M implies that n Z = E Ij(z)vj = O. j=l 
320 Chapter 5: Linear Operators on Normed Spaces Thus, M n N = {O}. Finally, for each x EX, we let n y=Efj(x)Vj andz=x-y. j=1 Clearly, y E M and x = y + z. We note that ZEN, since for each i fi(Z) fi(X - y) - fi(X) - fi(Y) - fi(X) - fi ( t h(X)Vj ) J=1 n - fi(X) - E fj(x)fi(Vj) j=1 n - fi(X) - E 6jifj(x) j=1 - fi(X) - fi(X) = 0 so that Z E N/i and hence zEN. . Alternately, if we define P : X  X by n Pv = Efj(v)vj for each v EX, j=1 then, by using (i) and (ii) of the proof of the last corollary, we can show that P is bounded, linear, Pv E M for each v E X, PVi = Vi, and Rp = M. Thus, M is the range of the continuous linear projection P on X. Hence, M is complemented in X (see Theorem 5.121). Our next result is to demonstrate how a linear functional allows us to decompose a given normed space. 5.143. Corollary. Let f : X  1F be a continuous linear functional on a normed space X. Then X = N/ E9 M, where M is a one dimensional subspace. Proof. Now N/ = {x EX: f(x) = OJ. As f is continuous, N/ is closed subspace of X. Assume that f # O. Then, by homogeneity of f, we may choose a point Xo # 0 such that f(xo) = 1. Now for each x E X, f(x - axo) - f(x) - af(xo) - f(x)(1 - f(xo)], if a = f(x), - O. 
5.15. Embedding of Normed Spaces 321 That is, z = x - axo E N/ whenever a = f(x). Thus, each x E X has a unique representation x=z+axo withzEN/anda=f(x) showing that X = N / €a span {xo}. . This corollary for Hilbert space setting is given in Corollary 7.7. Our final result shows that nonzero continuous linear functional always exists on every nontrivial normed space. 5.144. Corollary. Given a nonzero vector Xo in a normed space X, there exists a continuous linear functional F such that ) (5.145) F(xo) = IIxoll 'and IIFII = 1. In other words, if f(x) = 0 for all f E X* then x = O. In particular, X*, the set of all continuous linear functionals, separates the points of X. Proof. Choose M = {OJ, the trivial zero subspace of X and apply Theorem 5.135. To show that X* separates the points of X, let x,y E X with x :j:. y. Then Xo = x - y :j:. 0 so that Ilx - yll :j:. O. Therefore, by (5.145), there exists F such that F(x) - F(y) = F(x - y) = IIx - yll  0 and the conclusion follows. . 5.15 Embedding of Normed Spaces It is important to understand the notion of duality. First we recall that, if X is a normed space then, by Theorem 5.70(ii), the norm dual X* = B(X, F) of X is always a Banach space. Note that the algebraic dual of a vector space V is L(V, F), Le. the set consisting of all linear (need not be continuous) functionals on V. We wish to ascertain something about the dual space X* of a given normed space X. In this connection, it is therefore, natural to discuss the duals with the properties and structure of the successive spaces (X*)* := X**, (X**)*:= X***, and so on. In particular, the norm dual X** of X* is a Banach space and our special emphasis in the remaining part of this section is to look at the connection between the given space X and its second dual X**. Note that X** consists of all continuous linear functionals defined on X*. Let us start with certain preliminaries and try to describe the notion of duality in a simple way. For 
322 Chapter 5: Linear Operators in Normed Spaces the sake of convenience, let the elements of X, X* and X** be x, y, . . ., be f, g, · · ., and be F, G, . . ., respectively. Consider the following bilinear mapping (often referred as a pairing between X and its dual X*) via the formula X* x X  IF, (f, x) I-t f(x). This mapping formula illustrates also that, just as a linear functional f determines a map f : X  F, each vector x E X determines a functional X*  IF which is an element of X**. To see this, fix x E X and let f to vary over X*. Then for different f E X*, we obtain various values for f(x) and thus, every vector x E X gives rise to certain functionals Fx on X* via the formula Fx(f) = f(x) for every f E X*. At this place, it is appropriate to remember one of the basic results in linear algebra which says that "X is isomorphic to X** if X is finite dimensional". We shall soon see by examples that this result does not necessarily hold in a general normed space. However, as we shall see in Chapter 7, the situation is much better if X is a Hilbert space, instead of a normed space. We now show that Fx is a bounded linear functional on X*, where X is a normed space. Indeed, for f, 9 E X* and a E IF, Fx(af + g) = (af + g)(x) = af(x) + g(x) = aFx(f) + Fx(g), Le. the functional Fx is linear. Further, for every x EX, the inequality IFx(f)1 = If(x)1 < Ilfllllxll implies that IIFx II < lIxll and Fx E X**. This clearly defines a mapping J : X  X** , x I-t Fx. We show that J is a linear mapping. Now, for x, y E X and a E IF, we have Fax+y(f) = f(ax + y) = af(x) + f(y) = (aFx + Fy)(f) for every f E X*, so that F ax + y = aFx + Fy. That is, J(ax + y) = F ax + y = aFx + Fy = aJ(x) + J(y) and therefore, J is linear. We have already shown that IIFxll < IIxli. Next, we aim to prove the reverse inequality IIFx II > IIxll. 
5.15. Embedding of Normed Spaces 323 Figure 5.3: Natural embedding For x = 0, this is obvious. For x :j:. 0, by virtue of Corollary 5.144, there exists a continuous linear functional g E X* such that g(x) = IIxll and lIyll = 1. Using this, we find that { IFx(f)1 * } g(x) IlFz II = sup 11/11 : I EX, 11/11 =I- 0 > li9iI = IIxll and thus, IIFxll = Ilxll for every x E X which implies that J is an isometry. Since an isometry is one-to-one, J is an isomorphism. Finally, to say that X and X** are isomorphic, we need to show that J is onto (which is not true, in general, as we shall soon see by an example). Now, we make the following definition: "The mapping x I-t Fx is called the natural embedding of X into its second dual x**" , see Figures 5.3. In conclusion, we have proved the following important result (compare with Theorem 7.12). 5.146. Theorem. The natural embedding x I-t Fx from a normed space X into its second dual X** is a norm preserving linear injective map (and hence, X can be identified as a subspace of X**). In other words, the natural embedding x I-t Fx is a linear isometry. If the mapping J : X  X** is onto (i.e. J(X) = X** or we simply write X = X**), then the space X is called reflexive (such a space X is necessarily complete!). In general, the map x I-t Fx is not onto and hence, X (when embedded in X**) is a proper subspace of X**. To provide examples of dual spaces, we use the following notation. For a E C, let sgna = { :I for a = 0 for a:j:. O. 
324 Chapter 5: Linear Operators in Normed Spaces Clearly, Isgnal = {  for a = 0 for a:j:. 0 { o s n a = g e-iArg 0: for a = 0 for a:j:. 0 and { 0 for a = 0 a sgn a = I AI I  for a:j:. o. Our first result concerns the spaces dual to lP for 1 < p < 00. 5.147. 1 1 Theorem. For 1 < p < 00 and - + - = 1 (q = 00 when p q p = 1 ), we have (lP)* = ,q in the sense that there exists a linear isometry (i.e. isometric isomorphism) T which maps (IP)* onto lq. In particular, 1 2 is self-dual. Proof. Using Schauder basis {ek}kl, each element Z = {Zk}kl E lP has a unique representation 00 Z = LZkek, k=l since, if Sn = E=l Zkek = (Zl, Z2, . . . , Zn, 0, 0, . . .), we have IIz-snllp = 1I(0,0,...,Zn+l,Zn+2,...)lIp ( 00 ) IIp k  l1ZklP  0 as n  00. Now, let f E (IP)* = B(lP, IF) be arbitrary. Then, since f is linear, we have n f(sn) = L zkf(ek). k=l Moreover, since Sn  Z as n  00 and f is continuous, f(sn)  f(z) as n  00 so that each f E (IP)* is represented in the form (5.148) 00 f(z) = L zkf(ek), for each Z = {Zk}kl E lP. k=l 
5.15. Embedding of Normed Spaces 325 We claim that w = {/(ek)} E lq. Now we break the proof into two cases: Case (i) : Let p = 1 and q = 00. For convenience, we let ak = I(ek) and show that w = {ak}kl E 1 00 . We have IIwll oo = sup lakl = sup I/(ek)1 < 1I/IIII e kill = I1III < 00. k k On the other hand, (5.149) 00 00 I/(z)1 < E IZkakl < IIwll oo E IZkl = IIwll oo Ilzlll k=l k=l which implies that 11/11 < IIwll oo . Thus, 11/11 = Ilwll oo and therefore, the map T: (1 1 )*  1 00 , I t--+ W =-{ ak}kl, is an isometry- and is obviously linear. Next, we show that T defines an isomorphism of (1 1 )* onto 1 00 . We know that an isometry is one-to-one and therefore, it suffices to show that T is onto. To do this, we choose an arbitrary element /3 = {/3k} E 1 00 and define a linear functional 1/3 : 11  IF by 00 1/3(Z) = E zk/3k o k=l As in (5.149), we obtain 1//3(z)1 < 11/31100 IIzlll which shows that 11//311 < 11/31100 < 00, Le. 1/3 E (1 1 )* and 1/3(ek) = /3k for each k, Le. T(I/3) = (3. Thus, T is onto. Case (ii) : Let 1 < p < 00 and ! + ! = 1. As in Case (i), we assume that p q I E (IP)* and ak = I(ek). First, we show that w = {ak}kl E lq. To do this, for any fixed n, we consider the following element Zn E lP: Zn = (z,z... ,z,O,O,...) where lakl q z = laklq-1sgn ( a k) = - ak 
326 Chapter 5: Linear Operators in Normed Spaces so that Iz1 = lakl q - l . Thus, (5.150) n n n IIZnll = E IzIP = E laklp(q-l) = E lakl q k=l k=l k=l and n I(Zn) - E zak k=l n - E lakl q k=l < 1I/IIIIZnllp ( n ) l/p - IIfll  lakl q , by (5.150). Dividing through by the last term on the right, we find that (  la k1q ) l-l/p < IIfll which is equivalent to ( n ) l/q Sn =  lakl q < IIfll. This is true for each n and {8n} is seen to be a bounded monotone increasing sequence and theefore, converges. Hence, IIwll q < 11/11 < 00. In particular, w E lq. At the same time, by virtue of the Holder inequality, (5.148) shows that I/(z)1 < IIzllp IIwll q so that 11/11 < IIwll q and therefore, we obtain that 11/11 = IIwll q . Thus, the map T: (lP)*  lq, I I-t w = {ak}kl' is a continuous linear map which is an isometry. Clearly, T is one-to- one, since T is an isometry. As in the previous case, Holder's inequality immediately shows that T is onto. _ 5.151. Corollary. Let {an}n>l be a sequence of complex numbers and p > 1. Suppose that the series E  1 anZ n converges for each z = {Zn}nl E lP. Then we have 
5.15. Embedding of Normed Spaces 327 . {an}nl E 1 00 ifp = 1 . {an}nl E Iq ifp > 1 and q = p/(p - 1). Proof. Define In : lP  IF by n In(z) = E akZk for n = 1,2, · . .. k=l Then, by Theorem 5.147, In E (IP)* and maxlakl ifp=l (5.152) IIfnll = l ( $;k n $;n ) l/q  lakl q if p > 1. By hypothesis, since the series E  1 anZ n converges for each Z E lP, the family {/n(z) : n = 1,2,...} is bounded for each Z E lP. The Uniform Boundedness Principle implies that the sequence {In} is bounded and the desired conclusion follows by (5.152). . 5.153. Examples of reflexive and nonreflexive spaces. (i) Every finite dimensional Banach space is reflexive. (ii) For each 1 < p < 00, the space lP is reflexive. Indeed, by Theorem 5.147, for each 1 < p, q < 00 with 1 + 1 = 1, we have p q (IP)* = ,q and because of symmetry (lq)* = lP so that, by combining these two, (IP)** = lP for 1 < p < 00. (iii) Each of the spaces Co, 11 and 1 00 is not reflexive. Let us first prove that Co is not reflexive and the proof for the remaining spaces may be supplied with similar arguments. Recall that  = 11 and (1 1 )* = 1 00 and therefore, C ** _ ' 00 0-' Thus, we can identify Co with 1 00 and the canonical embedding J : Co  ' 00 is precisely the inclusion Co C 1 00 , as a closed subspace. But we know that this inclusion is proper and therefore, J is not onto. So, Co cannot be- reflexive. '. 
328 Chapter 5: Linear Operators in Normed Spaces 5.154. Theorem. If X is a normed space such that X* is separable, then X is also a separable space. Proof. Suppose that X* is separable. Then the unit sphere Sx. must contain a countable dense subset, say, S = {fn E X* : IIfnll = 1 for all n EN}. Note that for each n, IIfnll = sup Ifn(x)1 IIx 11=1 and therefore, for each n, there exists X n E X with IIxn II = 1 and Ifn(xn)1 > 1/2 (otherwise, this would contrad ict the fact that Il fnll = 1). Now, let M = span {x n : n E N}. Then M is a closed subspace of X. Now, we claim that M = X. Suppose not. Then there e:?Cists a point Xo E X \M and, by Corollary 5.138, there exists a continuous linear functional F such that IIFII = 1, F(M) = 0 and F(xo) # O. In particular, since X n E M, we have F(xn) = 0 for each n E N. Let n be such that IIfn - FII < 1/4 which is clearly possible since S is a dense subset of Sx.. Now, we find that IF(xn)1 - > > > Ifn(x n ) - (fn(xn) - Fn(xn))1 Ifn(xn)1 - Ifn(x n ) - Fn(xn)1 Ifn(xn)1 - Ilfn - Fnllllxnll 1/2 - 1/4 = 1/4, which is a contradiction, and we complete the proof. . The converse of Theorem 5.154 is false. For instance, consider X = 11 which is known to be separable. On the other hand, we also know that X* = (11) * = 1 00 , which is a nonseparable space. The space C[O, 1] is not reflexive. Consider the evaluation functionals 6t : f I-t f(t) which implies that 6t E C[O, 1]* for every t E [0, 1]. Now for every s # t, we have 116 8 - 6tll = 2 
5.15. Embedding of Normed Spaces 329 T 1F I f  Y f  f* f f 0 -.-. x Figure 5.4: Adjoint of an operator which shows that C[O, 1]* is nonseparable. Thus, if C[O, 1] = C[O, 1]** were true, then, by Theorem 5.154, C[O, 1]* would be separable and we arrive at a contradiction. Hence, the space C[O, 1] is not reflexive. 5.155. Adjoint of bounded linear operators. Let X and Y be two normed spaces and T : X  Y be an operator. Then we have a natural map T* : Y*  X* defined by T* f = f 0 T or T* f(x) = f(Tx) for all f E Y* and x E X. The mapping T* : Y*  X* so defined is called the adjoint (or conjugate) of T, see Figure 5.4. We have the following simple properties. 5.156. Proposition. Let X, Y and Z be three normed spaces and T E B(X, Y). Then we have (i) T* E B(Y*, X*) and IIT*IIB(Y*,X*} = IITIIB(X,y) (ii) The mapping T I-t T* (via B(X, Y) I-t B(Y*, X*) ) is linear, injective and an isometry. (Hi) 1fT E B(X, Y) and U E B(Y, Z), then (UT)* = T*U.. Proof. To prove the linearity of T, consider f, 9 E Y* and a, {3 E IF. Since T*(af+{3g)(x) = (af+{3g)(Tx) = af(Tx) + {3g(Tx) = aT* f(x) + {3T*g(x) for all x EX, we see that T*(af + {3g) = aT* f + {3T*g. So, T* is linear. Now, IT* f(x)1 = If(Tx)1 < IIflillTllllxll 
330 Chapter 5: Linear Operators in Normed Spaces and therefore, IT* fl < IIflillTl1 which shows that the operator T* is bounded and has a norm at most IITII. Also, by an application of the Hahn-Banach theorem (see Corollary 5.144), given a nonzero vector y E Y, there exists a continuous linear functional 9 E Y* such that Ig(y)1 = lIyll and IIgll = 1. Applying this with y = Tx (where x E X is arbitrary) gives IITxll = Ig(Tx)1 = IT*g(x)1 < IIT*lIlIgllllxll = IIT*lIlIxll which implies that IITII < IIT*II. Hence, IITII = IIT*II. (ii) For all x EX, we have (aT + (3S)* f(x) - f((aT + (3S)(x)) - f((aTx + /3Sx)) - af(Tx) + /3f(Sx) - aT* f(x) + /3S* f(x) - (aT* + (3S*)f(x) which proves the linearity of the mapping * : T I-t T*. We have shown in (i) that * is indeed an isometry and * is an injective map because T* = S* :::} IIT* - S*II = 0 => II(T - S)*II = 0 :::} liT - SII = 0 :::} T = S. (iii) The last part is easy to verify. Indeed, (UT)* f(x) - f((UT)(x)) - f(U(Tx)) - (U* f)(Tx) - (T*(U* f))(x) - (T*U*)f(x) so that (UT)* = T*U*. . Recall that to each x EX, we can correspond to a functional Fx E X** via the formula Fx(f) = f(x) for every f E X*. Write Fx = J(x) and call J : x I-t Fx the natural embedding of X into X**. 
5.15. Embedding of Normed Spaces 331 We recall the definition of an extension of an operator. Let T : DT C X  Y be a operator between the two vector spaces X and Y with D T as its domain. An operator S : X  Y is an extension of T if D T c Ds and Tx = Sx for all x EDT. If we let XE and YE are the images under the natural embedding of X, Y into X** and Y** respectively, and if T E B(X, Y), then we define TE E B(XE, YE) by the relation TEx e = Ye (xe E XE, Ye EYE) where Y = Tx. If S : Ds C X**  Y** such that XE C Ds and SX e = TEx e for all Xe E X E , then we call S (which is an extension of TE) an extension of T. If Ds = XE, then we write S = T. 5.157. Theorem. Let X, Y be two normed spaces and T E B(X, Y). Then the second adjoint T** := (T*)* : X**  Y** is an extension ofT. If X is reflexive, then T** = T. Proof. By Proposition 5.156(i), IITIIB(X,y) = IIT*IIB(Y.,X.) = IIT**IIB(X..,y..). Let J : x I-t Fx be a natural embedding of X into X** so that Fx(f) = f(x) and Fx = J(x). For x E X and 9 E Y*, we have T** Fx(g) - Fx(T*g), by the definition of adjoint, - (T*g)(x), by the definition of embedding, - g(Tx), by the definition of adjoint, - FTx(g) which shows that T** Fx = F Tx , or T** J(x) = J(Tx), Le. T** 0 J = JoT and therefore, T** is a norm preserving extension of T. If X is reflexive, then X = X** and so T = T**. . 5.158. Theorem. Let X be a Banach space, Y, a normed space and T E B (X, Y). 1fT has a bounded inverse T-l (with domain Y), then T* has a bounded inverse (T*)-l (with domain X*). Moreover, (T*)-l = (T- 1 )*. Proof. Suppose that T- 1 exists. Then, by Proposition 5.156(iii), Iy. = (TT- 1 )* = (T-1)*T* and Ix. = (T-1T)* = T*(T- 1 )* 
332 Chapter,5: Linear Operators on Normed Spaces where Ix. and Iy. are the identity operators in B(X*) and B(Y*), re- spectively. This proves that the inverse (T*)-l exists and it is given by (T*)-l = (T-l)*. . The converse of Theorem 5.158 also holds but we leave it as an exercise. It is indeed easier to prove the converse part in the case Y = X and it is a Banach space. 5.16 Exercises 5.159. Determine whether the following statements are true or false. Justify your answer. (a) Let 11.111 and 11.112 be two norms on X such that Ilxlli < Cllxll2 for each x E X. If a set A is dense in B with respect to 11.112, then it is dense with respect to II · 111. (b) The map T : C[O, 1]  JR, I(t) I-t 1(1), is continuous for just one of the metrics d l and doo on C[O, 1]. (c) Let X denote the set of all functions I in C[O,I] such that I'(x) exists and continuous on [0,1] with the supnorm. Then T : X  JR, I(x) I-t 1'(0) is linear but not continuous. (d) IfT:}R2 }R2 is defined by (x,y)  (xcos8-ysin8,xsin8+ycos8) with the Euclidean norm, then T is a bounded linear transformation with IITII = 1. (e) Let X = R[a, b] be the set of all Riemann integrable functions on [a, b] with the supnorm and Y = }R with the Euclidean norm. Then T : X  Y, I I-t J: I(t) dt, is continuous on X. (f) Let X be the vector space of all real valued functions on [0, 1] that have continuous first derivatives with the supnorm and Y = ]R with the Euclidean norm. Then T : X  Y, I(x) I-t 1'(0), is an unbounded operator. (g) Let X be the vector space of all real valued functions on [0, 1] that have continuous first derivatives with L 2 -norm and Y = C[O, 1] with L 2 -norm. Then T : X  Y, I I-t I', is an unbounded operator. (h) Let T : X  Y be a bijective linear operator between the Banach spaces X and Y. Then T-l is continuous iff m = inf{llTyll : Ilyll = I} > O. (i) If Y is a closed convex subset of a Banach space X, and if T : Y  Y is a nonexpansive map, then inf xEY IIx - Txll = o. (j) Every infinite dimensional Banach space X admits a discontinuous linear functional on X. 
5.16. Exercises 333 (k) There exists a discontinuous linear map T of a Banach space X into itself such that T-1 (0) is closed. (I) If X is a vector space and I,g E X* are such that N/ = N g , then 1 = Ag for some scalar A :j:. o. (m) A linear transformation between two normed spaces need not be con- tinuous. (n) If X = C[O, 1] equipped with L 2 -norm and if k(x, y) is continuous on the unit square [0, 1] x [0, 1], then the operator T defined in Example 5.79 is bounded. (0) For t E [0,1], the subset t 2 t n A = {In(t) = t + 2 + · · · + n : n E N} C PR[O, 1], where PR[O, 1] is the linear space of all real polynomials p(t) defined on [0, 1], is unbounded with respect to the supnorm II · 1100 but not w.r.t the L1- norm . The same statement holds for the subset A = {In(t) = nt(l - t)n : n E N} C PR[O, 1] (see also Example 5.81). (p) Let Y be the normed space of all the complex sequences {zn} such that L:  1 n!zn is convergent and the norm is given by II{zn}1I = L:  1 n!lznl. Define T .: , 1  Y by {zn}  {znjn!}. Then T is a continuous operator. (q) If X = (IR2,1I'1I00) and Y = {y = (Y1,Y2) EX: lIylloo < I}, then there are infinitely many points in Y nearest to x = (2,0) E X. (r) If Y is a finite dimensional proper subspace of a normed space X and if x EX, then there is a best approximation to x by the elements of Y. (s) The linear operator T : 1 2  1 2 defined by the formula Z = {zn} n 1 I-t W = {Wn}n1 such that Tz = w iff W n = Zn - 3z n + 1 v'iO 3z n - 1 + Zn v'iO if n is odd, if n is even, is an isometry. (t) If I(x) = 0 for every continuous linear functional 1 defined on normed space X, then x = O. (u) If {V1, V2, . . . , v n } is a set of linearly independent vectors in a normed space X, then, for any set a1, a2, . . . , an of real numbers, there exists a continuous linear functional 1 on X such that I(Vk) = ak for all k=1,2,...,n. 1 1 (v) For 1 < p,q < 00 and - + - = 1, we have (IP(n))* = lQ(n). p q 
334 Chapter 5: Linear Operators on Normed Spaces (w) (l1(n))* = loo(n) and (loo(n))* = 11(n). (x) (l1)* = 1 00 , but (1 00 )* :F l1. (y) The strict inclusions Coo S; lP S; Co S; c holds for all 0 < p < 00, where c, Co and Coo are the usual subs paces of lOO . (z) If X is a Banach space and Y, Z are two closed subspaces of X such that Y n Z = {OJ, then Y + Z is closed iff there exists a positive real number M > 0 such that Ilyll < Mlly + zll for all y E Y and z E Z. 5.160. Prove or disprov the following statements: (a) For 1 < p < 00, the spaces lP(n) and lP are strictly convex. (b) The spaces 11(n), 1 1 , loo(n), lOO and (C[a,b], 11.11(0) are not strictly convex. 5.161. If X = (]R2, 11.112) and Y = {(x,y) EX: x = 3y}, then find the norm of a general element x + Y in the quotient space X/Yo 5.162. Let Ck[a, b], k > 1, denote the space of all k-times continuously differentiable functions f(t) defined on the interval [a, b]. (Then Ck[a, b] is an incomplete normed space under the supnorm, see Exercise 3.73(ii) for k = 1.) Show that Ck[a, b] is a Banach space under the norm k IIfll = L IIf(m)lloo, where IIf(m) 1100 = max If(m)(t)l, f(O)(t) = f(t). m=O tE[a,b] Is the map T : (Ck[a, b], 11.11(0)  (Ck[a, b], 11.11) continuous? Note: Notice that the CO-norm is the supnorm and that then CO [a, b] is what we would have written as C[a, b] earlier. Recall that, f E C1[a, b] iff f is differentiable on (a, b), differentiable from the right at a and the left at b, and the derivative f' is continuously differentiable. Here f'(a) and f'(b) are the right hand and the left hand derivatives at a and b, respectively. Let us now answer the question for the case k = 1. Consider a sequence {fn(t)} of functions, where fn(t) = (l/n) sin n1rt. Then (with k = 1) sin n 1rt 1 1 1 IIfnlloo = sup = - and IIfnll = - + sup I cosn1rtl = 1 + -. tE[O,1] n n n tE[O,1] n This shows that {fn} converges to the limit f(t) = 0 in (C1 [0, 1], II · 11(0) but not in (C1[0,1],II'II). Thus, T: (C1[0,1],II'1I00)  (C1[0,1],II'11) is not continuous. 5.163. Let X = (C1[0, 1], II · II), where IIfll = IIflloo + IIf'lIoo, and Y = (C[O, 1], 11.11(0)' Define T : X  Y by (Tf)(t) = a(t)f'(t) + b(t)f(t), 
5.16. Exercises '335 where a and b are fixed members of Y. Show that T is continuous on X. 5.164. Define f : IR  C and (Tf)(x) = f(x)f(1 + IxDQ, a: > O. For which p > 1, is T : LP{IR)  L1 (IR) a bounded linear operator? Give an estimate for its norm. Can you find the norm? 5.165. Let X = (C[a, b], 11.11(0)' Show that each of the linear functional defined by 1 t-+ i b I(t) dt and n f I-t E akf(xk) (ak E IR, Xk E [a, b]) k=1 is continuous on X. 5.166. Define T : , 2  , 2 by (i) {Zn}n1 I-t {Z2n}n1 = {Z2, Z4, . . .} (ii) {Zn}nl t-+ { nl Zn Ll = Hz!' i Z 2'...} (Hi) {Zn}n1 I-t { n1 Zn+1} n1 = {2z 2 , Z3' . . .} (iv) {Zn}n1 I-t {zn + Zn+1}n1 = {Z1 + Z2, Z2 + Z3,...} (v) {Zn}n1 I-t {Wn}n1' where Z2n if n is odd, W n = n Z2n-1 if n is even. n Find IITII in each case. 5.167. For X = (C[O, 1], 11.11(0), let T : X  X be defined by (Tf)(t) = I t I(s) ds, t E [0,1]. Prove that IITnll < Ifn!. Is IITnll = IITlln? 5.168. Fix a E (0, 1) and consider the mapping T : C[O, 1]  IR by f(t) I-t f(a). Find metrics d and p on C[O,I] such that T is continuous with respect to d but not with respect to p. 5.169. Construct examples for an operator T to be (i) closed but not continuous 
336 Chapter 5: Linear Operators on Normed Spaces (ii) closed, where T from a Banach space into a normed space that is not continuous (iii) closed, where T from a normed space into a Banach space that is not open. 5.170. Check which of the following mappings are open? (i) T: }R3  ]R2, X = (Xl, X2, X3) I-t (Xl, X3) (ii) T: }R3  JR3 , X = (Xl, X2, X3) I-t (Xl, X2, 0). 5.171. In 1 2 , consider the following mappings: (i) {Zn}nl I-t {nZn}nl = {Zl, 2z 2 ,...} (ii) {Zn}nl I-t {nZn+l}nl = {Z2, 2Z3,...} (iii) {Zn}nl I-t {2-(n-l)(Zl + Z2 +... + Zn)} nl = {Zl, (Zl + z2)/2,.. .}. Show that the first two mappings are closed while the last one is bounded but not onto. 5.172. IT X is a normed space, c E X and A E IF, then show that each of the mappings X I-t X + c, X I-t AX is a homeomorphism. 5.173. Consider the space X = en with the norm IIzll = E  1 aklzAd for Z = (Zl, Z2, . . . , zn) E en, where {al, a2, . . . , an} is a fixed set of non- negative real numbers. Construct Y and X/Y in this case. 5.174. If X is a normed space and Y is a closed subspace such that both Y and X/Yare Banach spaces, then X is a Banach space. 5.175. Let X = Cl[a,b] with the norm 11/11 = maxtE[a,b] 1/'(t)l. Con- struct Y and X/Y in this case. 5.176. Let X be a closed subspace and Y, a finite dimensional subspace of a normed space. Then X + Y = {x + y : X E X, Y E Y} is closed. 5.177. Show by an example that the Banach isomorphism theorem (see Theorem 5.107) fails for incomplete normed spaces. Note: Consider the space X of all polynomials over III For n p(x) = L ak xk E X, k=O 
5.16. Exercises 337 let IIpll = maxokn lakl. Define T : X  X by n Tp(x) = ao+ L o; x k . k=l Show that X is an incomplete normed space, T E B(X), T-l exists but T-l is not bounded. 5.178. Show by an example that the completeness assumption of the domain of {Ta} in the Uniform Boundedness Principle (see Theorem 5.122) cannot be dropped. 5.179. Discuss the conjugates of LP-spaces and the space C[O, 1]. 
Part III HILBERT SPACES In the earlier parts, we have considered different kinds of topological spaces, namely metric spaces, normed spaces and Banach spaces. In this part, we shall first present the elementary properties from the theory of Hilbert spaces 26 , a well-developed area in the theory of inner product spaces. In the real case, an inner product is a symmetric bilinear form and positive definite on V x V  III In the complex case, it is a Hermitian symmetric, conjugate bilinear and positive definite on V x V  C. A (semi) inner product gives rise to a (semi) norm. Thus, an inner product space is a particular case of a normed space. Our particular emphasis in Chapter 6 will be on orthogonal projections and orthonormal bases. Unless otherwise is stated explicitly, all the Hilbert spaces mentioned are assumed to be complex Hilbert spaces whose definition is simple to state: a complete inner product space with a norm that is derived from the inner product; Le. Hilbert spaces are special Banach spaces. Further, because of the inner product structure, Hilbert spaces have extra features than Banach spaces, which make them more special and hence lead to several interesting properties. For example, some ideas for use in differential equations like Fourier Analysis do not fit well in Banach space setting but fits very well in Hilbert space setting. Chapter 6 begins with the definition of inner product and discusses sev- eral properties and applications of Hilbert spaces since the Hilbert space theory is very important in applied mathematics and, in particular in the study of quantum mechanics. In Proposition 6.13, we show how an inner product space on a real or a complex vector space can be made into a normed space with the help of a very important inequality (6.14) due to Cauchy-Schwarz-Buniakowski (briefly we use the notation CSB27 to refer this inequality), and probably, many others. Then we proceed to give ex- amples showing that a normed space need not be an inner product space, see for instance 6.33. The concepts such as continuity, convergence and completeness presented in Chapter 2 are valid in the inner product spaces 26Hilbert spaces were developed by the German mathematician David Hilbert (1862- 1943) and his school but not in an abstract fashion. In 1927, Yon Neuman gave the axioms for a Hilbert space which is named after David Hilbert. 27The CSB inequality was first proved for finite dimensional sums by Cauchy in 1821 (see Corollary 6.42), for integrals by Buniakowski (see Example 6.43) and rediscovered by Schwarz in 1885. 
340 because the family of all inner product spaces forms a proper subset of the family of all normed spaces. The polarization identity derived in Proposi- tion 6.13 expresses the norm of an inner product space in terms of the inner product. Indeed for the real inner product spaces (see Theorem 6.44), we have 1 (u, v) = 4 [IIu + vll 2 - lIu - v1l 2 ] and for complex inner product spaces one has (u, v) =  [(lIu + vll 2 - lIu - v1l 2 ) + i(lIu + ivll 2 - lIu - i v Il 2 )]. Among the standard examples of Hilbert spaces, there is the Euclidean space en with an inner product n (z,w) = LZk W k (z = (Zl,Z2"."Zn), W = (Wl,W2,...,W n )), k=l the space 1 2 of square-summable sequences indexed by N with an inner product 00 ({Zk}, {Wk}} = L Zn W n, n=l and the space L2[a, b] of all square-summable functions f : [a, b]  IF with an inner product (f, g) = i b f(t) g(t) dt. The most important among the features of Hilbert spaces is its underlying concept of orthogonality which is useful to introduce the notion of orthonor- mal basis (= maximal orthonormal set) (see Sections 6.5,6.7 and 6.6). We warn the readers that an orthonormal basis is not a Hamel basis in the usual sense, unless the underlying space is finite dimensional. Section 6.7 covers an important result-called the "projection theorem" (see Theorem 6.74); thus, in a Hilbert space X, it makes sense to talk about a point Xo in a closed convex set K of X that is closest to a given point x E X. In this section, we also use the orthogonality concept to split a Hilbert space up into a 'sum' of subs paces and indeed, we show that every closed subspace K of a Hilbert space X is complemented by Kl., Le. X = K E9K .1.. (see Corollary 6.82). However, in Examples 6.90 and 6.91, we point out that a closed subspace of a Banach space is not necessarily complemented unless the subspace is finite dimensional (see Corollary 5.142). In Section 6.8, we prove an interesting result (see Theorem 6.121) stating that "every separable Hilbert space X is isomorphic either to the 'Unitary space en for some n or to the space l2". In Section 6.10, we discuss certain basic concepts from Fourier Analysis and give an application of uniform boundedness principle. 
341 In Chapter 7, we discuss an important result known as the Riesz repre- sentation theorem which shows that there is a one-to-one correspondence between the elements of a Hilbert space X and the members of its dual X. (see Theorem 7.12). An important corollary to this fact is that each Hilbert space X and its dual X. are isometrically isomorphic to each other. More precisely, this result implies that the examples of bounded linear func- tionals f on a Hilbert space X determined by a vector p E X such that f(x) = (x,p) for each x E X, are indeed all bounded linear functionals on X. Finally, we discuss the notion of adjoint operator as a consequence of the Riesz representation theorem. A.t the end, we present several examples of adjoint operator along with some applications. 
Chapter 6 Inner Product Spaces In this chapter, we first introduce the basic concept of an inner product space and then proceed to discuss certain properties and examples of inner product spaces. In Section 6.1, we also consider Hilbert spaces and exam- ples of spaces that are not Hilbert spaces. Thanks to the inner product structure of a Hilbert space, we discuss the concept of orthogonality in Sec- tion 6.5. In Section 6.7, we present the deep property called the Projection theorems on Hilbert spaces that follows from the concept of orthogonal projections in Hilbert spaces. 6.1 Inner Product We begin with 6.1.. Definition. Let V be a linear/vector space over a field IF (where IF is either C or IR). By an inner product on V, we mean a mapping 28 f := (.,.) : V X V  F, (u, v) I-t (u, v) = f(u, v), that assigns for each (u,v) E V x V a value in F, denoted by (u,v), called a scalar valued expression (or simply the inner product of u and v), such that for each u, v, w E V and A E F we have [Hermitian/Conjugate symmetric] [Homogeneous] [Additivity] [Positivity] (11) (u, v) = (v, u) (12) (Au, v) = A(u, v) (13) (u, v + w) = (u, v) + (u, w) (14) (u, u) > 0, and (u, u) = 0 => u = o. In (11), the bar denotes the complex conjugation. 28S ome authors use the notation (.,.) or (.1.) to denote the inner product (., .). 
344 Chapter 6: Inner Product Spaces There are number of observations that can be made about these axioms. The fact that (u, u) must be real follows from (11) on taking v = u n (11). Further, (12) for A = 0 shows that (0, v) = 0 for every v E V, and in particular, we have u = 0 =>  u) O. We shall see in details how the function II . II defined by lIull = (u, u) makes V into a normed space. A vector space V together with its inner product (.,.), Le. the pair (V, (., .) ),29 is said to be an inner product space. An inner product space is also called a pre-Hilbert space. As in the case of normed spaces, there are really two definitions here and these depend on V over the real field and V over complex field, respectively. Thus, when the underlying field is JR, then V is called a real inner product space; otherwise it is called a complex inner product space. In the real case the axioms are the same, except that, (11) will be written without the bar over (v, u) since complex conjugation has no effect: (u, v) = (v, u). With this understanding, Definition 6.1 covers both real and complex cases. When (11) is combined with (13) and (12), we have (u + v, w) = (u, w) + (v, w), (u, AV) = A (U, v). Thus, the inner product is linear in the first variable, and conjugate linear in the second variable. As a consequence of this, we easily get that (6.2) (AU, J.tv) = A J.t (U, v) for A, J.t E F and in the sequel, we shall use this formula often. In fact, we deduce the fol- lowing general properties of the inner product which are indeed immediate consequences of the axioms (11)-(14) of Definition 6.1, and can be verified easily. If Ov denotes the zero vector in V and 0 denotes the scalar zero in F, then, for any vector v E V, we have 0 · v = Ov so that by the linearity property of the inner product (Ov, v) = (0 · v, v) = O(v, v) = O. Given a vector space V over F, a semi-inner product for V is a function on V x V which satisfies all the axioms of the inner product except (14) and instead of (14), satisfies just the condition (u, u) > O. Thus, a semi-inner product allows the possibility that for some u  0, (u, u) = O. For example, let V be the set of all sequences u = {Un}nl 29When there is no confusion, very often we refer the inner product space (V, (., .) ) simply by V itself with an understanding that there is an associated inner product (.,.) with respect to the vector space V. 
6.1. Inner Product 345 such that Un = 0 for all, but a finite number of values of, n. For U and v = {Vn}nl in this space V, define 00 (u,v) = LU2k V 2k. k=l Then, (.,.) defines a semi-inner product but is not an inner product. 6.3. Proposition. In an inner product space (V, (., .)), we have for each u,v,w E V and Al,A2 E]F (i) (0, u) = (u,O) = 0 (H) (Al U + A2V, w) = Al (u, w) + A2 (v, w) [Linearity] (iii) (U,AlV+A2W) = A l(U,V)+ A 2(U,W) [Conjugate linearity] (By (H) and (Hi), (.,.) is a conjugate bilinear function.) (iv) More generally, for all Uj, Vk E V and Aj, J.tk E ]F (j = 1,2, . . . , n; k = 1, 2, . . . , m), we have (6.4) (  )..jUj,  J.tkVk) =  )..j J.tk (Uj'Vk}' The bar over Al, A2 in (Hi) and also the bar over J.tk in (6.4) are to be omitted in the case of a real space as it has no effect. From Definition 6.1 and Proposition 6.3, we say that the key properties that characterize the inner product are 'Hermitian symmetric, conjugate bilinear and positive definite '. Clearly, for every w E V, (U, w) = (v, w) holds iff (u - v, w) = 0, which for the choice w = u - v E V gives (u - v, U - v) = 0, Le. U = v. The simplest example of an inner product space is C itself, with (u, v) = u v . Let us look at a simple generalization of this inner product space. 6.5. Standard inner product on r. Consider the unitary space en, i.e. the set of all n-tuples of complex numbers, see Example 2.32. For U = (Ul,U2,... ,un) and v = (Vl,V2,... ,v n ) in en, where the Uk and Vk are complex numbers, we define the complex valued function (.,.) on the space en x en by (6.6) n (u,v) = LUk Vk = (Ul U2 ... un) k=l V l V 2 V n 
346 Chapter 6: Inner Product Spaces (for vectors in }Rn the bar over Vk, denoting the complex conjugate, has no effect). Then it is easy to verify that the space (en, (', .)) is an inner product space and is often called standard inner product space on en. For an example of nonstandard inner product, see Exercises 6.135(c). We re- mark that in the real Euclidean space }Rn, the inner product defined above becomes (6.7) n (u, v) = U · v = E UkVk := (Ul U2 · .. un) k=l Vl V2 V n where the Uk and Vk in (6.7) are in }R and U · v denotes the familiar dot product in several variables. Note that the complex vector spaces behave very much like real vector spaces. But, in the complex case, one requires some extra care while defining the inner product. For example in }R2, we get (u,u) = u +U for U = (Ul,U2) E}R2 which fails to satisfy (14) if U = (Ul, U2) is considered as a vector in C2. In fact, if Ul and U2 are in C then u + u may be negative or complex. 6.8. Other inner products on r. One can define a general inner product on }Rn in the following manner: let U = (Ul, U2, . . . , un), v = ( Vl , V2, . . . , v n ) E }Rn and (u,v) =  Ui (  aijVj) - E aijUiVj li,jn where A = (aij) is a real positive definite n x n matrix. Here positive definiteness is to mean the following: . A is a symmetric matrix . L:j=l aijUiUj > 0 for each (Ul, U2, . . . , un) E }Rn and equality holds in this inequality only when Ul = U2 = · · · = Un = O. Analogously, one can define a general inner product on en by refining the above inner product by n (u, v) = E aijUi Vj i,j=l where A = (aij) is a complex positive definite matrix, Le. A is a Hermitian matrix, L:j=l aijUi Uj > 0 for each (Ul, U2,. . . , un) E en and that the equality holds in this inequality only when Ul = U2 = . · · = Un = O. 
6.1. Inner Product 347 6.9. Example. Let X and Y be two inner product spaces with the corresponding inner products (.,.) x and (., .) y, respectively. Then the Cartesian product X x Y defined by X x Y = {(x, y) : x E X, Y E Y} has a natural inner product on X x Y and is defined by ((XI,YI), (X2,Y2))XXY = (XI,X2)X + (YI,Y2)y which in turn gives us a natural norm on the inner product space X x Y: lI(x, y)llxxY = IIxll + lIylI}. Unless otherwise is stated explicitly, this will be the norm considered as the Cartesian product of two inner product spaces. The same idea can be considered for Cartesian product Xl x X 2 x.. . X n of finite number of inner product spaces Xl, X 2 ,. . · , Xn. . 6.10. Example. Consider Mnxn (C), the complex vector space of all n x n matrices with complex entries. For A, B E Mnxn(C), define (A, B) = trace (A* B), where trace of a matrix is the sum of its diagonal entries and ,*, is the Her- mitian conjugation (= conjugate transpose) of matrices. It is easy to verify that this is the standard inner product obtained by identifying Mnxn (C) with cn 2 . In particular, in the real vector space Mnxn(IR), (A, B) = trace (At B) defines an inner product on Mnxn(IR), where At stands for the transpose of the matrix A. . 6.11. Definition. (see Corollary 6.21) If V is an inner product space equipped with an inner product (., .), the norm 11.11 on V associated with the inner product is the nonnegative real number defined by the formula (6.12) lIuli = v (u, u) for u E V. In (6.12), we take the nonnegative square root. It is not immediately clear whether the definition (6.12) satisfies the triangle inequality. We shall soon check that (6.12) defines a norm which we say 'the norm induced by the given inner product' or an induced norm of the given inner product. Indeed, it is evident that for every u E V and A E C, by (6.2), we have the homogenei ty condition (AU, AU) = IAI2 lIull 2 , Le. IIAul1 = IAllluli. 
348 Chapter 6: Inner Product Spaces Also, from Definition 6.1, it follows that lIuli > 0 with equality iff u = o. This property is called positivity condition for the norm. We remark that the norm of a vector u E V, where V = en or }Rn , with respect to the standard inner product defined in 6.5 is the familiar Euclidean length: for z = (z 1 , Z2, . . . , Zn) E en, n Ilzll = v (z,z) = E I Z kl 2 k=l which we have already noticed as the Euclidean distance from the origin to the point z E en. This observation shows that the definition of inner prod- uct space arise naturally as a generalization from the concept of (Euclidean) length that is familiar for tne set of real or complex numbers. Next we derive an important inequality which has many interesting applications in the theory of inner product spaces, and as a consequence we obtain that each inner product space is a normed vector space with the norm II., II = V(0; Le. inner product generates this norm. In other words, the set of inner product spaces is a proper subset of the set of normed spaces. Moreover, there are several essential algebraic identities, variously and ambiguously called Polarization identities. We shall state and prove these identities in the following Proposition. These and other closely related identities are of constant use. Now we are in a position to state and prove the above mentioned important inequality known as the Cauchy-Schwarz- Buniakowski inequality (briefly, we say CSB inequality), and we shall also use this to define the concept of angle by means of a formula (6.25). 6.13. Proposition. (Cauchy-Schwarz-Buniakowski inequal- ity) Every complex inner product space (V, (', .)) equipped with the norm (6.12) satisfies the CSB inequality: for each u, v E V (6.14) l(u,v)1 < lIullllvll, and the equality holds iff u, v are linearly dependent, where 11.11 is as defined by (6.12). In particular, for every u, v E V, we have (i) lIu + vII < Ilull + IIvll [Triangle inequality] and the equality holds iff either u = 0 or u = AV, A > o. (ii) Ilu + vl1 2 + lIu - vl1 2 = 2(lIu1l 2 + Ilvll 2 ) [Parallelogram identity] (Hi) When 1F = C, i.e. for complex inner product space, we have (6.15) Ilu + vl1 2 - tlu - vl1 2 + i(lIu + ivl1 2 - Ilu - ivll 2 ) = 4 (u, v); or equivalently, 
6.1. Inner Product 349 4 E ikllu + i k vl1 2 = 4 (u,v). [Polarization identity] k=l (iv) When F = JR., i.e. for real inner product space, we have Ilu + vl1 2 - Ilu - vl1 2 = 4 (u, v). [Polarization identity] Proof. Clearly, if one of u and v equal to zero then (6.14) holds. There- fore, we can assume that neither of u, v is zero so that Ilull and IIvll both become nonzero. This means that (6.14) is equivalent to I(U, V)I < 1 where U = u/liull and V = v/llvll. For every complex quantity A, by (6.4) and by the fact that IIUII = IIVII = 1, we have o < IIU + AVI1 2 = 1 + IAI2 + A (U, V) + A(U, V) which, in particular for A = - (U, V), is equivalent to (6.16) o < 1 -I(U, V)1 2 and (6.14) follows from this. Let I (u, v) I = lIullllvll and let both u and v be different from zero (oth- erwise, u and v are trivially linearly dependent). Then, (6.16) implies that U + A V = o. Thus, U, V, and hence u, v are linearly dependent if equality in (6.14) holds. Conversely, if u # 0, v # 0 are linearly dependent then there exist A E F, such that u = AV. From this, we deduce that I (u, v) 1 2 = I (AV, v) 1 2 = IAI2 (v, v)2 = (AV, AV) (v, v) = (u, u) (v, v). (i) By (6.4) we easily get (6.17) (u + v, u + v) = Ilu + vll 2 = IIull 2 + 2 Re (u, v) + IIvll 2 so that using the CSB inequality (6.14), the above equality yields Ilu + vll 2 < lIul1 2 + 21(u, v)1 + IIvl1 2 < (Ilull + IIvl1)2 and the triangle inequality follows. Next, we consider the equality case of the last inequality. Assume that Ilu + vII = lIuli + Ilvll and v # o. Then, squaring the last inequality gives Ilull · Ilvll = Re (u, v) = I (u, v) I which means that lIull.llvll = (u, v) = I(u, v)l. 
350 Chapter 6: Inner Product Spaces Thus, from the equality case of (6.14), we see that there exists a nonzero A in 1F such that u = AV. Since I (u, v) I = (u, v) = A(V, v), we conclude that A > o. Conversely, if u = AV for A > 0 then Ilu + vII = 11(1 + A)vll = (1 + A)llvll = IIvll + IIAVIl = IIvll + lIuli. (ii) Recall the equation (6.17) Ilu + vl1 2 = IIul1 2 + 2 Re (u, v) + IIv11 2 . Similarly, we have (6.18) Ilu - vll 2 = lIull 2 - 2 Re (u, v) + IIvll 2 and therefore the parallelogram law follows just by adding the last two equations. (iii) In a similar vein, subtracting (6.18) from (6.17), we obtain lIu + vll 2 - Ilu - vll 2 = 4 Re (u, v). From replacing v by iv in the above equation and then using the formula (6.2) with A = 1 and J.t = i, we directly see that Ilu + ivl1 2 - Ilu - ivl1 2 = 4 Re (u, iv) = 4 Re {-i(u, v)} = 4Im (u, v). Now the polarization identity (Hi) follows from the last two equalities, since ( u, v) = Re (u, v) + i 1m (u, v) . . 6.19. Remark. As an alternate proof of the CSB inequality (6.14) where u and v are nonzero members of a real inner product spaces V, one can consider the linear combinations u v u V Wl =  +, W2 =  -  so that the positivity condition of the inner product gives (W j , W j) > 0 (j = 1, 2), which for j = 1,2 is easily seen to be equivalent to -lIullllvll < (u, v) and (u, v) < lIullllvll, respectively. From this, it follows that I(u, v)1 < lIullllvll. . 6.20. Corollary. IEV:F {OJ, then lIuli = sup I(u, v)l. 111111=1 
6.1. Inner Product 351 Proof. There is nothing to prove if lIuli = o. Therefore, we assume that U :j:. 0 so that by the definition lIuli = / u, - II U II ) < sup I (u, v) I \ U IIvll=1 and by the CSB inequality, we have lIull = sup lIull'lIvll > sup l(u,v)l. IIvll=1 IIvll=1 The conclusion follows by combining the last two inequalities. . 6.21. Corollary. If (.,.) is an inner product on a vector sp ace V , then V is a normed space, where the norm on V is defined by lIull = V (u, u). Proof. The axioms (N1) and (N2) are clear from the definition of lIuli while (N3) follows from the triangle inequality in Proposition 6.13(i). . 6.22. Observations. (i) In general, a norm in a normed space need not be induced by an inner product (see Example 6.33). However, the existence of a norm in an inner product space follows from Corollary 6.21 in a natural way. (ii) Let (V, (., .)) be an inner product space, considered as a normed space with 'induced norm'. Then the inner product (.,.) is a continuous function from V x V into the scalar field IF; for, if we let Un  u and V n  v in V, then, with the help of the CSB inequality, we have I(un,v n ) - (u,v)1 = I(un,v n - v) + (un - u,v)1 < I(un,v n - v)1 + I(u n - u,v)1 < lIunllllv n - vII + Ilu n - ullllvil so that the continuity of the inner product (.,.) follows from the fact that every convergent sequence in 1F is bounded, i.e. {liu n II} is bounded. In particular, for each point y E V, the map x I-t (x, y), i.e. fy(x) = (x,y), is continuous linear functional on V. However, several interesting properties of continuous linear functionals will be discussed in detail in Chapter 7. (Hi) We have observed that }Rn is an inner product space with the standard inner product (see 6.5). In }R3 , we can calculate the angle between two nonzero vectors 71 and 11 using the law of cosine from trigonometry (see Figure 6.1): ( 6.23) 1171 - 11 11 2 = 117111 2 + 1111 11 2 - 211lt 111111 II cos (} , 
352 Chapter 6: Inner Product Spaces B o Figure 6.1: Law of cosine where () is the acute angle between 7t and 1. It is then natural to raise the following question: "Is it possible to generalize the concept of angle between vectors in an inner product space (V, (', .) ) ?". For two nonzero vectors u and v in V, we have (see (6.18)) (6.24) lIu - vll 2 = lIull 2 - 2 Re (u, v) + IIvl1 2 and by the CSB inequality IRe (u, v) I < I (u, v) I < lIullllvll, so that Re (u, v) -1 < lIullllvll < 1. Therefore, if we define Re (u, v) cos,p = lIullllvll then (6.24) becomes lIu - vll 2 = lIull 2 - 211ullllvil cos cP + IIvll 2 which agrees with (6.23) when V = ]R3. Now it makes sense to define the angle between u and v by (6.25) ( Re (u, v) ) ,p = Arccos lIullllvll ' where Arccos denotes the principal value so that cP E [0,1r]. Thus, the CSB inequality is helpful in extending the concept of angle be- tween two vectors in plane geometry to inner product spaces. We also mention that such representation given by (6.25) is consistent with what we know about cosine in the context of real Euclidean space, in particular: for every u, v E }Rn , the dot product u . v given by u · v = Ilullllvll cos(). . 
6.2. Examples of Hilbert Spaces 353 6.26. Example. In}R2 equipped with an inner product (see Exercise 6.135( d)) (u, v) = U1 V1 + (U2V1 + U1 V2) + 2U2V2, the angle between (1,0) and (0, 1) is calculated as () ((1,0), (0,1)) 1 cos = II (1, 0) 1111 (0, 1) II = V2 and therefore, () = 1r /4. Note that the angle between (1,0) and (0, 1) with respect to the standard inner product on }R2 is 1r /2. This example shows that the angle between the two vectors in a given inner product space depends on the choice of the inner product of that space. . 6.27 . Natural metric. The "mean-square distance" d( u, v) between two elements u and v in an inner product space is defined by d(u,v) = lIu - vII = y (u - v,u - v). Recall that the reflexivity, symmetry, and the positivity conditions of this distance function are clear from the definition of inner product (.,.) while the triangle inequality d(x, z) < d(x, y) + d(y, z) is equivalent to lIu + vII < Ilull + IIvll (see Proposition 6.13(i) with u = x - y and v = y - z). This metric d defined by the 'induced norm', is called "mean-square metric", or "natural metric", see page 145. Note that if (u, v) = 0, then the mean-square distance gives d(u, v) = (11u1l 2 + IIvIl 2 )1/2. 6.2 Examples of Hilbert Spaces An inner product space V is called a Hilbert space if it is complete with respect to the mean-square metric (Le. with respect to the induced norm 11.llv). In other words, a vector space V over the field 1F is a Hilbert space iff the following two conditions hold: (i) there is an inner product on V (ii) every Cauchy sequence with respect to the induced norm is conver- gent. Most often we prefer to work with Hilbert spaces, since it is handy to have the limits of Cauchy sequences available. Further, it is clear that every Hilbert space is a Banach space whose norm is generated by its inner product. However, the converse is not true in general, as we see from Example 6.33 and also from several other examples. 
354 Chapter 6: Inner Product Spaces 6.28. Subspaces and closed subspaces. Among all the subsets of an inner product space (and Hilbert space), the only subsets which playa distinguished role are those which have certain special properties (eg. linear structure) on them. Now we consider these subset which we use quite often. Let X = (X, (., .)) be an inner product space and Y be a nonempty subset of X. Then (i) Y is said to be a subspace of X if Y is a linear subspace of X, equipped with the inner product reduced by X, Le. (x, y)y = (x, y) for all x, y E Y Evidently, Y is, likewise, an inner product space. (ii) A subspace Y of X is said to be a closed subspace of X if Y is closed in X considered as normed space under the induced norm !!'lIx. (iii) If X is a Hilbert space, then a subspace (closed subspace) Y of X regarded as an inner product space is said to be a subspace (closed subspace) of the Hilbert space X. It is important to emphasize that a subspace Y of Hilbert space X, according to our definition, need not be a Hilbert space, because Y mayor may not be complete as a normed space. 6.29. Examples. Every finite dimensional inner product space is a Hilbert space, since every finite dimensional normed space is complete (see Theorem 5.18). In particular, the Euclidean space en (JRn, respectively) is a Hilbert space. Recall that en is a Banach space with respect to the Euclidean norm (see 3.3) ( n ) 1/2 II ( Z 1 , Z2, · · · , Zn) 112 = E ! Z k 1 2 k=1 and we observe that this norm is derived from the inner product defined by (6.6). The space X of all real polynomials p(x) of degree less than or equal to rt defined over a closed interval [a, b] with the inner product {P, q} = lb p(t)q(t) dt is finite dimensional inner product space, and hence a Hilbert space. . Now we have the following simple result: 6.30. Proposition. Let X be a Hilbert space (resp. Banach space) over the field F. Let K be a subspace of X and K be its closure. Then K 
6.2. Examples of Hilbert Spaces 355 is also a Hilbert space (resp. Banach space) equipped with the same inner product (resp. norm) as X. Proof. From Corollary 3.9, we see that K is a subspace of X and it is clear that K , as a restriction of the inner product (resp. norm) of X, is an inner product space (resp. normed space). Since K is a closed subspace of the complete space X, it follows that K is complete. _ Moreover, the following result is easy to prove! 6.31. Proposition. Let X be a Hilbert space and Y, a subspace of X. Then Y is closed in X iff Y is complete. 6.32. Example. On }R2, consider the norm lI(x,y)lIl - This norm does not satisfies the parallelogram identity. Ixl + Iyl. . The parallelogram identity gives a criterion for a normed space to be- come an inner product space. This is an important property which we use below to obtain examples of Banach spaces which are not inner product spaces. 6.33. Is lP a Hilbert space? For 1 < p < 00, consider the IP-space defined in Example 2.32: 00 lP = {z = {Zn}nl : IIzlI = E IZkl P < oo}. k=l Suppose that 1 < p :j:. 2 < 00. Then, it is easy to see that there exists no inner product (IP, II · lip) such that its induced norm is II · lip. In fact, for Z = {-1,-1,0,0,...} and w = {-1,1,0,0,...} in lP, we have IIzllp = Ilwll p = 21/p and because of the coordinate-wise addition as a rule, we compute IIz + wll p = IIz - wll p = 2. Clearly, the parallelogram identity (see Proposition 6.13(ii)) is not satisfied for p :j:. 2, 1 < p < 00. Therefore, IP for 1 < p :j:. 2 < 00 cannot be an inner product space. However, we have seen in Example 2.40 that IP is Banach space for all p such that 1 < p < 00. Now, for p = 2, we define 00 (z, w) = E Zk W k, k=l z = {Zk}k>l and w = {Wk}k>l in 1 2 . - - 
356 Chapter 6: Inner Product Spaces Since E  llzkl 2 < 00 and E  llwkl 2 < 00 so that the product series E  1 Zk W k converges by the CSB inequality (6.14) for series, this definition is well defined. It clearly satisfies all the conditions of the inner product. Hence, 1 2 becomes an inner product space. Again for 1 2 , using the mean square distance, namely 00 d(z,O) = IIzll2 = v (z,z) = E I Z kI 2 , k=l the completeness property has already been proved in Section 3.4. We recall that the space 1 2 is a particular Li(X, p,) space, with X = N and p, is the counting measure. Even though every Hilbert space is a Banach space, but there exist plenty of Banach spaces which are not Hilbert spaces. Again, in the case of loo-space, the failure of parallelogram law in 1 00 may be used to show that 1 00 is not an inner product space in such a way that its induced norm is 11.1100; Le., it is impossible to define an inner product on 1 00 such that (z, z) = IIzll for all z E 1 00 . Thus, the only lP space which is a Hilbert space is 1 2 , an infinite dimensional space with the inner product defined above. 6.34. Is (CJF[a, b], II · 11(0) a Hilbert space? We show that it is im- possible to construct an inner product on the infinite dimensional space CF[a, b] such that the corresponding norm it induces is defined by the uni- form/maximum norm : IIflloo = max If(t)l. tE[a,b] To prove this, we consider t- a f(t) = 1 and g(t) = b _ a ' t E [a, b]. Then we have 1 = IIflloo = IIglioo = IIf - glloo = Ilf + glloo - 1 which shows that IIf + gll + IIf - gll = 5, 2(lIfll + IIgll') = 4, and therefore, the parallelogram identity is not satisfied. 6.35. Is (CF[a, b], 11.112) a Hilbert space? For any pair of functions f, 9 on the infinite dimensional space CF[a, b] (see Examples 1.33), define (6.36) (f, g) = lab f(t) g(t) dt, 
6.2. Examples of Hilbert Spaces 357 We recall that this inner product induces the norm 11.112. Also, we note that [a, b] is a bounded interval and therefore the continuous functions f and 9 are bounded. Then the product f(t) g(t) is also continuous (bounded) and therefore, integrable. The formula (6.36) assigns to every pair of functions a complex number and further, it is easy to verify that it possesses three properties, namely, 'Hermitian symmetric, conjugate bilinear and positive definiteness'. Therefore (CF[a, b], (., )) is an inner product space and the inner product defined above is often referred as standard inner product on CF[a, b]. Hence, (CF[a, b], (.,.)) is called the standard inner product space. It is shown in Section 3.7 that CF[a, b] in not complete with respect to 11.112 and hence, it is not a Hilbert space. Throughout, in the space C[a, b] (Le. the case 1F = IR), the bar over g(t) in (6.36), denoting the complex conjugate, has no effect. Thus, for functions in the infinite dimensional space C[a, b] the above formula becomes (6.37) (f, g) = l b f(t)g(t) dt. However, we note that the formula (6.37) no longer defines an inner product if we allow complex valued functions into the formula (6.37). In the complex case, one can directly see that (6.37) is neither Hermitian symmetric nor conjugate (linear in second variables); Le. Proposition 6.3 does not hold in the complex case if we use (6.37). 6.38. Is (LP[a, b], II · lip) a Hilbert space? We show that for 1 < p < 00, the only LP[a, b]-space which is a Hilbert space is L2[a, b]. The L2-space has several other special properties that are not shared by other LP-spaces. For simplicity, we let a = -1 and b = 1. Suppose that 1 < p  2 < 00. Then, it is not difficult to show that there exists no inner product on LP[ -1, 1] which induces the norm II . lip, see Section 3.6. For this, we consider f(x) = 1 + x and g(x) = 1 - x. Then f and 9 are (in equivalence classes) in LP[-I, 1] for all p > 1. More- over, / 1 / 1 2P+1 Ilfll = If(x)IP dx = (1 + x)P dx = -1 -1 P + 1 and / 1 2P+1 . IlglI = (1 - x)P dx = 1 · -1 p+ Because of the coordinatewise addition as a rule we find that IIf + gll = ill 1(1 + x) + (1 - x)IP dx = ill 2 P dx = 2 P + l 
358 Chapter 6: Inner Product Spaces and 1 1 1 1 1 1 2P+1 Ilf - gll = 2 P Ixl P dx = 2 P + 1 Ixl P dx = 2 P + 1 x P dx = . -1 0 0 P + 1 An elementary calculation shows that the parallelogram identity (see Propo- sition 6.13(ii)) is equivalent to the equation (p + 1)2/p - 3 = O. Clearly, p = 2 is a solution and it is a simple exercise to see that p = 2 is the only solution of this equation. Thus, we conclude that the parallelogram identity is not satisfied for p :j:. 2, 1 < p < 00. Therefore, LP[-l,l] (and hence, LP[a, b]) for 1 < p :j:. 2 < 00 cannot be an inner product space. However, we have observed in Section 3.6 that LP[a, b] is Banach space for all p such that 1 < p < 00. If f, 9 E L2[a, b], then from the Holder inequality, it follows immediately that lab I/(t)g(t)1 dt < 11/1I211g112 so that the integral J: f(t) g(t) dt exists as a finite number. Also, we observe that. (f,g) = lab f(t) g(t) dt defines an inner product on L[a, b] (Note that the bar over g(t) has no effect for the real space). Thus, L2[a, b] is an inner product space. As (f, f) = IIfll, the Hilbert space norm V (f, f) and the already existing Banach norm IIfl12 are therefore identical. Hence, L2[a, b] is a Hilbert space. 6.39. Strict norm. Recall that a norm is said to be strict if for x :j:. 0, y :j:. 0, IIx + yll = IIxll + lIyll implies that y = AX for some scalars A > O. Proposition 6.13(i) shows that the norm induced by an inner product is always strict. Note that the supnorm on 1 00 is not strict, because sup IZn + wnl = sup IZnl + sup Iwnl n n n does not imply that Zk = AWk for all k with A > O. For example, choose Z = {I, 0, 0, 1, 0, 0, 0, . . .} and W = {O, 1, 0, 1, 0, 0, 0, . . .} so that IIzlloo = 1 = IIwlloo and Ilz + wl loo - 2 but Z :j:. . AW. 
6.2. Examples of Hilbert Spaces 359 Similarly, the supnorm on C[O, 1] is not strict. In fact, the choice f(x) = x and g(x) = 1 in C[O, 1] imply that 1 = IIflloo = IIglloo, Ilf + glloo = 2 although there exists no A > 0 such that f(x) = Ag(X). The reason in both cases is that it is impossible to construct an inner product on these spaces so that the corresponding norm is an induced one (see 6.33 and 6.34). 6.40. Arithmetic-geometric-harmonic means. For Uk E IR+ (k = . 1, 2, . . . , n), let A ( u) : = A ( U 1 , U2, . . . , un) 1 n - - E Uk n k=l G ( u) : = G ( U 1 , U2, . . . , Un) _ ( iI Uk ) l/n k=l n - n 1 L.Jk=l Uk H ( U) : = H ( U 1 , U2, . . . , Un) denote the arithmetic mean, the geometric mean and the harmonic mean, respectively. Then, from Jensen's inequality (see Proposition 1.53) we ob- tain the Arithmetic mean and the Geometric mean (or briefly AM-GM) inequality: (6.41 ) ( n ) l/n n g Uk <   Uk, Le. G(u) < A(u). Indeed if we put Pk = n for every k = 1, 2, . . . , nand Xk = u/n = U/Pk in Exercise 1.79, we get (6.41). If we replace Uk by l/uk in (6.41), then we see that ( n ) l/n n n 1 < II Uk , Ek=l Uk k=l The last two inequalities show that the harmonic mean is less than or equal to the geometric mean, which itself is less than or equal to the arithmetic mean. Thus, we see that the Jensen's inequality (p. 46), which has a large number of many other applications, is actually an extension of the basic iI;lequality (6.41). Equality in the three means inequalities, Le. H(u) < G(u). H(u) < G(u) < A(u), holds if, and only if, Ul = U2 = · · · = Un. 
360 Chapter 6: Inner Product Spaces 6.42. Corollary. (CSB Inequality) For a given set of complex numbers U1, U2, . . . , Un and V1, V2, . . . , V n , we have  Uk V k < (  I Uk I 2 ) (  I Vk I 2 ). Equality holds iff there exists a A E C such that either Uk = AVk or VA: = AUk holds for each k. In particular for Uk E IR+ for each k = 1,2, . . . , n, we have n  n 2 n < .!.  U < Lik=1 Uk E n 1 - L....J k - , - n n k=1 Uk k=1 i.e. the harmonic mean of U1, U2, . . . , Un E IRt is less than or equal to the arithmetic mean of these numbers which is less than or equal to the root-mean square of these numbers. Proof. Recall that (en, (., .)) is an inner product space, where the inner product and the norm are defined by (6.6) and (6.12), respectively. Now setting U = (U1'...' un), V = (V1, . . . , V n ) as points in en, we observe that the CSB inequality, namely I(U, v)1 < lIulillvlI, is equivalent to the Cauchy inequality in Corollary 6.42. Equality case holds iff U and V are linearly dependent, Le. either U = AV or v = AU which is equivalent to either Uk = AVk or Vk = AUk for each k and for some A E C. The left hand side inequality follows just by taking Vk = l/uk for every k = 1,2,..., n while the right hand side inequality follows from choosing v k = 1/ n for each k = 1, 2, . . . , n. . Recall that a generalization of the Cauchy-Schwarz inequality of Corol- lary 6.42 is the Holder inequality (see Lemma 2.21). 6.43. Example. If we apply the CSB inequality for each pair of functions f, 9 on CF[a, b], then, according to the inner product defined by (6.36), we find that b ( b ) 1/2 ( b ) 1/2 fa f(t) g(t) dt < fa If(t)1 2 dt fa Ig(t)1 2 dt · If we apply this to the functions If I and Igl in places of f and g, we get b ( b ) 1/2 ( b ) 1/2 fa If(t)g(t)1 dt < fa If(tW dt fa Ig(tW dt 
6.3. Applications of Polarization Identity 361 (see also Holder inequality in Lemma 2.21(iii)). These are two important integral inequalities that were studied by different methods of proof. Fur- ther, the last inequality for g(t) = 1 yields that b ( b ) 1/2 llf(t)1 dt < v b - a llf(t)1 2 dt which means that a square integrable function on a finite interval [a, b] is integrable therein. . 6.3 Applications of Polarization Identity Proposition 6.13 raises a question whether every normed space V over 1F satisfying the parallelogram law, namely Proposition 6.13(iii), is an inner product space on V. Further, an interesting observation to note from Polar- ization identity is that if we know the norm in an inner product space, the inner product can be recovered, which in turn implies that not all Banach spaces are Hilbert spaces, but there are special ones in which parallelo- gram identity holds, and such spaces have enormous consequences. More precisely, we have 6.44. Theorem. A normed space satisfying the Parallelogram iden- tity is an inner product space, where the inner product is necessarily given by the Polarization identity (6.15). Proof. Let X be a normed space with a norm satisfying the Parallelo- gram identity lIu + vll 2 + lIu - vll 2 = 2(llull 2 + IIvI1 2 ), u, vEX. If X is a real normed space, define 1 (u, v) = 4 [IIu + vll 2 - lIu - v1l 2 ] and if X is a complex normed space, define (u, v) with Re (u, v) =  [IIu + vll 2 - lIu - vIl 2 ], 1m (u, v) =  [IIu + ivll 2 - lIu - ivll 2 ] . so that our formula for (u, v) can be rewritten as 4 (6.45) (u, v) = Re (u, v) + iIm (u, v) =  I: i k lIu + i k vll 2 , i = yCi k=1 which is in fact the Polarization identity. We assume that X is a complex normed space as the real case is easy, and show that the above definition makes X into an inner product space. 
362 Chapter 6: Inner Product Spaces (i ) To p rove (11), we take conjugation in (6.45) and obtain 4(v, u) - Ilu + vl1 2 - Ilu - vll 2 + i(lIv + iull 2 - Ilv - iull 2 ) - Ilu + vll 2 -Ilu - vl1 2 - i(lli(u - iv)1I2 -11- i(u + iv)1I 2 ) - Ilu + vll 2 - lIu - vl1 2 + i(lIu + ivll 2 - lIu - ivll 2 ) - 4(u,v). (ii) In view of the definition of (., .), for u EX, we have 4(u, u) = 411ul1 2 + ill + il 2 11ull 2 - ill - il 2 11ull 2 = 411ull 2 so that (14) is satisfied. (Hi) The proof of (12) and (13) require a little work. According to the Parallelogram identity, we have ( 6.46) II (u ::l: w) + v 11 2 + II (u ::l: w) - V 11 2 = 211 u ::l: w 11 2 + 211 v 11 2 . By (6.46), Ilu+w+vI1 2 + Ilu+w-vI1 2 -llu-w+vI1 2 -llu-w-vI1 2 = 2[llu+wI1 2 -llu-wI1 2 ] showing that Re [(u + v, w) + (u - v, w)] = 2Re (u, w). A similar argument implies that 1m [(u + v, w) + (u - v, w)] = 21m (u, w) and the combination of these two equations gives (6.47) (u + v,w) + (u - v,w) = 2(u,w). Since by (6.45) we have (0, w) = 0, from (6.47) with u = v, we deduce that ( 6.48) (2u, w) = 2(u, w). Taking u + v = x, u - v = y, w = z in (6.47) implies that (x, z) + (y, z) = 2((x + y)/2, z) = (x + y, z). which proves (13). From this and (6.48), by induction, we obtain that (nx, y) = n (x, y) for n E N. Thus, (12) holds in the particular case that A is a natural number. However, the same argument for x/n in place of x gives (x, y) = n(x/n, y}, i.e. (x/n, y) = .! (x, y) n 
6.3. Applications of Polarization Identity 363 so that for m, n E N we obtain m m - (x, y) = (-x, y). n n Thus, (12) holds for all positive rational numbers A. Since the set of ratio- nals is dense in ]R, (12) holds for all A E ]R, since (x, y) is continuous in x, by (6.45). In fact, for A < 0, we have A (x, y) - (AX, y) - A (x, y) - (IAI( -x), y) - A (x, y) -IAI(( -x), y) - A (x, y) + A(( -x), y) - A(X + (-x),y) = o. Consequently, A(X,y) = (X,AY) for all real A and x E X. Also, by (6.45), we have (x, iy) =  [!Ix + iyll2 - IIx - iyll2 + illx - yll2 - illx + y1l2] = -i(x, y}, or (ix, y) = i(x, y). Thus, for A = a + i{3, we have A(X, y) - a(x, y) + i{3(x, y) - (ax, y) + i({3x, y) - (ax + i{3x, y) - (AX, y). Therefore, (12) holds for all A E C. . From Proposition 6.13 and Theorem 6.44, we have 6.49. Corollary. A complex (real) Banach space (X, II.ID is a com- plex (real) Hilbert space with inner product (.,.) satisfying Ilull = y (u,u) iff Parallelogram identity holds. When this equality holds, the inner prod- uct is necessarily given by the Polarization identity Ilu + vll 2 - lIu - vl1 2 - i(lliu + vl1 2 -lliu - v1l 2 ) = 4 (u, v) for the complex Banach space (without imaginary part in this identity for the real Banach space). Finally, given a real Hilbert space, is it possible to construct a complex Hilbert space? (see Exercise 6.146). 6.50. Isometry. (see 2.100) Let X and Y be two inner product spaces and let T : X  Y be a transformation between them. The map T is cailed an isometry iff it is linear and preserves the lengths: IITxlly = IIxlix for all x E X. 
364 Chapter 6: Inner Product Spaces We say that X and Yare (linearly) isometric iff there exists an isometry T of one space onto the other. The map T is called an isomorphism if it preserves the Hilbert space structure, Le. if it is bijective and linear isometric. This map can also referred as isometric isomorphism. If there exists an isomorphism between the two Hilbert spaces, then they are called isomorphic or isometrically isomorphic. (i) An isometry is one-to-one because of Theorem 1.43(H) and the fact that Tx = 0 :::} IIxll = 0 :::} x = 0 :::} NT = {OJ. Further, if dim X = dim Y where X and Yare finite dimensional and if T is an isometry, then by the Rank-Nullity Theorem (see Theorem 1.44) we have dim RT = dim Y which shows that T is onto and hence an isometry is an isomorphism. An important observation to record is that, when X is an infinite dimensional space, an isometry T : X  X need not be onto. The standard example for this observation is the unilateral shift map T: 1 2  1 2 , Z = {Zl,Z2,Z3,...} I-t {O,Zl,Z2,.. .}. Note that 1 2 is Hilbert space. (H) It follows that if T is a linear map between two inner product spaces X, Y, then the identity IITxl1 = IIxll is seen to be equivalent to (6.51) (Tx, TX/) = (x, x') for all x, x' EX, because of Polarization identity (6.45): 4 (Tx, Tx') =  ::>k IITx + i k Tx'1I 2 k=l 4 _  I>kIlT(x + i k x'}112 k=l 4 -  L ikllx + i k x'1I 2 k=l - (x, x'). (Hi) With d(x, x') = Ilx - x'II, it follows that T is an isometry iff d(Tx, TX') = d(x, x') for all x, x' E X (see page 118). 
6.4. Completion of Inner Product Spaces 365 (iv) If (J is the angle between any two nonzero vectors x and x' in the inner product space X and if T is an isometry, then, by (ii), we have (x, x') (Tx, TX/) cos (} = IIxllllx'll = IITxIlIlTx'll and hence an isometry preserves angle but the converse is not true. 6.52. Example. Let Y be the normed space of all the complex sequences {zn} such that E  1 n!Zn is convergent and the norm is given by II{zn}11 = E  1 n!lznl. Define T : l1  Y by {zn}  {zn/n!}. Then 00 IIT( {zn})11 = lI{zn/n!}1I = L IZnl = II{ Zn}1I1 n=1 and therefore, T is a continuous operator with IITII = 1. Further, T is bijective and hence T is an isometric isomorphism. . 6.4 Completion of Inner Product Spaces We have discussed the completion of metric spaces and normed spaces in Sections 2.8 and 5.8, respectively. Further, we have already proved that every inner product space is a normed space with respect to the induced norm IIxll = V (x, x) for each x. Thus, it is natural to enquire whether a given incomplete inner product space can be made complete by means of introducing the notion of com- pleteness with the induced norm. We start with a basic result about the convergence of certain scalar sequence. 6.53. Lemma. If {xn} and {Yn} are two Cauchy sequences in an inner product space (X, (., .)), then the scalar sequence {zn}, Zn = (xn,Yn), is convergent. Proof. Let us show that {zn} forms a Cauchy sequence in IF, where IF is either C or III Now, we can write Zm - Zn - (xm, Ym) - (xn, Yn) - (xm,Ym) - (xm,Yn) + (xm,Yn) - (xn,Yn) - (xm,Ym - Yn) + (xm - xn,Yn). Therefore, by virtue of triangle inequality, we have IZm - znl < l(xm,Ym - Yn)1 + I(xm - xn,Yn)1 < IIxm IIIIYm - Yn II + IIxm - X n II llYn II 
366 Chapter 6: Inner Product Spaces where we have used CSB in the last step. Since every Cauchy sequence in a metric space is bounded, there exist two constants Cl, C2 > 0 such that Ilxnll < Cl and llYn II < C2 for all n. Therefore, with C = max{cl, C2} > 0, we have IZm - znl < c(IIYm - Ynll + IIxm - xnll). In view of the assumption that {xn} and {Yn} are Cauchy, we conclude that {zn} is a Cauchy sequence in the complete space IF. So, the sequence {zn} is convergent, Le. lim n -+ oo (xn, Yn) exists. - Now we state and prove the main theorem about the completion of an inner product space. 6.54. Theorem. Given an inner product space X, there exists a Hilbert space X. containing a subspace Xo with the following property: (i) Xo is everywhere dense in X. (ii) Xo is isometric with X: (X,Y)x.=(X,Y)x for all x,yEX. Proof. By Corollary 6.21, X is a normed space with respect to the norm IIxlix = V (x,xh. Therefore, by Theorem 5.89, there exists a Banach space X* and an isom- etry T : X  Xo where Xo C X* and X 0 = X*. We need to equip X. with an inner product and show that it is just the required Hilbert space. Let {xn} and {Yn} be two representatives (Cauchy sequences in X) of their equivalence classes x* and y* in X., respectively. In view of Lemma 6.53, lim n -+ oo (xn, Yn) x exists and therefore, we can set (6.55) (X*, y*) x. = lim (xn, Yn) x' n-+oo To prove that this definition is well defined, we need to show that the limit depends only on x. and y* and not on the chosen representatives {xn} and {Yn} that represent the classes x* and y*, respectively. Indeed, if {x n }, {x} E x* and {Yn}, {y} E y*, then, by the method used in Lemma 6.53, we see that there exists C > 0 such that I(xn, Yn) - (x, y)1 < C[IIYn - y1I + IIxn - x11]  0 as n  00 (since X n f"oJ x and Yn f"oJ Y). That is, lim (xn, Yn) = lim (x, Y) n-+oo n-+oo 
6.5. Orthogonal Family of Vectors 367 and therefore, (6.55) is well defined. Next, we shall show that (x*, y*) x. satisfies the axioms (11)-(13) of the inner product. (11): Clearly, (x*, x*) > 0 as it is a limit of sequence of nonnegative real numbers. If x* = 0*, then {O, 0, . . .} E x* and therefore, (x*, x*) x. = O. Conversely, if o = (x*, x*) = lim (xn, x n ) = lim IIx n l1 2 => lim Ilxnll = 0 n-+oo n-+oo n-+oo then {x n } ,....., {OJ = {O,O...}, Le., x* = 0*. Axioms (12) and (13) are obvious. Hence, (.,.) x. is an inner product on X*. In addition, Ilx* IIx. = lim IIx n l1 2 = lim V llx n l1 2 = lim IIxnllx n-+oo n-+oo n-+oo which means that the norm introduced above coincides with the norm in the Banach space X*. Hence X* is a Hilbert space. Finally, for each x EX, we can associate certain class x* E X* , namely, with the class that contains the stationary sequence { x } := {x, x . . .} (This identification enables one to say that X c X*). Let Xo be the set all such equivalence classes. H { x } and { y } are two such distinct stationary sequences (with x :F y), then Ilx - yll :F 0 and (x* , y*) x. = ({ x }, { y } ) x. = lim (x, y) x = (x, y) x · n-+oo The fact that Xo is dense in X* follows from Theorem 5.89. . 6.5 Orthogonal Family of Vectors The parallelogram identity in Proposition 6.13(ii) has a geometric content and carries its name from the analogous result one has in plane geometry. It says that if one considers the parallelogram, formed by two nonzero vec- tors u and v, in the plane generated by u and v, the sum of the squares of the two adjacent sides of the parallelogram is equal to half the sum of the squares of the diagonals, see Figure 6.2. As pointed out in the begin- ning, the inner product structure of a Hilbert space allows us to introduce the concept of orthogonality of vectors. Let X be an inner product space. For any two vectors u, v in X, if (u, v) = 0 then we say that u is orthog- onal/perpendicular to v, expressed symbolically by u ..L v. For example, (1, i) and (i,l) are orthogonal with respect to the standard inner product on C2. H u ..L v, then Proposition 6.13(iv) becomes Iliu + vll 2 = Ilu - vll 2 = lIu + vl1 2 = IIull 2 + IIv11 2 . In fact, in a real inner product space (V, II · II) we have u ..L v <==} Ilu + vll 2 = IIul1 2 + IIvl1 2 
368 Chapter 6: Inner Product Spaces - - - - _ _ _ _ _+v ' /' 1 /' I /' lIull '\. /' ' I /' /' I X /' , I  /' , I X /  /' , I  , /' 0 IIvll v Figure 6.2: Illustration for parallelogram identity which is the Pythagorean theorem in the classical case: The square of the hypotenuse of a right angle triangle in the plane equals the sum of the squares of the adjacent sides, see Figure 6.3. This simple result is useful for certain calculations. However, the converse of the above result is not necessarily true for complex inner product spaces. For example, for u = (0, i) and v = (0, 1) in C2 we have u+v=(O,l+i), u-v=(O,-l+i), iu+v'=(O,O), which give lIuli = Ilvll = 1, lIu + vll 2 = 2, lIiu + vII = 0 so that in this example, lIull 2 + IIvll 2 = Ilu + vll 2 holds even though u is not orthogonal to v because (u, v) = i :j:. O. However, if V is a complex inner product space and if u, v E V, then it is easy to show that (see also Exercise 6.135) ( 6.56) u ..L v {:::::} IIAu + J-tv1l 2 = IIAull 2 + lIJ-tvll 2 for all pairs of scalars A and J-t in C. For the proof of this simple two way implications, it suffices to note that II AU + J-tv 11 2 = II AU 11 2 + II J-tv 11 2 {:::::} Re { A J-t ( u, v)} = 0 for all A, J-t E C. 
6.5. Orthogonal Family of Vectors 369 v_________ u+v Ilvll o lIuli u Figure 6.3: Classical Pythagorean Theorem Now, taking A = J.t = 1 we obtain that Re (u, v) = 0, and choosing A = 1, J.t = i we get that 1m (u, v) = O. Now the orthogonality implication (6.56) follows. Now let us discus s som e basic properties of the concept of orthogonal- ity. Since (u, v) - (v, u), it follows that the orthogonality condition is symmetric: u ..L v <=> v ..L u. Therefore, we may say without ambiguity that u and v are orthogonal. Note that, u ..L 0 for all u, and u ..L u iff u = O. Further, if (u, v) = 0 for all v in an inner product space V then we have (u, u) = 0 which is equivalent to u = o. Thus, we conclude that the zero vector is the only vector orthogonal to every element of V: u ..L v for all v E V <=> u = o. Few other simple facts concerning the concept of orthogonality are: (i) u..L Vl and u ..L V2 implies u ..L AVl + J.tV2 (ii) u..L V n for n = 1, 2, . . ., and V n  v implies u ..L v. The notion of orthogonality makes it possible to think about vectors in a geometric way. For example, earlier we have defined the angle between two vectors. Indeed, from (6.25), we see that u ..L v iff () = 1r/2, which is also 
370 Chapter 6: Inner Product Spaces clear from the geometric interpretation because cos () = 0 iff () represents the right angle. This fact explain the consistency in the terminology in the plane case: Two nonzero vectors in}R2 are orthogonal precisely when the angle between them is 1r/2. Also, in the formula (6.7) for n = 3 with u = (Ul, U2, U3) and v = (Vl, V2, V3), the orthogonality condition is (u, v) = u · v = 0 which agrees with the elementary concept of orthogonality in the space }R3 . In an inner product space (V, (., .) ), we write {u}..L for the set of vectors which is orthogonal to u E V: {u}..L:={vEV: (u,v)=O}={v: u..Lv}. 6.57. Proposition. The set {u}..L is a closed subspace of the inner product space (V, (., .) ) . Proof. Consider the linear functional f : V  C, v I-t (v, u). Clearly, {u}..L = N/ and the conclusion is immediate in view of the fact that N/ is always a closed subspace. In fact, by the CSB inequality, If(v)1 < Ilullllvll which implies that f is bounded. Thus, for each u E V, the set {u}..L is the inverse image of the closed set {OJ of C under the continuous map v I-t (v, u) and hence, the set {u}..L is closed by Corollary 5.67. . The vector u is orthogonal to a set S C V if u ..L y for all yES. We write, u ..L S. A finite or infinite (countable) set of vectors  = {cPOl : a E A} in an inner product space (V, (., .)) is said to be orthogonal iff cPOl -L cP/3 whenever a, /3 E A, a :j:. /3. The set  is called orthonormal iff, in addition to being orthogonal, each vector in  is a unit vector, Le. (cPOl, cP/3) := 60l/3 where 60l/3 denotes the Kronecker symbol. The standard basis vectors {ej }ljn is the simplest example of an orthonormal set of vectors for the real inner product space (}Rn, (., .)) with standard inner product. 6.58. Example. Consider the space C[O, 1 with the standard inner product defined by (6.37) and norm by 11.11 = (', .). Assume that fn(t) = v'2cos27rnt and gn(t) = v'2sin21rnt, n E N. 
6.5. Orthogonal Family of Vectors 371 First we note that sinn1r = 0 and cosn1r = (-1)n for all n E N. Now for n,m E N we have f ( t) 2 cos 2 27rnt = 1 +  : (sin 47rnt), 1rn t g ( t ) - 2 sin 2 27rnt = 1 -  : (sin 47rnt), · 1rn t fn(t)fm(t) - cos 21r(n - m)t + cos 21r(n + m)t, m :j:. n "!! [sin 27r(n - m)t + sin 27r(n + m)t] m :j:. n, dt 21r(n - m) 21r(n + m) , gn(t)gm(t) - cos 21r(n - m)t - cos 21r(n + m)t, m :j:. n - !! [sin27r(n-m)t _ Sin27r(n+m)t] m :j:. n, dt 21r(n - m) 21r(n + m) , fm(t)gn(t) - sin 21r(n + m)t - sin 21r(n - m)t, m 1= n - !! [COS27r(n - m)t _ cos27r(n + m)t] m :j:. n. dt 21r(n - m) 21r(n + m) , Using these equalities, we easily obtain that Ilfnll = IIgnll = 1, (1, fn) = (1, gn) = 0 for each n E N. Further, for each n, mEN with n :j:. m, it follows that (fn, fm) = (Yn, gm) = (fm, Yn) = O. Therefore, we obtain that the set {I, V2 cos 21rnt, V2 sin 21rnt}n1 forms an orthonormal subset of C[O, 1] with respect to the inner product (6.37). More generally, one can easily see that the set { 1 f2 ( 21rnt ) f2. ( 21rnt ) } vb - a ' V  cos b - a ' V  SIn b - a nl forms an orthonormal set in C[a, b] with the standard inner product defined by (6.37) and the norm by 11.11 = . Note that these are orthonormal unit vectors. . From the definition of the inner product space it is clear that every countable orthogonal set {cPOl : a E A} of nonzero vectors in an inner product space can be orthonormalized by replacing each vector cPOl with cPOl/llcPOlII; Le. the collection { II::II : a E A} · Now we state a simple criterion for orthogonality, which is popularly known as the generalized Pythagorean theorem in abstract form. 
372 Chapter 6: Inner Product Spaces 6.59. Corollary. (Pythagorean theorem) Let V be an inner product space over F. If { 4Jl , . . . , 4Jn} E V is an orthogonal set, then n L Ak4Jk k=1 2 n = L IAk 1 2 II4Jk 11 2 , k=1 where AI, . . . , An are scalars. Proof. The proof follows easily from the orthogonality condition, and (6.4) with m = nand Aj = J.tj = 1 for j = 1,2, . . . , n. T!le conclusion also follows by induction. - Does the converse of Corollary 6.59 hold? An important consequence of Corollary 6.59 is that every orthogonal set {4Jk : 1 < k < n} of nonzero' vectors in an inner product space is linearly independent. How about the orthonormal family  = {4Jo: : a E A} of vector in an infinite dimensional inner product space? More generally, can we show that an orthonormal family {4JO:}O:EA of vectors in V is summable iff the family {114Jo:I1 2 }O:EA is summable? We shall answer these questions in an appropriate place at a later stage. . Two subsets SI, S2 C V are said to be orthogonal, denoted by SI 1.. S2, iff (u, v) = 0 for all u E SI and all v E S2. From the definition, it follows at once that two orthogonal subsets have the unique common element which is indeed the null element. This null element is orthogonal to every subset. Thus, if SI and S2 are orthogonal subspaces of an inner product space V then SI n S2 = {OJ, and hence, in this case, we have SI E9 S2 := SI + S2 c V as the orthogonal direct sum of SI and S2. If S is a nonempty subset of V, then the subspace {v E V : (u, v) = 0 for all u E S} C V is called the orthogonal complement of S, and is denoted by S.l.. Obviously, if S = 0 (empty set) then S.l. = V. Clearly, 0 E S.l. for any subset S. Moreover, the orthogonal complement of S is simply the intersection of the family of closed sets, namely, the collection of all vectors orthogonal to S: S.l. = n {u}.l.. uES Recall from Proposition 6.57 that for each u, {u}.l. is a closed subs pace of V and therefore, "S.l. itself is a closed subspace even if S is not closed". An alternate proof of this result is provided in the next proposition. 
6.5. Orthogonal Family of Vectors 373 For example, if S = {(Xl,X2,... ,Xn-l,O) E IRn : Xl,... ,Xn-l E IR} then SJ.. = {( 0, 0, . . . , 0, X n) E IR n : X n E IR}. In IR3 , if S is the xy-plane then SJ.. is the z-axis. Similarly, if S is the z-axis then SJ.. is the xy-plane. The following results, whose applications will be discussed later, rely very much on the notion of orthogonality. 6.60. Proposition. Let X be an inner product space and S be a subset of X. Then we have (i) SJ.. is closed subspace of X (even if S is not closed) (ii) S C SJ.. .1.. = (S .1.. ) .1.. (Hi) st c Sf- for subsets Sl, S2 of X such that Sl c S2 ( i v) SJ.. .1.. .1.. = SJ... Proof. (i) Clearly, SJ.. is a linear subspace. In fact, let Zl, Z2 E SJ... Then, for each XES, we have (Zl, x) = (Z2, x) = 0 which, for all A, J.t E IF, gives (AZI + J.t Z 2, x) = A(Zl, x) + J.t(Z2, x) = O. Therefore, AZI + J.tZ2 E SJ... Next, we show that SJ.. is closed. Let {xn} be any convergent sequence in SJ.. such that X n  a, where a EX. It suffices to show that the limit point a also belongs to SJ... Since (xn, x) = 0 whenever XES, from the CSB inequality, it follows that for n > 1 and for xES I(a, x)1 - I(xn - (xn - a),x)1 I(xn, x) - (xn - a, x)1 < l(xn,x)1 + I(xn - a,x)1 < o + Ilxn - alllixil  Oasnoo and therefore, (a, x) = 0 which shows that a E SJ... Hence, SJ.. is a closed subspace of X. Alternatively, since the inner product is continuous, we can argue that (a, x) = ( lim X n , x) = lim (xn, x) = 0 n-.+oo n-.+oo so that a E SJ... (H) Let a E S. Then (a, x) = 0 for all x E SJ.., and hence a E (SJ..)J.., Le. S C (SJ..) .1.. . 
374 Chapter 6: Inner Product Spaces (iii) If S1 c S2 and if a E st, then (x, a) = 0 for all x E 82. In particular, (x, a) = 0 for all x E 8 1 so that a E sf- and the assertion follows. (iv) By (ii) and (iii), we have (S.L.L).L C S.L. Also, by (ii), it follows that (S.L) C (S.L ).L.L = S.L.L.L. Hence, S.L.L.L = S.L. . Note that Proposition 6.60(i) assures us that 8.L is complete in the Hilbert space X, by Proposition 2.109. 6.61. Example. Consider M 2X2 (IR), the real vector space of all 2 x 2 matrices with real entries: M 2X2 (1R) = {A = ( :): a,b,c,d E IR} . Then, M 2X2 (IR) is an inner product space with respect to the inner product defined by (see also Example 6.10) (A, B) = trace (At B), for each A,B E M 2x2 (IR). Suppose that we want to find all matrices A E M 2X2 (IR) such that (A, B) = 0, where B is a specific matrix given by B = (1 ). Then, we find that (A, B) = trace (At B) = trace { ( ) (1 )} = b - c which shows that all matrices A such that (A, B) = 0 is of the form ( :)=a( )+b( )+d( ) for some a, b, d E III Hence, B-1 = {A E M 2X2 (1R) : (A,B) = O} = { (: :): a,b,d E JR} and is a linear subs pace of M 2X2 (IR) with dim B.L = 3.  . 6.62. Gram-Schmidt orthonormalization process. Suppose that we are given a nonzero vector y in IR2. Then, we know that each vector x E ]R2 has a unique representation as a sum of two vectors, one parallel to y and the other orthogonal to y. 
6.5. Orthogonal Family of Vectors 375 Our interest is to generalize this idea to inner product spaces. Let y be any nonzero vector in an inner product space X. We know that any vector in the direction of y will be of the form AY for some scalar A, and hence AY is in the subspace K spanned by y: K = span {y}. Now, for any x there exists a unique A such that x - AY is orthogonal to y. Indeed, the value of A such that x - AY is orthogonal to y is obtained from the orthogonality condition o = (x - AY, y) = (x, y) - A(Y, y) = (x, y) - Allyll2 which gives (x, y) A = lIylI 2 . It is this value of A which shows that, for each x, x - AY is orthogonal to y. Now, we can write AY = (x, 11:11 ) 11:11 = (x, ,p},p with 1I,p1l = 1. As with the case of Euclidean space}R2 (or more generally in ]in ), the vector projection of x on K = span {y} = span { cP} is defined by Px = (x, cP)cP. For example, the vector projection of z E en on 0 :j:. wEen is defined by w pz := (z, w) IIwll 2 = (z, ,p},p, w ,p = M' where (z, w) is the standard inner product on en is defined by (6.6). This aspect of Px or PKX will be the starting point for our generalization. In particular, this geometric intuition helps us to formulate the following re- sult: Every finite dimensional inner product space has an orthonormal basis. More precisely, we have 6.63. Theorem. H {U1, U2, . . . , un} is a linearly independent set in an inner product space V, then there exists an orthonormal set { cP1 , cP2, . . · , cPn} of vectors in V such that Lk := span {U1, U2, · · · , Uk} = span { cP1, cP2, . . . , cPk} for k = 1, 2, . . . , n. Proof. As U1 :j:. 0, we begin by normalizing U1 to cP1 of unit length: Ul = Cl,pl, i.e.,pl = II: II with lI,plll = 1. The projection P2 of U2 on K 1 = span {cP1} is P2 = (U2, cP1)cP1. 
376 Chapter 6: Inner Product Spaces Now, U2 - P2 1. 4>1 so that we can write the orthogonal component V2 as V2 = U2 - P2 = U2 - (U2, 4>1)4>1. Clearly, V2 :j:. 0, since otherwise we would have U2 E K 1, contradicting the fact that {U1, U2} is linearly independent. Thus, we can normalize V2 to obtain 4>2: 4>2 =  = U2 - P2 . II v 211 II u 2 - P211 It follows that span {U1, U2} = span {4>1, V2 + P2} = span {4>1, V2} = span {4>1, 4>2}. Moreover, ( V2 , 4>1) - (U2 - (U2, 4>1 ) 4>1 , 4>1) - ( U2, 4>1) - (U2, 4>1) ( 4>1 , 4>1 ) ( U2, 4>1) - (U2, 4>1) = 0 so that 4>2 1. 4>1 with 114>211 = 1. Next, define K 2 = span {4>1, 4>2} and work in three dimensional subspace containing 4>1,4>2 and U3. The projection P3 of U3 onto K 2 = span {4>1, 4>2} is then the sum of the individual projections ( U3 , 4>1) 4>1 + (U3, 4>2) 4>2 so that the orthogonal component V3 is given by V3 = U3 - P3 and therefore, 4>3 =  = U3 - P3 . II V 311 II u 3 - P311 At the m-th step, we work on Km-1 = span {4>1, 4>2, · · · , 4>m-1} where m-1 Pm = E (U m ,4>k)4>k k=l defines the projection of U m on K m - 1 . Again, V m = (u m - Pm) 1. K m - 1 , V m  o. Thus, we may normalize V m by A,. _ V m U m - Pm 'Pm - - , IIvmll lIu m - Pm II where span { 4>1, 4>2, . . . , 4>m} = span { U1 , U2, · · · , U m }, 
6.6. Projections on Finite Dimensional Spaces 377 and {cPl, cP2, . . . , cPm} is an orthonormal subset. This process can be contin- ued until the n-th step and as a consequence of this we obtain the desired orthonormal system {cPl, cP2, · . · , cPn}. . The idea used in the last theorem, which converts the linearly inde- pendent sets into an orthonormal ones with the same span, is called Gram- Schmidt orthonormalization process. The simplest example of an orthonor- mal system in the Hilbert space 1 2 is the Schauder basis {en} n 1. More importantly, if {Ul, U2, . . . , Un} is a linearly independent set such that K = span { Ul , U2, · . · , un} = span { cPl , cP2, . . . , cPn} then the following simple test is easy to verify. 6.64. Proposition. We have, U ..L K iff U ..L Uk for each k - 1, 2, . . . , n iff U ..L cP k for each k = 1, 2, . . . , n. 6.6 Projections on Finite Dimensional Spaces We shall soon see that the concept of orthonormal bases of in inner product space makes it easy to compute the Fourier coefficients of a vector, and this notion simplifies certain aspects of the theory. Let us start with a simple example. Recall that, for any vector y in an inner product space X, the (vector) orthogonal projection of x on K = span {y} = span {cP}, IIcPll = 1, is defined by PKx := (x, cP)cP. For example, if X = 12(N) and cP = ek, where ek = {6km}m1 with 6 km = 0 for k:j:. m and 6kk=1, then, for Z = {Zn}n1 E 1 2 , PKZ:= Zkek for K = span {ek}. This is an example of one dimensional projections. Next we observe that, if each vector x in the spanning set K is expressible as a linear combination of the orthonormal vectors {cPl, cP2, . . . , cPn} then this representation is unique; Le. if n X = E CjcPj for some Cj'S, j=1 then, for each k, the coefficient Ck (also called component of x with respect to the orthonormal basis) is uniquely determined by the formula (X,,pk) = (  Cj,pj,,pk) =  Cj(,pj,,pk} = Ck (k = 1,2,... ,n) 
378 Chapter 6: Inner Product Spaces so that each x in its span has the unique representation ( 6.65) n X = L (x, cPj)cPj. j=1 Now, the following proposition is trivial. 6.66. Proposition. Let {cP1, cP2, . . . , cPn} be an orthonormal basis for a subspace K of an inner product space X. Then for any two sequences {ak}k1 and {bk}k1 of scalars, we have (  ajj,  bii) =  ajbj, Equivalently for x, y E span { cP1 , cP2, . . . , cPn}, we have n (x, y) = L (x, cPj) (y, cPj). j=1 Equation 6.65 actually infers that the best approximation of x by a linear combination of the cPj'S (j = 1, 2, . . . , n) is the one having for com- ponents the scalars (x, cPj). More precisely, we have 6.67. Proposition. Let { cP1 , cP2, . . . , cPn} be an orthonormal basis for the subspace K of an inner product space X, i.e. K = span { cP1 , cP2, · · · , cPn}. If x E X \K, then the (orthogonal) projection of x on K is given by n k = L (x, cPj)cPj. j=1 (Again, remember that k is indeed the closest point of K to x). Equivalent formulation of Proposition 6.67 may be stated as follows. 6.68. Proposition. Let K be an n-dimensional subspace of an inner product space X. Then it has the orthogonal decomposition X = K E9 K.L . Proof. Clearly k E K and (k, cPj) = (x, cPj) so that, for each j - 1, 2, . . . , n, we have (x - k, cPj) = 0, Le. x:..... k ..L y for all y E K. This observation shows that x - k E K.L and therefore, we have the unique representation x = k + (x - k) with k E K and x - k = k.L E K.L. 
6.6. Projections on Finite Dimensional Spaces 379 Hence, k is the projection of x on K and X = K + K..L. Moreover, y E K n K..L implies that (y,y) = 0, Le. y = 0 and hence, X = K E9 K..L. . In particular, if X is finite dimensional, we have K = K..L..L and hence, in this case, K is the orthogonal complement of K..L. However (see Examples 6.90 and 6.91), this fact fails to hold if X is an infinite dimensional space. Let x and k be as in Proposition 6.67. Then for each y E K, we have k - y E K. Since k - y E K, we note that x - k E K..L => X - k .1.. k - y for each y E K. Thus, IIx - yl12 = Ilx - k + k - yl12 = Ilx - kl1 2 + Ilk _ yl12 so that Ilx - yll > Ilx - kll for all y E K. This is the meaning of the expression "closest vector to x from K". If IIx - yll = IIx - kll for some y, then, by the above equality, we have k = y. This confirms the uniqueness of k as the closest vector. 6.69. Example. Let X = IR3 and K = span {VI, V2}, where VI = (1,0,2) and V2 = (1, -1,0). It is easy to verify that K = span {<PI, <P2}, where 1 1 <P1 = . R (1, 0, 2), <P2 = . R (4, -5, -2), v 5 3v 5 and { <P1 , <P2} is the orthonormal basis for the subspace K. A typical element (a, b, c) E K can be obtained from solving the equation (a, b, c) = a(1, 0, 2) + ,8(1, -1,0). From the last equation we see that (see also Example 6.96) K = {(a, b, 2a + 2b) E IR 3 : a, b E IR}. According to Proposition 6.67, the projection k of x on K is given by k = (x, <PI )<P1 + (x, <P2)<P2. Again, we note that k E span {<P1, <P2} which is nearest to x. Now if we choose, for instance, x = (1,1, 1) E X \K, then the closest vector to (1,1,1) IS k - ((1,1,1),  (I,O, 2) )cPl + ((1,1,1), 3 (4, -5, -2) )cP2 1 1 - y'5 (1 + 0 + 2)cPl + 3y'5 (4 - 5 - 2)cP2 1 - 3 (1,1,4). 
380 Chapter 6: Inner Product Spaces Finally, we remark that the method of Example 6.81 would yield the same answer. . 6.70. Example. Let X = }R3 and K = {V1}, where V1 is a given nonzero single vector in }R3. Then, the orthogonal complement K.L of K is the plane through the origin perpendicular to the given vector V1. Note that the orthogonal complement of {OJ is the whole space, and vice versa. Here 0 means (0,0,0). If M = {V1, V2}, where V1 and V2 are two distinct nonzero nonparallel vectors. Then M.l is the intersection {V1}.L n {V2}.L and hence is a line through origin and perpendicular to the plane containing V1,V2. . 6.71. Example. Consider the space C[O, 1] with the standard inner product defined by (6.37) and the subs paces K 1 = span {1 } and K 2 = span {1, t}. Choose x = t 3 E C[O, 1] and Xo = a E K 1 . Then x - Xo E Kt if 1 1 1 o = (t 3 - a, I) = 0 (t 3 - a) · 1 dt = 4 - a, . 1 I.e. a = 4 . Thus, the projection of t 3 onto K 1 is 1/4. Now, we work on the subspace K 2 . For a general element a + bt E K 2 , the condition on a and b such that t 3 - (a + bt) E Kt is determined from the orthogonality conditions: ( 3 ) 1 1 3 1 b o = t - a - bt, 1 = 0 (t - a - bt) · 1 dt = 4 - a - 2 ' and Le. 4a + 2b = 1, 3 1 1 3 1 a b O=(t -a-bt,t) = (t -a-bt).tdt=-----, o S 2 3 Solving the last two equations for a and b, we find a = -l/S and b = 9/10 and therefore, the projection of t 3 onto K 2 is the function I.e. lSa + lOb = 6. 1 9 - S + lO t. Hence, we can write the decomposition of t 3 onto K 1 and K 2 respectively as t 3 = ! + ( t 3 - ! ) 4 4 ' 13 1 .1 where 4 E K 1 , t - 4 E K 1 , and 3 9 1 ( 31 9 ) t = 10 t - S + t + S - lO t where 9 1 3 1 9 .L 10 t - S EK 2 , t + S - 10 tEK2. 
6.6. Projections on Finite Dimensional Spaces 381 Finally, it is easy to verify that x = t 2 can be decomposed into the sum 2 1 ( 2 1 ) t = 3 + t - 3 ' 1 2 1  where 3 E K 1 , t - 3 E K 1 , and 2 1 ( 2 1 ) t =t--+ t +--t 6 6' 1 2 ( 1 )  where t - 6 E K 2 , t - t - 6 E K 2 · Since {I, t} is a linearly independent set, by Gram-Schmidt orthonor- malization process, we can orthonormalize this set to obtain K =. span { tPl , tP2 } , tPl = 1, tP2 = Vi2 (t - 1/2). Thus, the projection operator P on C[O, 1] is defined by Px = P(k + k) = Pk, for k E K 1 , Pk = 0 for k E Kt, and similarly for K 2 . Now we can compute IIt 3 - kll for k E K 1 (resp. for K 2 ). For example, for k = 1/4 E K 1 , we have 1 2 1 1 ( 1 ) 2 1 2 1 9 t 3 - 4 2 = 0 t 3 - 4 dt = 7 - 16 + 16 = 112 ' and, for a E K 1 , we have II 3 11 2 1 1 ( 3 ) 2 1 a a 2 ( 1 ) 2 9 t - a = t - a dt = - - - + - = a - - + -. 2 0 7 2 4 4 112 We observe that for each a E K 1 , we have IIt 3 - al12 > IIt 3 - 1/4112' Le. IIt 3 - 1/4112 = dist (t 3 , K 1 ). Similar conclusions can be drawn for the subspace K 2 (see Theorem 6.74). Once again, we recall that the problem of finding the closest element Xo = xo(t) = a* + b*t to x = x(t) E C[O, 1] on the subspace K = span {I, t} is obtained by determining a* and b* for which IIx - XOll2 < IIx - (a + bt)1I2 for all a, b E IR. Since K = span {tPl, tP2}, tPl = 1, tP2 = Vi2(t - 1/2), this problem is equivalent of finding two scalars c* and d* such that Ilx - (C*tPl + d*tP2)112 < IIx - (CtPl + dtP2)112, for all c, d E IR, where c* = (x, tPl), and d* = (x, tP2). Notice that {tPl, tP2} is the orthonor- mal basis for K, . 
382 Chapter 6: Inner Product Spaces 6.72. Example. Consider the real Hilbert space L2[-1, 1] and K = span {t k : k = 0, 1, 2, 3, 4}. Then K is the space of all polynomials of degree at most 4, and therefore, K is a 5-dimensional Hilbert subspace of L2[-1, 1]. Choose x = x ( t) = cos tEL 2 [ -1, 1]. Then we know that the orthogonal projection xo = Px on K is the best approximation of x by a polynomial of degree 4. Thus, IIx(t) - xo(t)112 < Ilx(t) - y(t)1I2 for all y E K or equivalently, 4 2 ill I cost - xo(tW dt < ill cost - I>kt k dt k=O for all ak E IR, 0 < k < 4. The procedure of finding xo is exactly same as in the previous example. Using Gram-Schmidt orthonormalization process, convert the basis B = {t k : k = 0, 1, 2, 3, 4 } into an orthonormal basis  = {tPk : k = 0, 1,2,3, 4} for K with respect to the L2-inner product. Then, we get 4 xo(t) = EaktPk(t) k=O where a k = (x, tPk) = Jl tPk(t) cos t dt. . 6.7 Orthogonal Projections on Hilbert Spaces We start with a general remark that one way in which Banach spaces differ from Hilbert spaces is that there are, in general, not enough continuous projections. One of the important basic properties of Hilbert spaces asserts that the distance from a point to a nonempty closed convex set in a Hilbert space is always attained, and this result is not valid for arbitrary Banach spaces. This fundamental result is called the 'Minimum principle' (also known as the 'Projection Theorem': "Every Hilbert space has the unique closest point property for convex sets", see Theorem 6.74. Beware of the fact that there is no analogous theorem for arbitrary Banach spaces. Recall 
6.7. Orthogonal Projections on Hilbert Spaces 383 that, for a subset K of a metric space X, a point Xo E K is said to be the closest point to x E X from K if d(x, xo) = inf d(x, y). yEK We also say that Xo minimizes the distance from x to K or Xo is the pro- jection of x on K. First we start with a result which characterizes the closest vector in terms of an orthogonality condition as given below. 6.73. Theorem. Let K be a subspace of an inner product space X. Suppose Xo E K and x EX. Then Ilx - xoll = dist (x, K) <==} x - Xo E KJ... Proof. '=>': Fix x EX, and suppose that Ilx - Xo II = dist (x, K). We claim that x. - xo..Ly for all y E K. Suppose on the contrary that there exists an y E K such that A = (x - Xo, y) # O. Without loss of generality we may assume that Ilyll = 1 since otherwise we can divide y by its norm. Since K is a subspace, z = Xo + AY E K. But then IIx - Zll2 - (x - z, x - z) - (x - Xo - AY, x - Xo - AY) 2 - 2 2 - IIx - xoll - A(Y, x - XO) - A(X - Xo, y) + IAI Ilyll - IIx - xol1 2 - IAI2, since Ilyll = 1 and A = (y, x - xo), < IIx - xoll 2 = (dist (x, K))2 which is a contradiction. Consequently, (x - xo,y) = 0 for all y E K, Le. x - Xo E KJ... '{::': Let x-xo E KJ... Then, for each y E K, the Pythagorean theorem yields IIx - (xo + y)1I2 = IIx - Xo - yl12 = IIx - xol1 2 + lIyll2 > IIx - xol1 2 which gives IIx - xoll = dist (x, K). . 6.74. Theorem. (Projection Theorem) Let X be a Hilbert space, K a nonempty closed convex subset of X and x EX. Then there exists a unique element Xo E K that minimizes the distance from x to K: IIx - Xo II = dist (x, K) = inf Ilx - yll. yEK 
384 Chapter 6: Inner Product Spaces Proof. We start with a small remark. Note that d = dist (x, K) > 0 for if d = 0, then x would be a limit point of K and therefore must belong to K, since K is closed. Moreover, the geometric intuition is not very much reliable in the case of infinite dimensional space. However, according to Theorem 6.73, there exists a point Xo E K closest to x such that minimum is attained iff the vector x - Xo is perpendicular to K. Now, we start proving the theorem. Note that K -x = {y -x : Y E K}. Since inf Ilx - yll = inf IIx - yll = inf lIy'li yEK y-xEK-x y'EK-x and dist (x, K) = dist (0, K - x), translating K to K - x if necessary, we may assume that x = O. So we must show that there is a unique element of K of minimal norm. Put d = dist (0, K) = inf IIYtI. yEK Since d > 0, it follows that there exists a sequence Yn E K with IIYnll  d. Let us show that {Yn} is Cauchy in K. For this, we first note that for any u, v E K, the parallelogram law gives 2 lIull 2 + IIvll 2 2 u + v 2 lIull 2 + IIvll 2 _ d 2 2 < 2 ' (6.75) u-v 2 because (u + v) /2 belongs to the convex set K, and II (u + v) /211 > d. Replacement of'U and v by Ym and Yn respectively in (6.75) shows that IIYm - Ynll 2 < 2(IIYm1l 2 + IIYnI1 2 ) - 4d 2  0 as m, n  00 and therefore, {Yn} is a Cauchy sequence. Consequently, the sequence {Yn} has a limit Xo (as X is complete) and Xo must belong to K, since K is closed. Since the norm is continuous, IIxoll = lim n -+ oo llYn II = d. This proves the existence of Xo E K. Again, the inequality (6.75) shows that any two Cauchy sequences con- verging to the point minimizing the norm on K must have the same limit. Indeed, if Xo,.Yo E K are such that IIxoll = IIYoll = d then (6.75) for Xo, Yo gives that IIxo - Yo II < 0, Le. Xo = Yo as required. - 6.76. Observations. We have the following: (i) We observe that Theorem 6.74 as such may fail if X is not complete. (ii) The existence conclusion of Theorem 6.74 is not true in some normed spaces. Finally, we also note that the uniqueness conclusion of The- orem 6.74 is not valid in normed spaces. For example, consider the Banach space }R2 with supnorm: X = (}R2, II .11(0), where II (Xl , x2)1I00 = max{lxll, I X 21} 
6.7. Orthogonal Projections on Hilbert Spaces. 385 (compare with Example 5.24). Note that there are infinitely many points in the closed convex set K = { (Xl, X2) EX: Xl > I} which are at minimal distance from the origin. Thus, in this case, X does not have the unique closest point property for convex sets. (Hi) From the proof of Theorem 6.74 it may be noted that Theorem 6.74 continues to hold if X is an inner product space X, and K is a nonempty convex subset which is complete in X. (iv) In Theorem 6.74, uniqueness of Xo need not be true for nonconvex sets. For example, consider the Euclidean space }R2 with respect to the standard product. This is a Hilbert space. Choose S = {(Xl,X2) E}R2 : x + x = I} and the point (0, 0) E }R2. Then all points in S are closest points to (0,0). Observe that S is not convex. . 6.77. Example. Let X = }Rn with the standard inner product and K = { X= (Xl,X2,,,,,Xn) EX: takxk = I } , k=l where (al, a2, . . . , an) is a nonzero fixed element in X. Then it is easy to show that n xo=A(al,a2,...,a n )EK, A=l/Ea, k=l is the unique element of minimum norm in K. . Since a subs pace is a convex set, by the definition, we have (see Figure 6.4) the following simple result from Theorem 6.74. 6.78. Corollary. Let X be a Hilbert space, K a closed subspace, and x EX. Then there exists a unique element Xo E K such that IIx-xoll = dist (x, K). Next, we have the following result which characterizes the projection. 6.79. Corollary. Let X be a Hilbert space over the field IF, K a nonempty closed convex subset of X, and x EX. Then a point Xo is a projection of x on K iff ( 6.80) Re (y - Xo, x - xo) < 0 for all y E K 
386 Chapter 6: Inner Product Spaces x . .. Xo o .. Figure 6.4: Unique approximating element on a subspace of a Hilbert space (Here, Xo is the element of K closest to x EX). Proof. '=>': Suppose that Xo is a projection of x on K. Then, by Theorem 6.74, Ilx - xoll = dist (x, K) < IIx - zll = II (x - xo) - (z - xo)11 for each z E K which, by squaring, is equivalent to 2Re (x - Xo, z - xo) < Ilz - xoll 2 for each z E K. Since K is convex, we have y. = (1 - A)Xo + AY E K for all A E [0, 1] and for arbitrary y E K. Then y. - Xo = A(Y - xo), and thus the last inequality applied to y. (in place of z) gives 2ARe (x - Xo, Y - xo) < A 2 11y - xol1 2 for each y E K. , This inequality is true for all A in [0, 1]. In fact, the last inequality is possible only under two conditions: A = 0 (but then Xo E K) and A > 0; in the later case dividing by A and then letting A  0+ the last inequality gives Re (x - Xo, Y - xo) < 0 so that (6.80) holds. '{::': Suppose that the inequality (6.80) holds for some point Xo E K and for every y E K. Then (see Figure 6.5) lIy - xll 2 - lI(y - xo) - (x - xo)1I 2 - lIy - xoll 2 + IIx - xoll 2 - 2Re (y - Xo, x - xo) > IIx - xoll 2 , for every y E K, and the conclusion follows from the definition of projection. . 
6. 7. Orthogonal Projections on Hilbert Spaces 387 y o H I  z Figure 6.5: Best approximating element on a convex set in a real Hilbert space 6.81. Example. Let X =}R3 and K = span {Vl,V2}, where VI = (1, 1,0) and V2 = (0, -1,2). Then the set {VI, V2} is linearly independent. A typical element (a, b, c) in K can be obtained from solving (a, b, c) = a(l, 1,0) + (3(0, -1,2) which gives a = a, b = a - /3, c = 2{3 = 2(a - b). Therefore, we can write K = {(a, a - b, 2b) E }R3 : a, b E }R}. Choose, for example, x = (1, 1, 1) E }R3 \K. Then, the condition for the vector x - xo = (1 - a, 1 - a - b, 1 - 2b) to belong to K J.. is given by (i) (x - XO,Vl) = 0 (ii) (x - Xo, V2) = 0 (Note that (i) and (ii) imply that (x - Xo, aVl + (3v2) = 0 for a, (3 E ]F). The conditions (i) and (ii) are equivalent to (1 - a) · 1 + (1 - a - b) · 1 + (1 - 2b) .0 = 0, Le. 2a + b = 2, and (1 - a) · 0 + (1 - a - b) · (-1) + (1 - 2b) · 2 = 0, Le. a - 3b = -1, respectively. Solving these two equations, we get a = 5/7 and b = 4/7. Hence, the projection of x = (1, 1, 1) E }R3 \K on K is given by Xo = (a, a - b, 2b) with a = 5/7, b = 4/7; that is Xo = (5/7,1/7,8/7). . 
388 Chapter 6: Inner Product Spaces K x = k + kJ.. .... k o Figure 6.6: Orthogonal decomposition of x E IR2: x = k + k.L z ....... .............1 ( x, y, z) y (x,y,O) x Figure 6.7: Orthogonal projection of (x, y, z) E IR 3 in the XY-plane 
6.7. Orthogonal Projections on Hilbert Spaces 389 One of the important consequences of Theorem 6.73 is that a closed subspace K of a Hilbert space X and its orthogonal complement K.L de- compose the Hilbert space X in the following sense: K together with K.L span the whole space X (see Figure 6.6 and 6.7). 6.82. Corollary. Let K be a closed subspace of a Hilbert space X. Then X = K E9 K.L; that is each x E X admits a unique representation (6.83) x = k + k.L with k E K and k.L E K.L. Proof. Let x E X be arbitrary. Clearly K ..L K.L. Since KnK.L = {O}, it suffices to show that X = K + K.L. By Theorem 6.74, there exists a unique element k E K of minimal norm in K and, by Theorem 6.73, we then have x - k E K.L. Thus x = k + k.L with k E K and k.L = x - k E K.L is a decomposition of the required form (if x E K, then k = 0). Further- more, k and k.L are determined uniquely by x. Indeed, if there was a second such decomposition, x = m + m.L with m E K and m.L E K.L then, by equating the two representations, we would have m - k = k.L - m.L E K n K.L = {O} which would then imply that m - k = 0 = k.L - m.L and hence, the decom- position is unique. - 6.84. Example. Consider the Hilbert space 1 2 with respect to the inner product defined in 6.33. Let K = span {e2} := {a{O, 1,0,0,...} : a E IF} and M = {z = {Zn}nl : Z2 = OJ. It follows that M = K.L and X = K E9 K.L, since each {zn} n 1 E 1 2 admits a unique representation {0,Z2'0,0,...} + {Zl,0,Z3,Z4,.. .}. . We have the following important points to remember. . If X is a Hilbert space and K is a closed subspace, then K is auto- matically a complete space. . Corollary 6.79 continues to hold if X is an inner product space and K is a complete subspace of X. 
390 Chapter 6: Inner Product Spaces . If X is an inner product space and K is not complete, then Corollary 6.79 may fail to hold even when K is a closed subspace of X, see 6.92. . If K is a closed subspace of an inner product space X, K :j:. K..L..L and x E K..L..L \K, then there exists no best approximation to x from K. Indeed, if there exists a best approximating element Xo E K to x then, from Theorems 6.73 and 6.74, Xo must satisfy the condition x - Xo E K..L. Now X-Xo E K..L..L as x, Xo E K..L..L. Thus, we arrive at a contradiction that x - Xo = 0 (Specific example of a closed subspace K such that K :j:. K..L1. is given in 6.92). . 6.85. Remark. Corollary 6.79 may also be used directly to obtain that x - k E K..L. Indeed, for each x EX, Corollary 6.79 gives that Re (w, x - k) < 0 for every w E K, where k E K is a projection of x on K. Because K is a subspace, the same is true when w is replaced by -k, ik, -ik. This observation yields that (k,x-k)=O, Le.x-kEK..L. . The equation (6.83) is known as the orthogonal decomposition of the Hilbert space X. The element kin (6.83) is called the orthogonal projection of x onto K, and the map T :)( E9 K..L  X defined by T(k, k..L) = k + k..L is an isomorphism of K E9 K..L onto X. Corollary 6.82, which establishes an essential geometric fact about Hilbert spaces, is also called the Projection theorem. We know that for a subspace K of a Hilbert space X (see Proposition 6.60(ii) ) K c K..L..L := (K..L)..L. However, as an immediate consequence of Corollary 6.82, we have K..L..L C K whenever K is closed. Indeed to show that this direction of containment result holds, we consider an arbitrary element x in (K..L)..L. Then, Corollary 6.82 guarantees the existence of k E K and k..L E K..L such that x = k + k..L . Since x is orthogonal to every vector in K..L, we have (x,k1.) = o. Therefore, o = (k + k..L, k..L) = (k, k..L) + (k..L, k..L) = IIk..L112. 
6.7. Orthogonal Projections on Hilbert Spaces 391 Hence, kJ.. = 0 which gives x = k E K; that is K J..J.. C K and hence, K J..J.. = K for the closed subspace K of the Hilbert space X. More precisely, the above discussion yields the following property of orthogonal complements in Hilbert spaces. 6.86. Corollary. Let K be a subspace of a Hilbert space X. Then K is closed iff K J..J.. = K. In particular, K = K J..J.. and K J.. = ( K ) J.. (see Proposition 6.60( iv) ). The projection theorem provides the following simple characterization for dense subspaces of a Hilbert space in terms of the orthogonality condi- tion. 6.87. Corollary. Let K be a subspace of a Hilbert space X. Then KJ.. = {OJ iff K is dense in X. Proof. '=>': Assume that the zero vector is the only vector orthogonal to K. Since ( K )J.. = K J.., Corollary 6.82 implies that K = K E9 {OJ = K E9 ( K )J.. which equals X, since K is closed subspace of X. Hence, K is dense in X. '{::': To prove the converse part, we assume that K = X and let kJ.. E K J... Then kJ.. ..L k for all k E K, and because ( K ) J.. = K J.., it follows that kJ.. ..L k for all k E K . But, by assumption, K = X and consequently, kJ.. ..L k for all k E X which gives k ..L k. This means that (k, k) = IIkl1 2 = 0 so that k = O. Hence, KJ.. = {OJ. . 6.88. Example. We see that C[a, b] is dense in £2[a, b] with £2_ norm. By Corollary 6.87, it follows that if f E £2[a, b] such that o = (f, g) = l b f(t)g(t) dt for all 9 E C[a, b], then f E C[a, b]J.. = {OJ so that f(t) = 0 in [a, b]. . 6.89. Corollary. If K is a proper closed subspace of a Hilbert space X, then K J.. contains a nonzero element. Proof. Let x E X\K. Suppose on the contrary that KJ.. = {OJ. Then, by Corollary 6.82, x = k + 0 with k E K and 0 E K J.. which is a contradiction. . In .our next example, we show that Corollary 6.82 is not valid when K is not a closed subspace of a Hilbert space. 
392 Chapter 6: Inner Product Spaces 6.90. Example. We show that there exists a subset K of 12-space such that K + K..L  1 2 . For this, we consider a subset K which consists of all sequences Z = {zn} E 1 2 with Zn = 0 for all but a finite number of n. Then, K is a subspace of 1 2 . With the inner product inherited from 1 2 , K is an inner product space. Consider the sequence {Zl, Z2, . . .} c K where Zn is itself a sequence, say { II 1 } Zn = 1, 2 ' 2 2 ' · · · , 2 n ' 0, 0, · · · E K for each n = 1,2, . ... Then it is a Cauchy sequence, since for n > m IIZn - Zmll = { II 1 } 2 0,0, · · · , 2m+l ' 2m+2 ' · · · , 2 n ' 0, 0, · · . 2 n 1 L 2 2 11: k=m+l 00 1 L 2 2 11: k=m+l 1 1 -  0 as m  00. 2 2 (m+l) 1 - 1/4 < The sequence {Zn} converges to {I, 1/2, 1/2 2 , . . .} E 1 2 \K and, therefore, K is not a closed subspace of 1 2 (Alternatively, for each n E N we can consider W n = { 1,  ,  ' . . . ,  ' 0,0, . . .} E K. Then, the sequence {Wn}n>l of elements in 1 2 converges to W = {1/k }k>l E 12\K as n  00, since - - 2 { II } IIW n - WII2 = 0,0, . · . ,0, 1 ' 1 ' · · · - n+ n+ 00 1 -"  0  m 2 m=n+l and hence K is not closed). Next, we claim that K..L = {OJ. We know that for each k, ek = {0,0,...,1,0,0,...} E K, where ek is the element in the sequence with 1 in the k-th place and zero elsewhere. Suppose that there exists a vector Z = {Zn}nl E K..L. Then, Z ..L ek for each k, and therefore for each k o = (Z, e k) = ({ Z 1 , Z2, . . . , Z k, . . . }, {O, 0, . . . , 1, 0, 0, · · .}) = Z k so that Z = {O, 0, 0, . . .}. Thus, K 1. = {OJ and hence, 1 2  K + K..L so that 1 2 :j:. K E9 K..L (see Corollary 6.82). Also, we observe that, K is a 
6.7. Orthogonal Projections on Hilbert Spaces 393 dense subspace of 1 2 , see Corollary 6.87. Hence, Corollary 6.82 fails to hold if K is not a closed subspace of a Hilbert space X. Since every vector is orthogonal to zero vector, it follows that (K1-)1- = 1 2 :F K. This observation shows that the equality (K 1-)1- = K fails to hold if K is not a closed subspace. However, this equality always holds whenever K is a subspace of finite dimensional space X (because in this case K is automatically a closed subspace). . 6.91. Example. Consider the space C[O, 1] with the standard inner product defined by (6.37), Le. U, g) = 1 1 J(t)g(t) dt, J, 9 E C[O, 1]. Let K be the subspace of C[O, 1] consisting of all differentiable functions on [0, 1]. Then K is a subspace of C[O, 1], but not a closed subspace. First we show that K1- = {OJ. Suppose 4> E K1-. If 4>(c) > 0 for some point c in [0, 1], then because 4> is continuous, 4>(t) > 0 for every t in some interval ( a, b) c [0, 1]. Choose any function 'ljJ E K which is positive in ( a, b) and zero elsewhere. For instance, { sin 2 ( 1r (t - a) / (b - a)) 1jJ(t) = o if t E (a, b) if t E [0, 1] \ ( a, b) will serve our purpose. Then, (,p,1jJ) = 1 1 ,p(t)1jJ(t)dt > 0 with 1jJ E K and,p E KJ.. This contradiction shows that 4> cannot be positive anywhere in [0, 1]. Sim- ilarly, 4> E K1- cannot be negative in [0,1]. This means that K1- = {OJ and C[O, 1]  K E9 K1-. Hence, Corollary 6.82 fails to hold, and since every vector is orthogonal to zero vector, we obtain that (K1-)1- = C[O, 1] :F K. This observation also shows that Proposition 6.68 fails for infinite dimen- sional spaces. . 6.92. No best approximation. Now, we give an example to demon- strate that the completeness of the subspace K plays a crucial role in Corol- lary 6.82 in the sense that Corollary 6.82 is not valid for inner product space 
394 Chapter 6: Inner Product Spaces X unless K C X is a complete subspace. Thus, our aim is to define an inner product space X and construct a closed subspace K (which is not complete) with a property that X :j:. K E9 K.L. This can be achieved by constructing a proper closed subspace K of X with the property that K.L = {OJ. Take X to be the subspace K of 1 2 defined in Example 6.90. Then this X is an incomplete inner product space. With the help of the standard inner product on l2, namely, 00 (Z, W) = L Zk W k, k=1 we choose a special element W = {Ilk} E 1 2 and define M = { z = {ZA:hl EX: (Z,  ) = f:  = O } · k=1 Obviously, M is a nonempty subspace of X. Next, we prove the closedness of M. Let {Zn} be a sequence in M such that {Zn} converges in X where each Zn is itself a sequence. For the sake of convenience we write Zn = {Zn(k)}k1' We need to show that {Zn} converges in M. For this, we use the standard procedure: (i) Find a candidate limit Z; (ii) Show that Z E M and that Zn  Z. Let Zn  Z in X and let N 1 E N be such that Zk = 0 for k > N 1 . Then, for m > N 1 , we have f z; = f ZA: - Zn( + zn(k) k=1 k=1  Zk - zn(k)  zn(k) < Li k + Li k k=1 k=1 ( m ) 1/2 ( m ) 1/2 < :2  IZA: - Zn(kW + ( 00 ) 1/2 m <  :2 IIZ - Znll +  Znk) 111 -  IIZ - Znll + L Znt) · v6 k=1 f Znk) k=1 
6.7. Orthogonal Projections on Hilbert Spaces 395 Since Zn  Z in X, for a given € > 0, there exists an N E N such that IIZn - ZII < € whenever n > N. Since ZN = {zN(k)}kl E M has the property f: ZNk) = 0, k=l there exists an N 2 E N such that  zN(k) L....J k < € whenever m > N 2 . k=l It follows that for m > max{N 1 ,N 2 } 00 m  Z k k _ _  Z k k 1r L....J L....J < v'6 € + € k=l k=l which implies that 00 L  = O. k=l Therefore, the candidate limit Z = {Zk}k>l belongs to M and hence, M is closed. ,- As in Example 6.90, we show that M.L = {OJ. Suppose that there exists a vector Z = {Zn}nl E M.L. For each n, define En = {En(k)}kl as follows: 0 if k :F nand k :F n + 1 En(k) = n ifk=n -n-l ifk=n+l that is El - {El(k)}kl = {1,-2,0,0,0,...} E 2 - {E 2 (k) }kl = {O, 2, -3,0,0, . . .} E3 - {E 3 (k) }kl = {0,0,3, -4,0,...} En - {En(k)}kl = {O,...,n,-n-l,O,...} Clearly, En belongs to M. As Z ..L En for each n, we must have 0= (Z,E n ) = ({Zl,Z2," .}, {O,... ,n, -n -1,0,.. .}) = znn - Zn+l(n + 1) 
396 Chapter 6: Inner Product Spaces so that n Zn+l = 1 Zn for each n. n+ This equation, in particular, implies that Zn :j:. 0 for each n even if Zj :j:. 0 for one j. In this case, this contradicts the fact that Z ..L M c X and therefore, we must have Z = {O, 0, 0,.. .}. Thus, MJ.. = {OJ, Le. M €a MJ.. = M :j:. X. Hence, if M is a closed subspace of an inner product space X then it is not necessarily true that X = M E9 M J.. . From Theorems 6.74 and 6.73, we recall that if K is a closed subspace of a Hilbert space X, then, for each x EX, there exists a unique Xo E K such that x - Xo E K J.. , or equivalently IIx - xoll = dist (x, K). The unique nearest element Xo appearing in Theorem 6.74 is the best ap- proximation to x on K, and is called the projection of the element x E X \K on K. We often write Xo = Px or simply by PKX. When K is a closed subspace of Hilbert space X, the corresponding operator defined by PK : X  K, x I-t k, is a linear projection map/operator from X into K: PK(ax + (3y) = aPKx + /3PKY for every a, (3 E 1F and for each x, Y EX. With this, k = PKx is called the orthogonal projection of x on K, since x=k+(x-k) withkEKandx-kEKJ... The map PK : X  K, x I-t k, is called the orthogonal projection of X onto K, in short, the orthoprojectorof X on K (Remember that x = k+kJ.. with k E K and kJ.. E K J.. is the unique representation of x E X = K E9 K J.. ). Such an operator is continuous, since by the Pythagorean theorem, IIxll 2 = IIkll 2 + IlkJ..1I 2 > IIkll 2 = IIP K xIl 2 , i.e. IIPKII < 1 where equality is attained for K :j:. {OJ. In fact, if there exists a k' :j:. 0 with k' E K, then IIPKllllk'll > IIPKk'll = Ilk' II, i.e. IIPKII > 1, so that IIPKII = 1. Also, as PKX E K, it follows that Pkx = PK(PKX) = PK(k) = k = PKX 
6.7. Orthogonal Projections on Hilbert Spaces 397 and therefore, the projection operator satisfies P K 0 P K = P K , Le. Pi< = PK so that P K is an idempotent operator. Hence, we have the following re- sult. 6.93. Proposition. If K is a closed subspace of a Hilbert space X, then the projection operator P K defined by P K : X  K, x t--+ k, is continuous, idempotent and has the operator norm 1. From the definition, we find that PK.LX = k.l. = X - k = (I - PK)X, Le. PK.L = 1- PK, where I denotes the identity operator. Here PK.L = I - PK is called the complementary projection to P K , and this formula leads to an easy proof of the relation K = K.l..l.. Thus, for each x EX, the elements PKX and (I - PK)x are orthogonal. Further, the decomposition x = PKX + (I - PK)X, Le. 1= PK + (I - P K ), is the unique representation of x into elements of K.l. and K.l..l.. In con- clusion, we have 6.94. Theorem. If K is a closed subspace of a Hilbert space X, then for each x E X x = PKX + PK.LX, i.e. I = PK E9 PK.L, where P K and PK.L are the projection map on K and K.l., respectively. The direct decomposition of this theorem may be expanded to a finite number of mutually orthogonal closed subspaces K i (i = 1,2, . . . , n). Now, we consider the converse part of Theorem 6.94. Suppose P : X  X is a projection. Then the range and the null spaces defined respectively by Rp = {x: Px = x} and N p = {x: Px = O} = {x: (I - P)x = x} are clearly closed subspaces of X. Further, if x is an arbitrary element of X then x = Px + (I - P)x, 
398 Chapter 6: Inner Product Spaces where Px E Rp because P(Px) = p2x = Px, and (1 - P)x E Np, because P((1 - P)x) = (P - p2)x = O. This observation shows that if P : X  X is a projection, then x = Rp E9 N p and hence, N p and Rp are algebraically complementary subspaces of X. Moreover, (1 - p)2 = 1 - 2P + p2 = 1 - P so that Q = 1 - P is also a projection. Note that RQ = N p and NO = Rp. The above discussion gives the following basic characterization theorem. 6.95. Theorem. Each closed subspace K of a Hilbert space X is complemented in X iff there is a projection P of X onto K. 6.96. Example. Let X =}R3 and K = span {Vl,V2}, where VI = (0, 1, 1) and V2 = (1,0,2). It is easy to see that the set {VI, V2} is linearly independent. Then a typical element (a, b, c) E K can be expressed as (a, b, c) = a(O, 1, 1) + ,8(1,0,2) which gives a = ,8, b = a, c = a + 2,8 = 2a + b. Therefore, we obtain K = {(a,b,2a+ b) E}R3 : a,b E JR}. H k.l = (a', b', c') E K.l, then for (a, b, 2a + b) E K we must have o = aa' + bb' + (2a + b)c' = a(a' + 2c') + b(b' + c'). As a and b are not related to each other, the last equation implies o = a' + 2c' = b' + c', i.e. a' = -2d, b' = -c', so that K J.. = {( - 2d, -d, c') E JR3 : c' E IR} = span { (2, 1, -I)}. . 6.97. Example. Let X = C3 with the standard inner product and K = {(,1],') :  = OJ. Define P: X  K by P(, 1], () = (0,1], () for all (, 1], () EX. 
6.8. Orthonormal Basis and Bessel Inequality 399 Note that p2(, 1J, () = P(O, 1J, () = (0, 1J, () = P(, 1J, () showing that p2 = P. Also, we observe that (, 1J, () = (0, 1J, () + (, 0, 0) where (0, 1J, () E K and (, 0, 0) E K.1. Further, the operator 1 - P is given by (1 - P)(, 1J, () = (, 0, 0) and is an orthogonal projection onto the space {(, 0, 0) :  E C}. . 6.8 Orthonormal Basis and Bessel Inequality Our main aim in this section is to describe an orthonormal basis in a Hilbert space. To start with, we recall finite dimensional Hilbert spaces such as the Euclidean space }Rn and the unitary space en. For such finite dimensional Hilbert spaces, we have the notion of orthonormal basis; for example, the standard orthonormal basis {el, . . . , en} for }Rn. Indeed, if we define  k = {el, e2, · · . , ek} then, for each k (1 < k < n), k is an orthonormal set. Note that l can be extended to 2, 2 to 3, and finally n-l to n' However, it is not possible to extend n by adjoining to it any other unit vector, say {e}, so that the resulting set  = {el,e2,...,e n ,e} becomes an orthonormal basis for the space }Rn. In this sense, the set n is called a complete orthonormal basis for }Rn. Thus, for any finite dimen- sional inner product space, the idea of complete orthonormal basis is clear. However, the concept of a basis for an infinite dimensional space is a little problematic one since any orthonormal basis of this space will contain an in- finite set of vectors which may be countably infinite or uncountable. Hence, our goal is to find a way to generalize the notion of complete orthonormal basis to arbitrary Hilbert spaces. We recall that to distinguish the ordinary bases from such notions, an ordinary basis (for a finite dimensional space) is called a Hamel basis. Again, a set B of elements is a Hamel basis for a vector space V if these elements are linearly independent and span the whole space V. In Section 6.5, we had defined the meaning of orthonormal basis. More important than orthonormal bases are the complete orthonor- mal bases which are infinite dimensional analogous of the finite dimensional bases. Now, let us first state the precise meaning of the maximal orthonormal set/system in an inner product space where dirnX = 00. 6.98. Definition. Let  = {tPO:}O:EA be an orthonormal set/system in an infinite dimensional inner product space X. We say that the system  is maximal (or complete) orthonormal in X if there is no unit vector tP 
400 Chapter 6: Inner Product Spaces in X that is orthogonal to each cPo:; Le.  is maximal for X if the only member of X which is orthogonal to every cPo:, a E A, is the zero vector: (cP, cPo:) = 0 for all a E A implies cP = O. This notion of complete orthonormality is different from the complete- ness concept in a general metric space. Most of the text books on this topic refer a complete orthonormal set in a pre-Hilbert space X simply as orthonormal basis, and we also adopt the same convention in this text. Again, we note that the notion of orthonormal basis of an inner prod- uct space X is not to be confused with that of a Hamel basis of a finite dimensional vector (sub )space. A Hamel basis {cPo:} is maximal with respect to linear independence because each x E X is uniquely representable as a finite linear combination of the cPo:'s whereas an orthonormal basis is max- imal with respect to being orthonormal (of course, complete orthonormal basis is also a linearly independent set, see Proposition 6.99). Clearly, an orthonormal basis  need not be a Hamel basis and this can happen only if  is infinite. For example, consider subset £00 = {ek E , 2 : kEN} where ek = {O, 0, . . . , 1, 0, . . .} in which 1 appears only at the kth-slot. We may write this ek = {ekm}ml with ekm = 0 for k  m and ekk=l. Clearly, these vectors are linearly independent. Moreover, £00 is an or- thonormal basis for 1 2 . Indeed, for each Z = {Zn}nl E 1 2 and for all k, we calculate 00 (z, ek) = E Zmekm = Zk. m=l Thus, (z, ek) = 0 for all k iff Zk = 0 for all k, Le. when Z = 0 which shows that £00 is a complete orthonormal set for 1 2 . In particular, using the fact that (z, ek) = Zk, we see that 00 00 IIzlI = E I Z nl 2 = E l(z,e n )1 2 . n=l n=l On the other hand, £00 is not a Hamel basis for 1 2 since, for example, the sequence {l/n}n>l cannot be written as a finite linear combination of the ek's. Nevertheless each element of x E 1 2 can be written in the form 00 x = EXkek k=l 
6.8. Orthonormal Basis and Bessel Inequality 401 but such a sum makes no sense in spaces with usual purely algebraic struc- ture because an infinite series requires to be specified with the natural notion of convergence which is a topological concept as the later one uses the topology to allow the infinite linear combinations. To overcome this difficulty we can add a topological structure to [2 and this is what precisely done in Proposition 6.115. Thus, an orthonormal basis in a Hilbert space is a special example of a Schauder basis which we have discussed in Section 5.5. 6.99. Proposition. Let  = {<Po: : a E A} be an orthonormal sys- tem in an inner product space X. Suppose that the infinite sum EaEA cacPa is convergent with scalars Co: in 1F such that EO:EA Co: cPo: = O. Then  is a linearly independent set. Proof. Suppose that EO:EA co:cPa = o. Then, for a given € > 0, there exists a large enough finite set Ao of A such that for any finite subset Al c Ao, we have L co:cPa < €. o:EAl For each {3 E A, we can --e!llarge the set Ao so that <P13 E <I> with {3 E Ao and therefore, the orthogonality condition implies that ( L co:cPo:, cPl3 ) = L ca(cPa, cPl3) = c{3) = c/3. o:EAl o:EAl On the other hand, in view of the CSB inequality, we have IC{31 = ( L Co: cPo: , cP(3 ) < L co:cPa IlcP{311 = L co:cPa < € o:EAl o:EAl o:EAl As € > 0 is arbitrary, we must have c{3 = 0 which holds for all indices {3. . If a basis of a Hilbert space X is finite or countably infinite, then X is usually referred to as a separable Hilbert space (Le. a Hilbert space containing a countable dense set). Most Hilbert spaces of practical interest which arise "naturally" in analysis are separable. But at the same time, we do have "respectable" nonseparable Hilbert spaces in the theory of almost periodic functions which is out of the scope of the present text. Now, we start proving the Bessel inequality. Let {cPk : 1 < k < n} be a finite orthonormal subset of the inner product space X and K = span { <PI , <P2, . . . , <Pn}. 
402 Chapter 6: Inner Product Spaces For x E X \K, as an immediate consequence of Proposition 6.67, we have the finite sum n k = L(x,cPj)cPj E K j=1 which gives the orthogonal projection of x on K so that x - k ..L Y for all y E K, i.e. x - k E K.L. Hence, by the Pythagorean theorem, we have IIxl1 2 = IIx - kll 2 + IIkll 2 which implies that (6.100) n IIxll 2 > IIkl1 2 = L I(x, cPk)1 2 . k=1 This inequality is known as the Bessel inequality (6.100) for finite dimen- sional case. Suppose that {cP1, cP2, . . .} is a countable orthonormal system in an infi- nite dimensional inner product space X. Motivated by the above discussion, for each x EX, we can form an infinite series 00 L cncPn n=1 and, in particular, the series 00 L (x, cPn)cPn. n=1 Now, we ask when this series converges to x, Le. our problem is to ask when can we write 00 x = L (x, cPn)cPn. n=1 We note that the sum such as EaEA' where A is an indexed set, has to be interpreted as an unordered infinite sums in the sense of the discussion of p. 161. However, we will mostly deal only with separable Hilbert spaces (so that the number of terms in the sums will be countable). We start by proving 6.101. Lemma. H  = {cPa}aEA is an orthonormal subset of an inner product space X, then for each x EX, (x, cP) is nonzero for at most a countable number of vectors cP in . Proof. Consider the orthonormal set  and let x EX. Set s = {cP E : (x, </J) # O} 
6.8. Orthonormal Basis and Bessel Inequality 403 and index the elements in S such that, for each n E N, we form Sn = {cP E S: l{x,cP}l> 1I12 }. This set contains at most n - 1 elements, because otherwise Sn would contain n or more elements, say {cPl, cP2, . . . , cPm } with I(x, cPk)1 > IIxl1 2 In for k = 1,2,..., m (m > n). This means that f I{X,cPkW > f I{X,cPk}1 2 > n C'12 ) = IIxll 2 k=1 k=1 which would then contradict the Bessel inequality for finite sum namely (6.100). Therefore, Sn must contain at most n - 1 elements and hence, for each n E N, Sn is a countable set. Again, we note that if cP is an arbitrary element of S, then I(x, cP)1 > 0, and therefore, we can choose large n such that I(x, cP)1 > IIxll 2 In which means that cP E Sn for some n. Consequently, S = U  1 Sn and hence, S being a countable union of countable sets is countable. - Suppose that X is an infinite dimensional space, and  = {cPo:} o:EA, an orthonormal set in X. Then for each x E X, by Lemma 6.101, it is possible to give a meaning to the symbol E(x,cPo:)cPo: ( := E (x,cP)cP ) . o:EA E Since, by Lemma 6.101, the set {a E A : (x, cPo:) :j:. O} is countable, by rearrangement (if necessary) of the members of , we may write 00 E (x, cPo:)cPo: = E (x, cPn)cPn. o:EA n=1 6.102. Lemma. (Riesz-Fisher lemma) Let {cPk}k1 be an or- thonormal sequence in a Hilbert space X. Then the series of the form L:  1 CkcPk (Ck E C) is norm convergent iffL:  l1 c kl 2 converges, i.e. {Ck} E l2. The sum L:  1 CkcPk is independent of the order in which the terms are arranged. Moreover, if {cPO:}O:EA is an orthogonal system then 2 E co:cPo: o:EA = E I c o:I211cPo:I12 o:EA 
404 Chapter 6: Inner Product Spaces and if { cPa} aEA is orthonormal, then one has 2 L CacPa aEA = E Ic a l 2 aEA for arbitrary C a for which EaEA ICal < 00. Proof. By hypothesis, X is complete, and we know that l2 is complete. Therefore, we need only to observe that the appropriate partial sums are Cauchy. Now, if we let Sn = E Z 1 CkcPk and an = EZ=l I C kI 2 , then, for n > m > 1, we have IIsn - smll 2 = an - am which implies that {sn} is Cauchy iff {an} is Cauchy. According to the Cauchy convergence criterion, since X is complete, E  1 CkcPk (Ck E C) converges iff {Sn} is Cauchy. Also, since IR is complete, E  lick 1 2 converges iff {an} is Cauchy. Therefore, the desired conclusion follows from the last two observations. For the proof of the second part, it suffices to show that the conver- gent sum EaEA cacPa is independent of the order in which the terms are arranged. For this, we assume that EaEA Ic a l 2 < 00 and let y = L ca-y cPa-y be an rearrangement of the original series x = L cacPa. aEA Then (6.103) IIx - yll2 = IIxll 2 - (x, y) - (y, x) + Ily112, where IIxl1 2 = Ily112. For any finite set Ao, set SAo = E CacPa, aEA o tAo = E Ca-y cPa-y aEAo so that the continuity of the inner product gives (x,y) = lim(SAo,tAo) = E Ic a l 2 = (y,X) aEA and therefore, (6.103) shows that IIx - yll = O. Le. x = y. . 6.104. Theorem. (Bessel's Inequality) Let  = {cPa: a E A} be an orthonormal subset of an inner product space X. Then for each x E X 
6.8. Orthonormal Basis and Bessel Inequality 405 we have the Bessel inequality3° (6.105) L l(x,<Pa)12 < Ilx11 2 . aEA Moreover, EaEA (x, <Pa)<Pa is a convergent sum with the limit x and x - x E .L. More generally, if x, yare two arbitrary elements in X then we have (6.106) L I(x, <Pa)ll(y, <Pa)1 < Ilxllllyll. aEA Proof. We have already proved the Bessel inequality (6.105) for finite dimensional case. However, a direct proof for this finite dimensional case follows from Proposition 6.67. Indeed, if { <P1, <P2, . . . , <Pn} is an orthonormal set and if K = span { <P1 , <P2, . . . , <Pn} then, for each x E X and for all scalars C1, C2, . . . , Cn, we have 2 n X - LCk<Pk k=1 n n n - IIxl1 2 - LCk(X,<Pk) - LCk(<Pk,X) + Ll c kl 2 k=1 k=1 k=1 - IIxll 2 - 2Re [  Ck (X,4>k) ] +  I C kl 2 n n - L I(X,<Pk) - ckl 2 + IIxl1 2 - L I(x,<Pk)1 2 > O. k=1 k=1 Note that x and <Pk are fixed whereas Ck'S are allowed to vary such that ECk<Pk E K. Thus, Ilx - E=1 Ck<Pkll has its minimum when Ck = (X, <Pk); that is n (6.107) dist (x, K) = inf x - '" Ck<Pk - Ck ' sEF L....J k=1 n X - L (x, <Pk)<Pk k=l In particular, if we choose Ck = (x, <Pk), it follows from the last inequality that n Sn = L I(X,<Pk)1 2 < IIxI12 k=l (see also (6.100)) and the equality sign in this inequality holds iff x E K. The Bessel inequality for infinite sum, namely 00 L I (x, <Pk) 1 2 < IIx1I2, k=1 30When the index set is (infinite) not denumerable, then the inequality (6.105) shows that the set of indices 0 with l(x,4>o:)1 > l/n must be finite for each n > o. Hence, Ax = {o E A: (x, 4>0:) f:. O} is at most countable, Le. denumerable union of finite sets. 
406 Chapter 6: Inner Product Spaces follows on letting n  00. Indeed, the sequence {Sn}n1 of partial sums is bounded above by IIx1l 2 . Since {Sn}n1 is a monotonically increasing sequence of nonnegative real numbers which is bounded above by IIx11 2 , it converges to a finite sum and therefore, 00 lim Sn =  I(x, tPk)1 2 < Ilxll 2 . n-+ 00 L....J k=1 In fact, for a given x E X, by Lemma 6.101, the set {a E A: (X,tPOl)  O} is countable, we can give a meaning to the sum E I(x,tPOl)12. OlEA If A is countably infinite, we may thus take a typical orthonormal sequence to be indexed by N. Therefore, we can simply write E I(x, tPOl)1 2 = E I(x, tPOl)1 2 OlEA OlEN so that the index form of the Bessel inequality holds for any subset J of A (An important point here is that the order of absolutely convergent series does not matter since any particular enumeration of A would simply yield the rearrangement of the series. Hence, by the definition of the summability the desired inequality (6.105) is immediate for a E A). Now, the convergence of x = EOlEA (x, tPOl)tPOl follows by an application of Lemma 6.102. Further, for tP/3 E , the continuity of the inner product gi ves (x - x, tP/3) = (x, tP/3) - E (x, tPOl)(tPOl, tP/3) = (x, tP/3) - (x, tP/3) = 0 OlEA so that x - EOlEA (x, tPOl) tPOl E .L. Finally, since (6.105) holds also for y E X with A = N, by the CSB inequality for the 12-space, we obtain 00 ( 00 ) 1/2 ( 00 ) 1/2  I(x , 4>k}ll(y, 4>k}1 <  I(x, 4>k}1 2  I(y, 4>kW < IIxlillyll and therefore the index form (6.106) follows because all but countably many of the terms of the left hand side of the inequality (6.106) are zero. _ Equation (6.107) is equivalent to the following result. 
6.8. Orthonormal Basis and Bessel Inequality 407 6.108. Corollary. H {cPo: : 0: E A} is an orthonormal set in a Hilbert space and {co:} Al is an arbitrary sequence of scalars, where A 1 is a finite subset of A, then x - L co:cPo: > x - L (x, cPo:)cPo: . o:EAl o:EAl Let us now return to the problem of writing (6.109) 00 x = L (x, cPn)cPn n=1 where {cP1, cP2, . . .} is an orthonormal system of the inner product space X and x E X. First, we observe that (6.109) means n X - L (x, <Pk)cPk  0 as n  co. k=l Secondly, Bessels's inequality and Lemma 6.102 between them ensure that the right hand side of (6.109) is a convergent series. However, without additional assumption on {cPk}, we are not sure whether the limit of the series is x. 6.110. Corollary. (Parseval Relation) Let {cPo: : 0: E A} be an orthonormal family of vectors of the Hilbert space X. Then IIxll 2 = L I(x, cPo:) 1 2 <==} X = L (x, cPo:)cPO:' o:EA o:EA The relation in Corollary 6.110 is also called Parseval's identity. The proof of Corollary 6.110 is a special case of Proposition 6.115. 6.111. Generalization of 12-space. Let A  0 be an arbitrary in- dexed set, and let 12(A) be the space of all complex valued functions on A such that {If(0:)12}O:EA is summable; that is, 12(A) = { {aO:}O:EA : f(o:) = ao:, for 0: E A and L lao:I2 < CO } o:EA with the meaning that for each given a = {ao:}, the set S = {o: E A : ao: # O} is at most countable. Thus, 12(A) is linear subspace of 12(N), and for a = {ao:}, b = {bo:} E l2(A), {ao:bo:} is a summable family. If we define (a, b) = L ao:bo:, o:EA 
408 Chapter 6: Inner Product Spaces then 12(A) becomes an inner product space, and that 1 2 is 12(N). If f(o:) = e a is the function on A, where e a = 1 for 0: E Sand 0 otherwise, then it is easy to see that {e a } is an orthonormal set. If A is uncountable (eg. A = IR, or C, or (a, b) with a < b), then {e a } is obviously uncountable. 6.112. Corollary. (Riemann-Lebesgue lemma) Let {cPn : n E N} be an infinite orthonormal family of vectors in an inner product space X. Then for any x E X, Urn (x, cPn) = O. n-+oo Proof. From the Bessel inequality, the series E  1 I (x, cPn) 1 2 converges. Now, the desired conclusion follows from the fact that the n-th term of a convergent series approaches to zero as n  00. . 6.113. Theorem. Let X be a Hilbert space and  = {cPa : 0: E A} be an orthonormal system in X. Then the following statements are equivalent: (i)  is an orthonormal basis: x ..L cPa for every 0: E A implies that x = 0 (ii) The linear span of  is dense in X. Proof. Let K I span  = span {cPa : 0: E A}. Then the closure K of span  is closed in X. (i) => (ii): If the closure K of K were a proper subspace of X, then K # X so that we have the decomposition (see Corollary 6.82) X = K fB KJ.. with KJ.. # {OJ. Thus, if 0 # x E K J.. then x ..L cPa for every 0: E A so that  could not have been maximal (as  U {x/llxll} becomes an orthonormal basis), and hence (i) fails. Therefore, we must have K = X, Le. K is dense in X. (ii) => (i): Suppose (x, cPa) = 0 for each 0: E A. Now, it is clear that x ..L Ko for each Ko = span {cPa: 0: E Ao} and for each finite subset Ao of A. Thus, x 1.. K. Therefore, the continuity of the inner product shows that x is orthogonal to the closure K , which is X. In particular, x ..L x and hence, x = O. . 6.114. Example. Let us demonstrate the fragile nature of the com- pleteness of the orthonormal system by an example. (i) Consider (see Example 6.58) the orthonormal set S = {v'2 COS 27rnt: n > 2} U { v'2 sin 21rnt : n > I}. Then v'2 cos 27rt ..L S and therefore, SJ.. # {OJ. This observation shows that S is not an orthonormal basis for C[O, 1]. 
6.8. Orthonormal Basis and Bessel Inequality 409 (ii) Consider the system c) = {  Lez for the Hilbert space LH-1I",1I"] with the standard inner product defined by (6.36). The system  forms an orthonormal basis for L[ -1r, 1r] because the linear span K of , K=span{  : kEZ}, of the polynomials EZ=-n  eikt, n E N, is dense in L[-1r, 11"]. . 6.115. Proposition. Let  = {tPa : a E A} be an orthonormal system in an inner product space X. Then the following statements are equivalent: (i) The linear span of  is dense in X, i.e. every vector in X is a limit of a sequence of vectors from this span. (ii) Parseval's relation IIxll 2 = E I(x, tPa)12 aEA holds for each x EX. (Hi) For each x EX, x = E (x, tPa)tPa aEA as a norm convergent series. (iv) Plancherel relation (x,y) = E (x, tPa)(Y, tPa) aEA holds for each x, y EX. Proof. For an arbitrary finite subset S of A, set Xs = E(X,tPa)tPa and K = span{tPa: a E S}. aES Then, for each given x EX, we have (x - xS,y) = (x - E (x,tPa)tPa,Y) = 0 for all Y E K . aES which shows that Y ..L x - Xs and, by orthonormality (see Lemma 6.102), 2 E(X,tPa)tPa aES = E I(x,tPa)12. aES 
410 Chapter 6: Inner Product Spaces In particular, the map PK : x I-t L (x, cPOl)cPOl OlES is the orthogonal projection of X onto K and (6.116) IIxll 2 = L I(x, cPOl)1 2 + IIx - PKxl1 2 OlES holds for each x EX. Now we are in a position to prove (i)  (ii)  (iii)  (i,,)  (ii) (i)  (ii): Let x E X ,and € > 0 be given. By (i), there exists a finite subset Al of A such that for all y = EOlEAl aOlcPOl E K = span {cPOl : a E AI} we have IIx - yll < €. Thus, by (6.116), o < IIxll 2 - L I(x, cPOl)1 2 = IIx - PKxll 2 = dist (x, K) < IIx - yll2 < €2. OlEAl Here € is arbitrary and therefore, (ii) follows as €  0 and Al  A. (ii) => (iii): By (6.116), for every x EX, and for each finite subset Al of A we have 2 x - L (x, cPOl)cPOl OlEAl - IIxll 2 - L 1 (x, cPOl) 1 2 OlEAl - L I(x, cPOl)1 2 - L I(x, cPOl)1 2 , OlEA OlEAl L I(X,cPOl)12. OlEA \Al by (H), If this being so, the equivalence of (ii) and (iii) is immediate (in fact, as in Lemma 6.101, the set {cPOl : (cPOl, x) :j:. O} is countable so that {(x, cPOl)cPOl} is summable to x, and hence the above equation can be written as n X - L(X,cPk)cPk k=l 2 00 = L I(x, cPk)1 2 k=n+l and the equivalence of (ii) and (Hi) follows from the last equation and from the definition of the norm convergence). (Hi)  (iv): Let x = L (x, cPOl)cPOl OlEA and y = L (y, cP/3)cP/3' /3EA 
6.8. Orthonormal Basis and Bessel Inequality 411 The continuity of the inner product gives (x, y) = E E (x, cPa )(y, cP/3) (cPa, cP/3) aEA /3EA - E (x, cPa ) (y, cPa) aEA and (iv) holds. (iv) => (ii): This part follows if we take x = y in (iv). So to prove (iv) => (i), it suffices to prove (ii) => (i). In fact, by Theorem 6.113, it is enough to show that if (ii) holds, then  is an orthonormal basis. Suppose this is not true. Then there must exist a nonzero vector x such that x ..L K, (x, cPa) = 0 for all a E A, where K = span {  }. Substituting this x into the Parseval relation, it follows that IIxl1 2 = E I(x, cPa)12 = 0 aEA which is a contradiction. Consequently,  should be an orthonormal basis and this completes the proof. _ Our important point here is that if  = {cPa : a E A} is an orthonormal basis of a Hilbert space X, then each x E X can be expressed uniquely in the form x = E cacPa, C a = (x, cPa). aEA The right hand side expression is sometimes called an abstract Fourier expansion or Fourier series of the vector x in this system. The scalars Co: = (x, cPa) are called the components of x with respect to the orthonormal basis , and are also called Generalized Fourier coefficients of x relative to the orthonormal basis . In the discussion of Fourier series, we use the trigonometric system as an example of orthonormal basis to develop the theory of Fourier series for periodic functions. However, examples of other orthonormal bases that are widely used in practice are Bessel functions, Chebyshev polynomials, Harr functions, Hermite polynomials, Jacobi polynomials, Laguerre polynomials, Legendre polynomials, Rademacher functions and Walsh functions. In Sec- tion 6.10, we provide an application of uniform boundedness principle to prove that there exists a continuous periodic function whose Fourier series diverges. Theorem 6.113 and Proposition 6.115 together give characterizations for orthonormal basis of a Hilbert space. 6.117. Theorem. Let X be a Hilbert space and  = {cPa : a E A} be an orthonormal system in X. Then the following statements are equivalent: 
412 Chapter 6: Inner Product Spaces (i) The system  is an orthonormal basis (H) The closed linear span of {cPOl : a E A} is X (Hi) For each x E X, x = L (x, cPOl)cPOl OlEA (iv) For each x EX, IIxll 2 = L I(x, tPOl)1 2 OlEA (v) For each x, y EX, (x, y) = L (x, cPOl) (y, cPOl). "- OlEA 6.118. Corollary. Let X be a Hilbert space, x E X and let K b a closed subspace of X. If  = {cPn : n E N} is an orthonormal basis in K, then the best approximation to x in X is 00 k = L(X,cPn)cPn. n=l Proof. By hypothesis, K is a closed subs pace of a Hilbert space X and hence, K itself is a Hilbert space with respect to the restriction of the inner product on X. Moreover, by Corollary 6.82, each x E X admits a unique representation x = k + kJ.. with k E K and kJ.. E K J.. . By Theorem 6.117, k E K has the representation 00 k= L(k,cPn)cPn. n=l Since (kJ.., cPn) = 0 for all n, it follows that (x, cPn) = (k + kJ.., cPn) = (k, cPn) + (kJ.. , tPn) = (k, cPn) for all n, and hence 00 k = L (x, cPn)cPn n=l as required. . 6.119. Example. Let {cPn : n E N} be a complete orthonormal family of vectors in an inner product space X. We show that a linear operator T on X defined by 00 Tx = L 3- k (x, cPk+l)cPk k=l 
6.9. Cardinality Theorems for Orthonormal Bases 413 is in B(X). To verify the boundedness of T on X, we compute the operator norm IITII. Indeed, by the Parseval identity, we get 00 IITxl1 2 = E 3- 2k I (x, tPk+l) 1 2 k=l 00 < 3- 2 E I(x, tPk+l)1 2 k=l < IIxIl 2 /9, by the CSB inequality, so that IITII < 1/3. Taking x = tP2, we conclude that IITxll = IItPl/311 = 1/3, and hence, IITII = 1/3. . 6.9 Cardinality Theorems for Orthonormal Bases In view of Zorn's lemma, we can obtain that "every nonzero Hilbert space X has an orthonormal basis." Indeed, if we consider the class of orthonormal sets in X with a partial order defined by inclusion then, by Zorn's lemma, there exists a maximal orthonormal set, say K. Since K is maximal, it is complete. By Corollary 6.89, the closed linear span of K is X. It follows that K is an orthonormal basis. How large can an orthonormal set be in a separable Hilbert space? The following result shows that "each orthonormal set in a separable Hilbert space is at most countable." 6.120. Theorem. (Existence theorem for orthonormal bases) If X is an infinite dimensional Hilbert space, then the following three con- ditions are equivalent: (i) X is separable (H) X contains a countable orthonormal basis (Hi) Every orthonormal system is countable. Proof. We prove "(i) => (iii)" and leave the remaining assertions as an exercise. Let X be a separable Hilbert space and let  = {tPo: : a E A} be an orthonormal system in X. Then, for a :j:. {3, IltPo: - tPtill 2 = (tPo: - tPti, tPo: - tPti) = IItPo:II2 + IItPtill 2 = 2. Thus, each pair of distinct elements in  are a distance V2 apart. Let B denote the collection of open balls with center tPo: and radius 1/2: B = {Eo: = B(tPo:; 1/2) : a E A}. Clearly, the balls Eo: are disjoint which implies that each such ball must contain a distinct point in a countable dense subset. Thus, if  were not 
414 Chapter 6: Inner Product Spaces countable then every dense subset of X would be uncountable which is a contradiction. _ Another useful theorem is a consequence of the existence of orthonormal bases in separable Hilbert spaces. This result provides a useful tool in identifying all separable Hilbert spaces since it permits us to represent each element of the space as a unique linear combination of elements of the corresponding orthonormal basis. Now, we can ask whether there exists any simple relationship between the elements of two separable Hilbert spaces. 6.121. Theorem. (Equivalence of Hilbert spaces) Any two infi- nite dimensional separable Hilbert spaces X and Y (each over IF) are iso- metrically isomorphic. In particular, every infinite dimensional separable Hilbert space is isometrically isomorphic to the space 1 2 . Proof. Let {tPn}nl and {1Pn}nl be two orthonormal bases for X and Y, respectively. Given a point x EX, we may write 00 x = E antPn, an = (x, tPn), n=l and, by the Parseval identity, we find that IIxll 2 = E  1 la n l 2 < 00. Now we show that the correspondence between the spaces X and Y is established by the map T : X --t Y defined by 00 Tx = E a n 1Pn. n=l Clearly, T is linear. As {1Pn}nl is the orthonormal basis for Y, we get Tx = 0 => an = (x, tPn) = 0 for all n => x = 0 which shows that T is one-to-one. Again, by the Parseval identity, it follows that 00 IITxll 2 = Elan 1 2 = IIxll 2 n=l which proves that T is an isometry. It remains to verify that T is onto. Suppose that y E Y is given. Then, we can write 00 y = E b n 1Pn, b n = (Y,1Pn), n=l and, since E  1 Ib n l 2 < 00 by the Parseval identity, the series E  1 bntPn must converge to some point x' EX. Then, for each n = 1,2, . . ., we have ( fbkrPk,c/Jn ) = b n for m > n. k=l 
6.9. Cardinality Theorems for Orthonormal Bases 415 By the continuity of the inner product, this equality yields a = (x', cPn) = oo ( f bkl/Jk, l/Jn ) = b n . k=l That is, we have 00 00 Tx' = L a1/Jn = L bn'l/Jn = Y n=l n=l which means that each element y E Y corresponds to some x' E X according to the formula for T and therefore, T is onto. - A proof similar to the above implies that every n-dimensional Hilbert space 1F is isomorphic to r and we leave it as an exercise. Our final result of this section deals with the cardinality of complete orthonormal system of a (separable) Hilbert space. 6.122. Theorem. Any two orthonormal bases in a separable Hilbert space X have the same cardinal number. . Proof. Let  = {cPOl : a E A} and q, = {1/J/3 : {3 E f} be two arbitrary orthonormal bases for X. H A or f is finite, then both have the same cardinality which is nothing but the (linear) dimension of X. Suppose that A and f are infinite sets. Then we consider the dense set S that consists of finite linear combinations, namely the set {b{31/J{3: for each finite set J c r } " where the coefficients b/3 must be chosen in the following way: If IF = C, then bl3 = c/3 + id/3, c/3, d/3 E Q, and if IF = JR, then we assume d/3 = O. We use a standard result (which is not an easily verified statement) that card (S) = card (f). Recall that for a :j:. a', 2.;2 lIl/Ja -l/Ja,1I = V2 > 3" . If B denote the collection of open balls with center cPol and radius 1/2, then the balls B(cPOl; 1/2) are disjoint which implies that for each cPol' there exists an element SOl in S such that .;2 IIcPOl - sOl11 < 3. 
416 Chapter 6: Inner Product Spaces Thus, there is a one-to-one correspondence between the elements of  and S. Indeed, if sa1 = Sa2 then we have IItPa1 - tPa211 < IItPal - salll + II s a1 - sa211 + II Sa 2 - tPa211 - IItPa1 - sa111 + II Sa 2 - tPa211 V2 V2 2V2 < 3+3=3 which implies that Ctl = Ct2. Hence, card (A) < card (S) = card (f). If we interchange the roles of f and A, the we arrive at card (f) < card (S) = card (A) and conclude that card (A) = card (f) . 6.10 Applications of Uniform Boundedness Principle Theorems like the uniform boundedness principle often useful in producing a continuous function whose Fourier series diverges. In this section, we pro- vide such a typical example as an application of the Uniform Boundedness Principle to the theory of Fourier series. 6.123. Complex Fourier series. The complex Fourier series of 21r- periodic continuous function f : IR  C is given by f(t) ,....., L Ck eikt kEZ where Ck is defined by (6.124) 1 l 1t' . Ck = _ 2 e-1,kx f(x) dx, 1r -1t' , k E Z, and are called complex Fourier coefficients of f(t). The complex Fourier series is also called the Fourier exponential series. Periodic functions are often restricted to (-1r, 1r] or (0, 21r]. In fact, if g(t) is a function which has an arbitrary positive period w (Le. g(t + w) = g(t) for all t E IR), then the complex Fourier series expansion of 9 ( 2 t) has a period 21r. Hence, by a change of variable, the Fourier series for a w-period continuous function 9 is 00 g(t),....., L Ckei()t, k=-oo 1 l c+w . 21fla C k = - g(x)e- t ( w )x dx w c 
6.10. Applications of Uniform Boundedness Principle 417 where C is any real number. Consider the n-th partial sum (snf)(t) of the complex Fourier series of f defined by n (snf)(t) = L Ck eikt , k=-n where Ck is given by (6.124). Note that for each n E N . sn(f) is continuous . sn(f) is 27r-periodic function; Le. sn(f)(t + 21r) = sn(f)(t). An important consequence of the uniform convergence theorem (see The- orem 3.43) is "if the Fourier series of f converges uniformly to f, then f must be continuous on [-7r,7r] with f(1r) = f( -1r)". Then the follow- ing basic questions arise: Is there an important relationship between f(t) and its Fourier series? Does the partial sum sn(f)(t) approximate f(t) for large values of  in some sense? Does the Fourier series EkEZ Ckeikt con- verge to f(t)? We are not aiming to discuss these questions, but wish to show that there exists a continuous 21r-periodic function with a divergent Fourier series. We remark that the problem of deciding whether or not the Fourier series EkEZ Ck eikt converges at a specific point (or everywhere) is difficult as it usually requires some degree of smoothness (differentiability and uniform convergence condition). 6.125. Dirichlet's kernel and its properties. We need a formula for calculating the sum (6.126) 1 n Dn(t) = 2 + Lcoskt. k=l Now, let us see how this finite sum is useful in examining the behaviour of the partial sum of the Fourier series of a periodic function. Actually, it is easy to see that Dn(t) has the following simple form: sin (n + !) t 2 sin ! t 1 2 + n (6.127) Dn(t) = for t :j:. 2m1r, m E Z, for t = 2m7r, m E Z, and this function is called Dirichlet's kernel. Indeed, if t = 2m1r for some m E Z, then, from (6.126), it follows that D n (2m1r) equals ! + n. IT t :j:. 2m1r, then sin! t :j:. 0 so that multiplying both sides of the equality (6.126) by 2 sin! t, we have n 2D ( ) . 1 · 1 " 2 k . 1 n t SIn 2 t = SIn 2 t + L.J cos t SIn 2 t k=l 
418 Chapter 6: Inner Product Spaces n sin  t + :)sin (k +  ) t - sin (k -  ) t} k=l - sin(n+  )t. Thus, (6.127) holds. For an alternate proof of (6.127), we write z = e it and observe that 1 n 1 n ikt + -ikt 1 n 1 2 + L cos kt = 2 + L e 2 e = 2 L e ikt = 2 L zk k=l k=l k=-n Ikln and therefore, we see that the Dirichlet kernel takes an equivalent form (6.128) 1  k Dn(t) = 2 L....J z , Ikln z = e it . Alternatively, for z :j:. 1, we have L zk - Ikln 1 n  (zk+l _ zk) z-l L.J k=-n - Z  1 (  (Zk+l - zk) + k n (Zk+l - Zk)) zn+l - z-n z-l zn+l/2 _ z-n-l/2 Zl/2 - Z-1/2 so that, by substituting z = e it (t :j:. 2m1r, m E Z), we obtain L zk = sin (  ) t Ikln SIn 2 t and the desired conclusion (6.128) (and hence, (6.127)) follows. Let us now formulate some of the preliminary properties of Dirichlet's kernel. We observe that Dn is even. If we substitute t = 2m1r in the right hand side expression in (6.126), we see that 1 D n (2m1rt) = n + 2 and, in particular, 1 Dn(O) =n+ 2 ' 
6.10. Applications of Uniform Boundedness Principle 419 From the right hand side of Dn(t) in (6.127), it follows that the points t = 2m1r (m = 0, 1,.. .), where both the numerator and denominator vanishes, are ,the points of removable discontinuity, because I " D ( ) - 1 . ( l ) cos(n+!)t _ (n+!)(-I)m _ 1 1m n t - 1m n + - - - n + -. t-+2m1l" t-+2m1l" 2 2.!cost (-I)m 2 Moreover, since Dn(t) is even, 1 1 11" 2 1 11" 2 [ 1 1 11" n 1 11" ] - Dn(t)dt=- Dn(t)dt=- - dt+2: cosktdt =1 1r -11" 1r 0 1r 2 0 . k=1 0 and that, by triangle inequality, 1 n 1 IDn(t)1 < 2 + 2:lcosktl = 2 +n. k=1 Thus, the basic properties of Dn(t) may be summarized as 6.129. Lemma. The Dirichlet kernel Dn(t) defined by (6.127) has the following properties: . Dn(t) is even . Dn(t) is a Coo 21r-periodic function . IDn(t)1 < ! +n 1 1 11" 2 1 11" . - Dn(t) dt = - Dn(t) dt = 1. 1r -11" 1r 0 The Dirichlet kernel plays a key role in our next calculation. 6.130. Lemma. Let f : IR  C be a 21r-periodic function and inte- grable on [-1r,1r]. Then, we have 1 1 211" 8n(f)(t) = - f(t + x)Dn(x) dx, 1r 0 where 8n(f)(t) is the n-th Fourier partial sum of f. Proof. Using (6.124), we can rewrite our formula for 8n(f) as follows: sn(f)(t) = n ( 1 1 11". ) . 2: - e-1,kx f(x) dx e tkt 21r -11" k=-n - 2.. r f(x) ( t eik<t-Z) ) dx 21r J -1f' k=-n 
420 Chapter 6: Inner Product Spaces 1 l 1t' - 7r _11" f(x) Dn{t - x) dx, by (6.128), 1 l 1t' -t ). - 7r -1I"-t f(t + u)Dn(-u)du (x - t = u, Dnl u ) is even) 1 (1t' - 7r J-1I" f(t + x)Dn(x) dx. Here both f(t) and Dn(t) are periodic functions of period 21r and therefore, the last equality is a consequence of the fact that for every 21r-periodic function 9 and for each real a, we have i:: g(u)du = i: g(u)du. The desired representation follows. . By Lemma 6.130, the value of the n-th partial sum of the Fourier series of f at 0 is given by 1 1 2 1t' Tn(f) := 8n(f)(0) = - f(x)Dn(x) dx. 1r 0 Now let X denote the Banach space of 21r-periodic continuous functions f : IR  ]R endowed with the supnorm. It should be clear that Tn defines a bounded linear functional on the Banach space X and obtain the following result. 6.131. Lemma. The linear operator Tn : X  IR defined by 1 1 2 1t' Tn(f) = - f(x)Dn(x) dx 1r 0 is bounded, and (6.132) 1 1 2 1t' IITnll = - IDn(x)1 dx. 1r 0 Proof. The boundedness of Tn follows from Example 5.64. Let g(x) be the function which is +1 for those x for which Dn(x) > 0 and which is -1 for those x for which Dn(x) < O. Then, for every € > 0, 9 may be modified to be a continuous function f of norm 1 so that Tn(f) - .! 1 211" IDn(x)1 dx = .! 1 211" (f(x) - g(x))Dn(x) dx < €. 1r 0 1r 0 and hereby we obtain the norm equality (6.132). . 
6.10. Applications of Uniform Boundedness Principle 421 Now, we aim to show that 1 21r IDn(t)1 dt  00 as n  00 and then use the Uniform Boundedness Principle to conclude that there exists a continuous periodic function whose Fourier series diverges at t = O. 6.133. Lemma. We have - 1 21r sin (n + ! ) t In = . 1 2 dt  00 as n  00. o SIn 2" t Proof. Since 1 sin tl < t for all t E JR, we have In > r 21r 2 1 10 t sin (n+ 2 )tdt r(2n+l)1r 1 · 91 - 2 10 SI; d(} (let 0 = (n + 1/2)t) 2n+l j k1r I . 91 - 2 L sm d(} k=l (k-l)1r 9 2n+l 1 j k1r > 2 L k I sin 01 d() k=l 1r (k-l)1r 4 2n+l 1 j k1r - - L k ' since I sin 01 dO = 2. 1r k=l (k-l)1r This, by the definition of II Tn II, implies that 2 2n+l 1 IITnll > 11"2 L k ' k=l Consequently, IITnll  00 as n  00, since L: %" 1 t diverges. . Finally, we are ready to prove the following important result on Fourier series. 6.134. Theorem. There exists a continuous real-valued 21r-periodic function 1 : [0, 21r]  ]R with 1(0) = 1(21r) such that its Fourier series diverges at O. Proof. Let X be the space of all real-valued 21r-periodic continuous functions 1 on IR endowed with the uniform norm: 11/1100 = sup I/(t)l. tE[O,21r] 
422 Chapter 6: Inner Product Spaces It is not difficult to see that X is a closed subspace of the space CR(IR), the space of all bounded continuous functions from IR into III Therefore, X is a Banach space. Now, consider the sequence of bounded linear functionals {Tn} defined in Lemma 6.131 on X. Let f E X. Then, by Lemmas 6.131 and 6.133, we observe that (snf)(O) = EZ=-n Ck fails to converge as n  00. . 6.11 Exercises 6.135. Determine whether the following statements are true or false. Justify your answer. (a) Neither the Cartesian norm Ilzll = max{lzII,..., IZnl} nor the I-norm IIzlh = E j I IZj I, Z = (Zl, · · · , Zn) E r, come from an inner product. (b) For U = (Xl, X2) and v = (YI, Y2) in IR 2 , if we define (u, v) = ax + X2(bxI + CYI) + dy then there exist no real values of a, b, c, d such that the above equation defines an inner product on IR 2 . (c) For U = (UI,U2) and v = (VI,V2) in C2, if we define (u, v) = aUI V I + U2( bv l + CV2 ) + dUI V2 then there exist complex values of a, b, c, d such that the above equa- tion defines an inner product on C2. (What are the resl1riction on these parameters so the above (u, v) defines an inner product?) ( d) For U = (UI, U2) and v = (VI, v) in IR 2 , if we define (u, v) = aUI VI + U2(bvl + CV2) + dUI V2 then there exist real values of a, b, c, d such that the above equation defines an inner product on IR 2 . (What are the restriction on these parameters?) (e) If u, v belong to an inner product space V over C, then U ..L v iff U ..L AV iff AU ..L v, for A E C\ {OJ. (f) In an inner product space V over C, and u, v E V, we have U ..L v iff lIu + J.tvll = lIu - J.tvll for all scalar J.t E C. (g) If, in the unitary space (en , (., .)) with standard inner product defined by (6.6), we have lIu + vII = lIu - vII, then it is not necessary that u ..L v, where u, v E en . (h) The sequence {  L>l in the space Cc[-1I",1I"] with the standard inner produt defined by (6.36), forms an orthonormal set. 
6.11. Exercises 423 (i) The sequences { COB nt } and { sin nt } are orthonormal in the .;:i n>l .;:i n>l space C[0,21r] with respect to the inner product defined by (6.37). (j) In C[-1r,1r] with respect to the inner product defined by (6.37), con- sider the sequence of points {un, Vn}n>l, where Un = sin nt and V n = cos(n - l)t. Then Un ..L V m for all m, n > 1. ( k ) In L 2 [ -1r 1r ] the set {  cos nt sin nt : n E N } forms a basis with , , .;:i , .;:i , .;:i respect to the inner product of (6.37). (I) In an inner product space (V, (.,.)), we have 00 00 U = LUn => L(un,v) = (u,v) for each v E V. n=l n=l (m) If {Zn}nl is an orthogonal sequence in a complex Hilbert space, then {Re Zn, 1m Zn}nl does not necessarily form an orthogonal sequence. (n) H {Ul, U2, . . .} is an orthonormal collection in an inner product space (V, (.,.)), then every U E span{ul,u2,...,U n }, where n is fixed, can be written as n u= L(U,Uk)Uk. k=l (0) In }Rn, (u, v) = Ul Vl + 2U2V2 + 3U3V3 + . · · + nUn V n defines an inner product on }Rn . (p) Let Pn(C) denote the space of all polynomials of degree n on C with complex coefficients. For p(z) = EZ=oPkZk and q(z) = EZ=o qk zk , define n (p, q) = LPk qk . k=O Then (Pn(C), (', .)) is an inner product space (and is in fact a Hilbert space) . (q) For p(z) = EZ=oPkZk E Pn{IF), define n Ilpll = L Ip(k)l. k=O Then, this definition makes the space Pn{IF) a Banach space, but there is no inner product such that (P,p) = IIpI1 2 for all P E Pn(IF). (r) For p(z) = E=oPkZk and q(z) = E=o qkzk in P2(IF), the space of all quadratic polynomials in z E C with complex coefficients, if we define 2 (p, q) = L Pj q k = Pl q l + Pl q2 + P2Ql + P2 q2 j,k=l then (.,.) does not define an inner product on P2 (IF). 
424 Chapter 6: Inner Product Spaces (s) The set S = {el,...,e n }, where ek is the n-tuple with 1 in the k- th place and zero elsewhere, is a complete orthonormal basis for the Hilbert space 2(n). (t) The set S = {ek : kEN}, where ek is the sequence with 1 in the k-th place and zero elsewhere, is a complete orthonormal basis for the Hilbert space 1 2 . (u) In the usual inner product space JR2, if K = {(a, b)}, where (0,0)  (a, b) E }R2, then KJ.. = {(x, y) E }R2 : (x, y) = A(b, -a), A E JR}. (v) The orthogonal complement of the set of all even functions in the standard inner product in C[ -1, 1] is the set of all odd functions in C[-l,l]. (w) If K is a closed subs pace of a Hilbert space X, and if PK denotes the orthogonal projection onto K, then K = {x EX: IIPKxll = Ilxll}. (x) Let B = {x EX: IIxll < I} be the closed unit ball in a Hilbert space X. H {x n } and {Yn} are two sequences in B such that (xn,Yn) -+ 1 as n  00, then lim n -+ oo X n = lim n -+ oo Yn. (y) If K = {f EX: f(t) = 0, t E [a, c] for a fixed c, c < b}, where X = (CF[a, b], II .112), then K J.. = {f EX: f(t) = 0, t E [c, b]}. (z) Each Z E 1 2 has the representation z = L:  1 Zkek, for some {Zk}kl E 1 2 , where ek is the standard orthonormal system in 1 2 . 6.136. Let X be an inner product space and u, v EX. Let Wk = e21rik/n (k = 0, 1, . . . , n - 1) denotes the n n-th roots of unity. Check whether the following equalities hold. n-l (i) (u, v) =  I: wkll u + wk v ll 2 for n > 3 n=O (ii) (u, v) =  1 211" lIu + eillvll2eill dfJ. 21r 0 Note: (i) corresponds to averaging over a cyclic group of n-th root of unity whereas (ii) is a consequence of (i) whenever n  00. 6.137. For U = (Ul, U2) and v = (Vl, V2) in ]R2, define (i) (u, v) = Ul Vl - U2V2 (ii) (u, v) = Ul Vl - U2 V l - Ul V2 + 3U2V2 (Hi) (u, v) = Ul Vl - U2 V l - Ul V2 + 4U2V2 (iv) (u, v) = 2Ul Vl + 3U2V2 (v) (u, v) = 2Ul Vl - U2 V l - Ul V2 + 5U2V2 
6..11. Exercises 425 (vi) (u, v) = 2Ul VI + U2V2. Check whether each of (i) to (v) defines an inner product on }R2. Explain. Note: Compare with Exercises 6.135(d). 6.138. Given an n-dimensional (complex) ellipsoid {  IZk 1 2 } (Zl , Z2, . · · , Zn) E en : L....J  < 1 k=l k for some ak > 0 for all k, introduce an inner product on en so that the ellipsoid becomes the unit ball. 6.139. For I, 9 E Cc[O, 1], define (i) (f, g) = 1 1 If( t) + g(t) I dt. (ii) (f, g) = f(1) g(1) + 1 1 f(t) g(t) dt. (iii) (f,g) = a 1 1 f (t)9(i) dt, wher e a is a fixed nonzero real number. (iv) (I, g) = 1(1/4)g(1/4) + 1(1/2)g(1/2). Determine which of the above defines an inner product on the space Cc[O, 1], and which does not? Explain. 6.140. For I, 9 E C[O, 1], define (I, g) = I(O) g(O) + 1'(O)g'(O) + 1'(1)g'(1). Check whether this defines an inner product on the space C[O, 1] and on the su bspace K = span {1, t, t 2 } , respectively. 6.141. Consider the space Ct[O, 1] with an inner product defined by (i) (f,g) = f(O) g(O) + 1 1 f'(t) g'(t) dt, f,g E Ct[O, 1] (ii) (f, g) = 1 1 f(t) g(t) dt + 1 1 f' (t) g'(t) dt, f, 9 E Ct[O, 1]. Orthonormalize the set of vectors {1, t, t 2 } with respect to each of the above inner products. 6.142. For the inner product define by (6.37) for I, 9 E C[O, 21r], find the mean-square distance for: 
426 Chapter 6: Inner Product Spaces (i) f(t) = 1 and g(t) = sin t (ii) f(t) = sin t and g(t) = cos t. 6.143. Let al, a2, . . . , an be given n positive real numbers. In}Rn, define n (u,v) = Lakukvk (u = (Ul,U2,...,U n ), V = (Vl,V2,...,V n )). k=l Show that (}Rn, (., .)) is an inner product space. Check whether the space an with respect to the above inner product is a Hilbert space. Show that if anyone of the numbers an is zero or negative, then the above definition does not define an inner product on }Rn . 6.144. Let a = {ak}kl' where ak > 0 for each k. Let 12(a) be the space of all sequences Z = {Zk}k>l of complex numbers such that E  1 akl z kl 2 < 00. For Z E 12(a) and w- = {Wk}kl E 12(a), define 00 (z,w) = Lakzk w k. k=l (This inner product is referred as weighted inner product on 1 2 .) Show that 12(a) is a Hilbert space. If anyone of the numbers an is zero or negative, then verify whether the above definition makes the corresponding space to be an inner product space. 6.145. Let w(t) be a fixed positive continuous functions on C[a, b]. Show that the formula (f,g) = i b f(t)g(t)w(t) dt defines an inner product on CF[a, b]. Does it become a Hilbert space? 6.146. Given a real inner product space X, suggest a method of form- ing a complex inner product space. 6.147. Using the standard inner product on C[O, 7r], find the projection of each of sin t and cos t on K = span { t}. 6.148. Let w be any(fixed) cube root of -1, and let K = span {z, (} C C3, where Z = (1,w,w 2 ) and (= (1,w 2 ,w). Determine which point in K is nearest to (1, 1, 1) E C3 . 6.149. Let w be any(fixed) fourth root of unity, and K = span {z, (} C Ci, where z = (1,w,w 2 ,w 3 ) and' = (1,w 2 ,w,w). Determine the nearest point in the closed subspace K to (1, -1,0,1) E Ci. 
6.11. Exercises 427 6.150. Show that an orthonormal basis for the solutions space of the linear equation x - 2y - z = 0 is {tPl, tP2}, where tPl = (2, 1,0) / v'5 and tP2 = (1, -2,5) / y'3O. 6.151. Show that a Hilbert space is finite dimensional iff every complete orthonormal set in X is a basis. 6.152. Let X = ]R2 with the standard inner product and K = {(x, y) EX: 3x + y = OJ. Find K..L and show that every (x, y) E X can be written as the sum of an element from K and an element from K..L. 6.153. Show that there exists an incomplete inner product space X and a proper closed subspace K of X with the following properties: (a) K..L = {OJ, but K is not dense in X (b) K = K # K.L.L ( c) there exists no best approximation in K to any x E X \K. Note: See Examples 6.90, 6.91 and 6.92. 
Chapter 7 Representation of Linear Functionals We know that a linear functional on a Hilbert space X is a linear map from X to 1F (C or IR), which is in fact a special case of linear operators between Banach spaces. At first in Section 7.1 we discuss the structure of linear functionals on Hilbert spaces by proving the Riesz theorem 31 which is widely used in various branches of mathematics. This theorem is ap- preciated as a central property of Hilbert spaces and gives rise to a simple representation for a bounded linear functional on a Hilbert space. This representation is given by taking a suitable inner product (see Theorem 7.3) . 7.1 Riesz Representation Theorem To start with, we consider an inner product space X and let y be a fixed nonzero vector in X. Consider the functional I y : X  1F via (7.1) Iy(x) = (x, y). Clearly, by the linearity of the inner product in the first argument, I y is linear so that every inner product space X gives rise to a collection of linear functionals on X. In view of the CSB inequality I/y(x)1 = l(x,y)1 < IIxlillyll = Cllxll (C = lIyll), we have Illy II < C, and hence I y is bounded. Further, the formula (7.1) for x = y implies that C 2 = IIyll2 = I y(Y) < Illy IIlIyll, Le. II/yll > C. Thus, we have 31 There exists a set of theorems that have the name Riesz attached to them, and several theorems known as the "Riesz representation theorems." 
430 Chapter 7: Representation of Linear Functionals 7.2. Proposition. Every vector y belonging to an inner product space X defines a bounded linear functional fy : X  IF, x I-t (x,y), such that IIfyll = lIyll. In particular, fy E X* for each y E X. The converse of this proposition is not true in general. However, as we shall see next, converse does hold for Hilbert spaces. At this place, it would be appropriate to recall that if p, q > 1 then (IP)* = lq, 1 1 - + - = 1, p q and, in particular, 1 2 is the self-dual. Also, it is possible to obtain a direct geometric proof for any Hilbert space, much quicker than the one passing through 1 2 identification. Indeed, using a very beautiful argument, Riesz obtained that for a given bounded linear functional cP on a Hilbert space X, it is always possible to find a vector y in X such that cP has the form (7.1) . 7.3. Theorem. (Riesz Representation Theorem) Let X be a Hilbert space. Then, for every f E X. = B(X,]F), there exists a unique element p E X such that (7.4) f(x) = (x,p) for all x E X, and IIfll = IIpll ( The element p is called the representation of f). Proof. Let f E X. be arbitrary. To show that it can be represented by the formula (7.4), we consider the kernel K = N/ = {x EX: f(x) = OJ. Note that if K = X, then IIfll = 0, Le. f = 0 with f(x) = 0 for all x EX. In this case, p = 0 satisfies (7.4) for all x EX. Let K :j:. X. Then, there exists a vector x E X such that f(x) :j:. O. Note that K is a proper closed subspace of X. In fact, if {x n } is a sequence in K such that X n  x then If(x)1 = If(x n ) - f(x)1 = If(xn - x)1 < IIfllllx n - xii = Cllx n - xII which tends to 0 as n  00. This observation implies that f(x) = 0, Le. x E K so that K is closed. Also, we observe that if for any p E X for which o = f(x) = (x,p) for all x E K we must have p E K.L. Further, since K is a proper closed subspace of X, by Corollary 6.89, K.L :j:. {OJ. Therefore, we can pick a nonzero vector q E K.L. Then, (k,q) = 0 for all k E K. As q  K, we note that f(q) :j:. 0 
7.1. Riesz Representation Theorem 431 (Otherwise, K.l. = to}, which implies that K = X which is impossible because f(x) :F 0). Further, since X = K f1) K.l., each f(q)x E X can be represented as f(q)x = k + o:q, where k E K, aq E K.l.. We need to find conditions on 0: such that k E K. The linearity of f gives o = f(k) = f(f(q)x - o:q) = f(q)f(x) - af(q) so that the condition on 0: for which k E K is 0: = f(x). Also, since q E K.L, we have o = (k, q) - (f(q)x - o:q, q) - f(q)(x, q) - o:(q, q) - f(q)(x, q) - f(x)lIqI12, since 0: = f(x), which gives f(q) ( f(q) ) f(x) = (x, q) IIqll2 = x, IIqll2 q · Hence, f(q) (7.5) p = IIqll2 q = >.q (say) which is the desired element of X satisfying the required property (7.4). For the uniqueness, we note that if there exist p, p' such that (x,p) = f(x) = (x,p') for all x E X, then (x, p - p') = 0 for all x EX; letting x = p - p' gives that lip - p'11 2 = 0, or p = p'. Thus, p is uniquely determined. The norm equality is trivial when p = O. So we assume that p :F o. Now, by the Cauchy-Schwarz inequality, we have If(x)1 < I(x, p) I < IIxlillpll, and therefore, Ilfll < IIpli. To show the reverse inequality, we consider the unit vector e = pillpil and obtain f(e) = (e,p) = ( II:II 'P ) = IIpll < IIflix. so that we actually have IIfll = Ilpli. . Next we show by an example that Riesz representation theorem does not hold for arbitrary inner product spaces. 
432 Chapter 7: Representation of Linear Functionals 7.6. Example. Consider X = (Coo, 11.112) as a subspace of (1 2 ,11'112). Then (note that X is not a Hilbert space) for z = {Zn}n1 E X, we see that 00 ( 00 ) 1/2 ( 00 ) 1/2  I  I <  I Z nl 2  2 = IIzll2  and therefore, f : X  1F defined by 00 j(z) = I: Zn n n=1 defines a bounded linear functional on X. Suppose on the contrary that the Riesz representation theorem holds for X. Then there exists a unique element p E X such that f(z) = (z,p) for all z EX. But for ej = {6 ij }i1 E X and p = {Pn}n1 EX, we have 00 1 (ej,p) = I: 6 ij Pi = Pj = f(ej) = -:-, i=1 J Le. Pi =  for each j E N, J which gives a contradiction as {1/n}n1  X. . The Riesz representation theorem gives a relationship between a Hilbert space and its dual. In particular, (7.5) immediately yields the following fundamental geometrical fact. 7.7. Corollary. If X is a Hilbert space and f E X., then the complement of N/ is a one-dimensional subspace. In particular X = N/ E9 (N/).1., the direct orthogonal sum of the kernel of f and a one dimensional subspace. Proof. Assume that f :j:. O. Then, by the Riesz representation theorem, there exists a unique element p E X such that f(x) = (x,p) for all x E X. Let M = span {pl. We have that x E M.1. {::=} (x, y) = 0 for all y E M {::=} (x, ap) = 0 for all a E 1F {::=} a (x,p} = 0 for all a E 1F 
7.1. Riesz Representation Theorem 433 which can happen iff x E N/. Thus, Ml. = N f . Since M is a one dimen- sional subspace (and therefore closed), it follows that (N/).L = M.L.L = M. Hence, we have the desired decomposition x = N/ ffi M = N/ ffi Nt = M.L ffi M and therefore, X is the direct orthogonal sum of the kernel of f and a one dimensional subspace. - This corollary also holds when X is a normed space and f : X  1F is a continuous linear functional on X (see Corollary 5.143). 7.8. Example. Suppose that f : 1 2  1F is defined by f ( z) = Z 1 + Z2, for Z = {z 1 , Z2, . . .}. Then If(z)1 = IZl + z21 < IZll + I Z 21 < V2 v' l z lI 2 + I Z 21 2 < V2llzl12 which shows that f is a bounded linear functional on the Hilbert space 1 2 . According to Theorem 7.3, there exists a unique point p E 1 2 such that 00 f(z) = (z,p) = E Zn P n. n=l In fact, for Z = {Zn}n>l we have zk=(z,ek), k=1,2,..., so that Zl + Z2 = (z,el) + (z,e2) = (z,el + e2). Hence, by the Riesz theorem, we can take P = {I, 1,0,0,. . .}. . The following is a consequence of the Riesz representation theorem. 7.9. Theorem. H F(f) is a bounded linear functional on the Hilbert space L2[a, b], then there exists a unique function 9 E L2[a, b] such that F(f) = lab f(t) g(t) dt for all f E L2[a, b], and IIFI12 = IIg112. Conversely, each 9 E L2[a, b] de- termines a bounded linear functional F(f) on L2[a, b] having the above representation. 
434 Chapter 7: Representation of Linear Functionals Let X be an inner product space and p EX. Define fp(x) = (x,p). Clearly, fp E X. and IIfpll = Ilpll. Therefore T : X.  X, fp I-t p, is an isometry. The map X  X., P I-t f p , via (x,p) = fp(x), is called a duality map or a Riesz map. We note that the Riesz map depends on the inner product as well as the space. The identification of dual spaces, in general, can be quite tricky and in Section 5.15 we have discussed the notion of duality on normed spaces. However, as pointed out in Section 5.15, the identification of dual spaces is easy in the case of Hilbert spaces as we see below: If X is a Hilbert space and if we associate with each Y E X a map cP : X  X. defined by cP(y) := cPy, where cPy(x) = (x, y) for all x EX, then Theorem 7.3 asserts that the map y I-t cPy is a bijection from X onto its dual X.. Several properties of this map can be obtained from the definition of cPo Indeed, the ontoness property follows from Theorem 7.3 whereas the one-to-one property is obvious, because for Yl, Y2 EX, cPYl = cPY2 => (x, Yl) = (x, Yl) for all x E X => (x, Yl - Y2) = 0 for all x E X => Yl - Y2 = 0 by the choice x = Yl - Y2 EX. Further, a straightforward verification shows that cPYl +Y2 = cPYl + cPY2 , AcPy = </>x Y , II cPY II = II Y II and therefore, cPY is a conjugate (if X is a complex Hilbert space) linear isometry of X onto X.. Note that if X is a real Hilbert space, then this map is simply a linear isometry of X onto X.. Thus, the Riesz represen- tation theorem establishes the natural isometric one-to-one correspondence between the spaces X and X.. We are not aiming to consider the duals of other Banach spaces such as C[a, b] and LP-spaces, as we leave it as exer- cises (see 7.26) for interested readers. However, we observe that a Banach space, in general, is quite different from its dual whereas Hilbert spaces are special in nature in this respect. Hence, it is natural to raise the following questions: Is the dual of an inner product space again an inner product space? If so, is the dual a complete space? Through the map cPY' we define an inner product on X. by (7.10) (cPY' cPz) = (z, y) for cPY' cPz E X. which makes X. a Hilbert space. The fact that (7.10) defines an inner product on X. follows from the following verification: 
7.1. Riesz Representation Theorem 435 (11) (cPy, cPz) = (z, y) = (y, z) = (t/Jz, t/J.) (12) (AcPy, cPz) = (</Jxy, cPz) = (z, AY ) = A(Z, y) = A(4)y, <Pz) (13) For cPy, cPz, cPw E X*, we have (cPy, cPz + cPw) - (t/JJI' tPz+w) - (z + w, y) - (z,y) + (w,y) (tPlI' tPz) + (tPlI' tPw) (14) (cPy, cPy) = (y, y) = IIyll2 = IIcPyll2 so that IIcPyll > 0, and cPy = 0 iff y = O. Further, since the norm on X* is induced by the inner product defined as (7.10), we observe that X. is a Banach space under the norm IIcPlIl1 = sup IcPy(x) I = sup l(x,y)1 = Ilyli. xex,IIxll=1 xeX,IIxll=1 Hence, X* is a Hilbert space. More precisely, the above discussion gives the following result. 7.11. Corollary. H X is a Hilbert space, then so does the space X*. The map cP : X  X* given by cP(y) := cPy, where cPy(x) = (x, y), is an isometric embedding of X onto X* . Finally, (X*)* = X** is also a Hilbert space with the inner product defined by (FJ,Fg)x.. = (g,/)x. for g,1 E X* where FJ,Fg correspond to I,g E X* respectively, and they are obtained from Theorem 7.3 for the Hilbert space X*. If y, z E X correspond to I, 9 E X* respectively, then (FJ,Fg)x.. = (y,z)x which gives the following theorem. 7.12. Theorem. Each Hilbert space X is isometrically isomorphic to its second dual X**. In particular, every Hilbert space is reflexive, i.e. X** = X. This result does not hold for normed spaces (see Theorem 5.146). 
436 Chapter 7: Representation of Linear Functionals 7.2 Adjoint Operators on Hilbert Spaces Let X and Y be two Hilbert spaces and T E B(X, Y). Note that each fixed Y E Y induces a continuous linear functional I on X via the map x I-t (Tx, y). Indeed, I/(x)1 = I(Tx, y)1 < IITxllllyll < IITllllxllllyll for all x E X so that the functional I is bounded and 11/11 < IITlillyli. Moreover, by the Riesz representation theorem, there exists an element y* E X such that 11/11 = Ily* II and I(x) = (Tx, y) = (x, y*) for all x E X. Here y* is uniquely determined by T and therefore by y itself. This obser- vation shows that we can relate y* with y and write this association with the formula y* = T*y, where T* is an operator with (7.13) (Tx, y)y = (x, T*y) x for all x EX. Since y was any vector in Y, we see that the domain of T* is Y. The operator T* defined in this manner is called the adjoint operator of T. The terms like "Hilbert space adjoint" and "Hermitian conjugate" are also in use in the literature. It can be easily seen that the adjoint operator T* is unique and linear. Indeed, the equality (x, T*Yj) = (Tx, Yj) for j = 1,2, implies that (x,T*(aYl+{3Y2)) - (Tx,aYl+{3Y2) a (Tx, Yl) + {3 (Tx, Y2) - a (x, T*Yl) + {3 (x, T*Y2) - (x, aT*Yl + {3T*Y2) and therefore, T*(aYl + {3Y2) = aT*Yl + {3T*Y2' So, T* is linear. Similarly, for each a E IF, (aT)* = aT*. Note that this definition coincides with that of the definition of adjoint on normed spaces (as given in the previous chapter), provided we use the identification of X, Y with their dual spaces. Moreover, IIT*II = lIy*11 = 11/11 < IITIIIIYII for all y E Y. 
7.2. Adjoint Operators on Hilbert Spaces 437 This inequality shows that T* is bounded and has. a norm at most liT II. Thus, IIT*II < liT II. With the same reasoning, it follows that T** := (T*)* is a bounded linear operator and that IIT**II < IIT*II. Since (T*x,y) = (y,T*x) = (Ty,x) = (x, Ty), the uniqueness of the adjoint T* implies that T** = T. So, IITII = IIT** II < IIT* II . and therefore, it follows that IIT* II = IITII. Suppose that T is an isometry, Le. (Tx, Tx')y = (x, x') x for all x, x' EX. Then, by the definition of adjoint and (7.13), the last equation is equivalent to (x, T*(Tx')) = (x, x'), that is, T*T = Ix, the identity operator on X. Further, if the isometry is onto then T* = T-l. An operator T E B(X, Y) such that T* = T-l is called unitary, and is called self-adjoint, if T* = T. More precisely, "an operator T : X  Y, where X, Yare Hilbert spaces, is called unitary if it is linear, bijective and presenJes inner products: (Tx, Tx') = (x, x') for all x, x'''. Thus, two Hilbert spaces are said to be isomorphic as Hilbert spaces if there exists an unitary operator between them. Further, from the above discussion, it is straightforward to obtain the following result. 7.14. Proposition. A linear surjection T between the Hilbert spaces X and Y is an unitary operator iff it is isometry, i.e. IITxll = Ilxll for all x E X. In addition to the above mentioned results we can easily obtain the following results for adjoint operators. 7.15. Proposition. Let X, Y and Z be three Hilbert spaces. Sup- pose that T, S E B(X, Y), U E B(Y, Z) and a, {3 E IF. Then we have 
438 Chapter 7: Representation of Linear Functionals (i) (aT + (3S)* = aT* + {3S * (ii) (UT)* = T*U* (iii) T*T = 0 iffT = 0 (iv) IIT*TII = IITII2. Proof. The cases (i)-(iii) are easy and therefore, we leave the proof as an exercise. For the proof of (iv), we let IIxll < 1. Then IITxl1 2 - (Tx, Tx) - (x, T*Tx) - (T*Tx, x) < IIT*Txllllxll _ IIT*Tllllx1l 2 so that IITII2 < IIT*TII. Moreover, using the submultiplicative property of the operator norm (see Theorem 5.70(iii)) we obtain IIT*TII < IIT* IIIITIl = IITII2 and (iv) follows if we combine the last two inequalities. . Let us give an example of a bounded linear operator on an inner product space that does not have an adjoint. 7.16. Example. Let X = (Coo, II · 112) and j : X  IF be as in Example 7.6: 00 zn j(z) = L...J -. n n=l Using this bounded linear functional, we have 1 f{e n ) = - for each n E N. n Now, we define T ({zn}) = {f{z), 0, 0,. . .}. Clearly, T is a bounded linear operator and Ten = {l/n, 0, 0, . . .} for each n E N so that 1 (Ten, el) = -. n 
7.2. Adjoint Operators on Hilbert Spaces 439 Suppose that T* E B(X) exists. Then (Tx, y) = (x, T*y) for all x, y E X so that (en,T*el) = (Ten,el) =.! for each n EN. n Thus, if we let T*el = {an}nl then the last equation would yield a n = (en, T*el) = .!, n 1 Le. an = - for each n E N. n But then T*el = {ljn}nl ft X and hence, T* does not exist. . We know that the range space RT of a bounded linear operator T on a Banach space X is not necessarily closed (see Example 5.68). Thus, it is natural to obtain a suitable condition so that RT is closed. 7.17. Proposition. Let X be a Hilbert space and T E B(X) be an isometry. Then RT is closed in X. Proof. We know already that RT is a subspace. To show that RT is closed, we consider a sequence {Yn} in RT which converges in X. Let Yn = TXn, X n E X. Then, as T is an isometry, we see that IITx m - Tx1I = IIxm - xnll = IIYm - Ynll from which it follows that {xn} is a Cauchy sequence in the Hilbert space and therefore, there exists an element x E X such that X n  x as n  00. Continuity of T implies that lim Yn = lim TX n = Tx E RT. n-+oo n-+oo Thus, RT is closed. . We know that the nullspacejkernel NT of an operator T or the ortho- complement S.L of a subs pace S, SeX, is guaranteed to be closed, but the range need not be. The following results describe the relationship among the range space and nullspace of T, T*. 7.18. Theorem. Let X, Y be two Hilbert spaces and T E B(X, Y). Then we have the following statements: (a) R;} = NT. (b) R;}. = NT 
440 Chapter 7: Representation of Linear Functionals .L - (c) NT. = RT .L - (d) NT=RT.. Proof. (a) Let Y E NT.. Then for all x EX, (Tx, y) = (x, T*y) = (x,O) = 0, Le. y E Rf, and therefore, NT. c Rf. Conversely, if y E Rf then the same equation implies that y E NT. and so (a) follows. (b) Since T = T** and RT. = Rf:-, we have R .L.L - R - 1\7.L 1\7.L T. = T. - J.VT.. = J.VT which proves (b). The remaining assertions follow from (a) and (b). . 7.19. Examples of adjoint operators. We list down a set of simple examples of bounded adjoint operators. (i) The zero operator and the identity operator are clearly self-adjoint operators: 0* = 0 and [* = I. In fact, these follow from (Ix, y) = (x, y) = (x,Iy) and (Ox,y) = (O,y) = 0 = (x,Oy), respectively. (ii) Define T : 1 2  , 2 by Tx = {AkXk}kl = (AIXl, A2X2,...) for x = {Xk}kl E 1 2 , where {Ak}k>l is a fixed sequence of scalars from 1 00 . Then, for x E 1 2 and y = {Yk}kl E 1 2 , we have by the inner product on 1 2 (Tx, y) - ({AIXl, A2 X 2,.. .}, {Yl, Y2.. .}) 00 - L AkXk Y k k=l 00 - LXk( A kYk) k=l - (x, T*y) where T*y = { A kYk}kl gives the adjoint for T. In particular, if Ak'S are all real and nonzero for each k > 1, then T is clearly a self-adjoint operator. 
7.2. Adjoint Operators on Hilbert Spaces 441 (iii) Let T : l2  1 2 be defined by Tx = {AkXk}km := (AmXm, Am+lXm+l'...) for x = {Xk}kl, where {Ak}km (fixing m > 1) is a fixed sequence of scalars from loo. Then one can easily find T* using the above procedure. Indeed, for x E l2 and Y = {Yk}kl E 1 2 , we see that (Tx, y) - ({AmXm, Am+lXm+l .. .}, {Yl, Y2,.. .}) 00 - E Ai+m-lXi+m-l Y i i=l 00 - E Xi+m-l ( A i+m-1Yi) i=l - Xl. 0 + X2 .0+ . · . + Xm-l . 0 + X m · ( A mYl) + · . . - (x, T*y) where T*y = {O, 0,...,0, Am Yl, A m +lY2,.. .}. Note that Am Yl on the right hand side occurs at the m-th position in the coordinates of T*y. One can derive several special cases by choosing special values for Ak'S and m. For example, if Ak is a nonzero real number for each k > I (e.g. Ak = Ilk, I/k2,1/2k) and m = 1 then, in this case, T* = T so that T becomes a self-adjoint operator on 1 2 . (iv) Suppose that T, S : 1 2  l2 are defined by Tx = {X3,X4,...} and Sx = {0,0,Xl,X2'. .}, x = {Xk}kl' respectively. Then the adjoints T* and S* are given by T*y = {O, 0, Yl, Y2,...} and S*y = {Y3, Y4,...} for Y = {Yk}kl E 1 2 . ( v ) If T : 1 2  1 2 is the left shift operator defined by Tx = {X2, X3,...} for x = {Xk}kl' then, with the method of the previous items, the adjoint T* of T is clearly given by T*y = {O, Yl, Y2,...} for Y = {YIi}kl' Note that T* is the right shift operator on l2. 
442 Chapter 7: Representation of Linear Functionals ( vi) Similarly, if T : 1 2  l2 is the right shift operator defined by Tx = {O, Xl, X2 ...} for X = {Xk}kl, then 00 (Tx, y) = L Xi Y i+1 = (x, T*y) i=l showing that T* y = {Y2, Y3, . . .}. Thus, T* in this case is the left shift oper ator . (vii) There is an interesting class of bounded linear operators on a Hilbert space given by weighted shifts. Let X be a separable Hilbert space with  = {4JI, 4J2, · · .} as its orthonormal basis. Then each x E X has the form 00 x = L ak4Jk, k=l 00 with L lakl 2 < 00, ak = (X,4Jk)' k=l If T : X  X is the right shift operator with weight given by a bounded sequence {Ak} by 00 Tx = L Ak a k4Jk+l, k=l then it is easy to see that IITII < mF IAkl. Similarly, we can also define the left shift operator with weight given by a bounded sequence {Ak} by 00 Sx = L Ak a k4Jk-1 (4Jo = 0) k=l and find that IISII < mFIAkl. It is easy to see that 4JI ft RT and 4JI ENs. If Ak = 1 for all k, then it follows that ST=I whereas T S is the projection on the orthogonal complement of 4JI and TS :F I. 
7.2. Adjoint Operators on Hilbert Spaces 443 (viii) Let X = en with the standard inner product on en. Then X is a Hilbert space of dimension n. Suppose that T E B(X) is defined by Tz = Az for Z = (Zl, Z2,. . . , zn) E X, where A = (aij) is an n x n matrix with entries from C. Identifying z E cn with the column matrix of order n x 1, we find that n (TZ)i = E aijZj = (ail ai2 . . . ain) j=l Zl Z2 Zn and Tz = ((TZ)I, (TZ)2,..., (Tz)n). Therefore, for all z E X and W = (WI, W2, . . . , W n ) EX, we have n (Tz, w) = E[(Tz)i] Wi i=l - t ( t aijZj ) Wi i=l j = 1 n n - EEaijZj W i i=l j=1 - t Zj ( t a ijWi ) j=l i=1 - (Z, T*w). Thus, for all w = (W1, W2,..., w n ) E X, n (T*W)i = E aj iWj j=1 which shows that T* is given by T*w=A*w for all W=(W1,W2,...,W n )EX, where A* = ( aj i)' This implies that the matrix that represents the adjoint T* of T is simply the conjugate transpose of the matrix that represents T. In particular, one has the following: (a) If A is a Hermitian matrix (Le. aij = aj i for all i, j = 1,2, . . . n), then T*w = A*w = Aw which shows that T is a self-adjoint oper ator . 
444 Chapter 7: Representation of Linear Functionals (b) If X = IRn, the Euclidean space, and if T E B(IRn) is defined by Tx = Ax where A = (aij) is some n x n real matrix, then the. adjoint operator T* of T is represented by the transpose matrix AT = (aji) so that T*w = AT w. In particular, if A is given by the symmetric matrix, then A = AT and therefore, T* = T; Le. T is a self-adjoint operator. (ix) Let T : 1 2  l2 be defined by Tz = Az, or equivalently, 00 (TZ)i = L aijZj = (AZ)i, j=1 where Z = {Z1, Z2, Z3,...} E 1 2 , Tz = {(TZ)1, (TZ)2, (TZ)3,"'} E 1 2 , and A is an infinite matrix with the scalars aij as (i,j)-th entries of the matrix. Here (Az) i denotes the (i, 1)- th entry in the infinite dimensional column matrix of Az. If E  1 E j 11 a ijl2 < 00, then T is a bounded operator. For Z = {Zi}i>1 and W = {Wi}i>1, we have - - 00 (Tz,w) = L{(Tz)i} Wi i=1 00 00 - L L aijZj W i i= 1 j= 1 00 00 - LZjL ai jWi j =1 i=1 - (z, T*w) where 00 (T*w)j = L ai jWi = (A*w)j. i=l Here (A*w)j denotes the (j,l)-th entry of the column matrix A*w which shows that T*w = A*w where A* is the conjugate transpose of the infinite matrix A. In particular, if A is an infinite Hermitian matrix (Le. aij = aj i for all 1 < i, j < 00), then T becomes a self-adjoint operator. (x) Let X = L2[a, b] and Y = L2[c, d]. Suppose that T E B(X, Y) is defined by (T f)(t) = lab k(t, 8)f(8) ds, c < t < d, 
7.2. Adjoint Operators on Hilbert Spaces 445 where the kernel k(t, s) is a continuous function from [c, d] x [a, b] into C. It is easy to derive the boundedness of this operator and so we let it as an exercise. Then, for f, 9 E B(X) (by Fubini's theorem the order of integration below is justified), (T I, g) = l d (l b k(t, s)/(s) dS) g(t) dt - l b I(s) ds (l d k(t, s)9(i) d t) - l b I(s) (l d k(t, s) g(t) dt) ds - l b 1( )(T.g(s)) ds - (f, T*g) where (T*g)(s) = l d k(t, s) g(t) dt for each 9 E L2[a, b]; or interchanging the role of t and s, (T*g)(t) = l d k(s, t) g(s) ds. In particular, if k{t, s) = k(s, t) for all s, t and a = c, b = d then T* = T so that T becomes a self-adjoint operator. For example, if k(t, s) = (t - S)2 then T* = T holds and hence, in this case, T is a self-adjoint operator. (xi) Let cP(t) be a fixed complex valued continuous function on [a, b] and T E B(L2[a, b]), a multiplication operator defined by (Tf)(t) = cP(t)f(t). Then, for all f, 9 E L2[a, b], it follows that (T I, g) = l b q,(t)/(t) (g(t)) dt = (f, T. g) where (T*g)(t) = cP(t)g(t). Note that T* is also a multiplication operator and T* is obtained from the operation of multiplication by the complex conjugate of <p. In particular, if <p(t) (eg. t or t 2 ) is a real valued continuous function on [a, b] then T becomes a self-adjoint operator. 
446 Chapter 7: Representation of Linear Functionals We have already shown that every projection operator is idempotent. Now we prove the following simple result. 7.20. Theorem. Every projection operator on a Hilbert space X is self-adjoint. Conversely, every self-adjoint operator that is idempotent on X is a projection. Proof. Let PK : X  K be a projection, where K is a closed subspace of the Hilbert space X. Then for every Xl, X2 EX, we have the unique representation Xi = Yi +Zi, Yi E K and Zi E KJ.. (i = 1,2). Now, (PKXl, X2) - (Yl, Y2 + Z2) - (Yl, Y2) + (YI, Z2) - (Yl, Y2), since (Yl, Z2) = 0, - (YI + Zl, Y2), since (Zl, Y2) = 0, - (Xl, PK X 2), and therefore, PK is self-adjoint. For the converse part, let P be an operator on the Hilbert space X such that p2 = P and P = P*. Define K = P(X). Clearly, K is a subspace of X and is also closed. Indeed, if Yn = PXn  Z then pYn = p2xn = PXn = Yn. Continuity of P shows that Z = lim Yn = lim pYn = P z, Le. z E K, n-+oo n-+oo and therefore, K is a closed subspace of X. Since P is self-adjoint and idempotent, we have (x - Px,py) = (P*(x - Px),y) = (Px - P 2 x,y) = 0 for all Y EX so that X - Px E K J... Therefore, each X E X has the unique decomposition x=Px+(x-Px), PXEKandx-PxEKJ... Hence, P is a projection on K. . We now prove a result that illustrates a method of finding the norm of a self-adjoint operator. 
7.2. Adjoint Operators on Hilbert Spaces 447 7.21. Theorem. (Rayleigh) If T is a self-adjoint operator on a Hilbert space X, then IITII = sup I(Tx,x)l. IIxll=1 Proof. Let 0: = sUPllxII=11(Tx,x)l. H Ilxll = 1, then I (Tx, x) I < IITxllllxll < IITllllxl1 2 = IITII so that 0: < IITII. Let us prove the reverse inequality. Obviously, it is enough to prove that IITzll < 0: Ilzll for all z for which Tz :j:. O. For all x, y in X, we find that (T(x + y), x + y) - (T(x - y), x - y) - 2(Tx, y) + 2(Ty, x) - 2(Tx, y) + 2(y, T*x) - 4Re(Tx,y), sinceT=T*. We may multiply y by a suitable complex number of modulus one, so we can assume that (Tx, y) > O. This observation shows that I(Tx,y)1 1 - 4 [(T(x+ y ),x+ y )-(T(x-y),x- y )] 0: < 4 (lIx + yll2 + IIx - y1l2) _  (lIxll 2 + lIyIl2). If we choose Ilxll = 1 and y = Tx/IITxll, the last inequality gives IITxll < 0:. It follows that IITII < 0: and the conclusion follows. . Finally, we provide a sufficient condition for an operator to be a zero oper ator . 7.22. Proposition. H X is a complex Hilbert space and T E B(X), then T = 0 iff (Tx,x) = 0 for all x E X. Proof. Suppose that (Tx,x) = 0 for all x E X. Then for each x,y E X, we have (7.23) 0 = (T(x + y),x + y) = (Tx,y) + (Ty,x). Similarly, o = (T(ix + y),ix + y) = i(Tx,y) - i(Ty,x) 
448 Chapter 7: Representation of Linear Functionals so that (7.24) 0 = (Tx, y) - (Ty, x). Adding (7.23) and (7.24) we find that (Tx,y) = 0 for all x,y E X. Substitution of y = Tx in the last equation yields IITxll 2 = 0 for all x E X and therefore T = O. The converse part is trivial. . 7.3 Exercises 7.25. Suppose that j : 1 2  1F is defined by j(Z) = Z3, Z = {Zn}n1' Show that j is a bounded linear functional on 1 2 . Find the unique vector p E 1 2 such that j(z) = (z,p) for all z E 1 2 . 7.26. Prove the following statements: (i) For 1 < p, q < 00 and ! + ! = 1, (£P[a, b])* = Lq[a, b] p q (ii) For 1 < p < 00, LP[a, b]-spaces are reflexive and separable (iii) ( L 00 [a, b]) * :F L 1 [a, b]. (iv) (L 1 [a, b])* = Loo[a, b]. 7.27. If X is a Hilbert space and T E B(X), then show that T is self-adjoint iff (Tx, x) real for all x EX. 7.28. If X = JF'1 with the standard inner product and P : X  X is defined by P(Z1,Z2,... ,zn) = (O,Z2,". ,zn), then show that P is a self-adjoint operator. 7.29. If X is a Hilbert space, T E B(X) is non-zero and self-adjoint, then show that Tn is also non-zero and self-adjoint. 
Bibliography [Ah] L.A. AHLFORS, Complex analysis, 3d Ed., McGraw-Hill Book Com- pany, Inc., 1979. [AVV] G.D. ANDERSON, M.K. VAMANAMURTHY, AND M. VUORINEN, Conformal Invariants, Inequalities, and Quasiconformal Maps, John Wiley & Sons, 1997. [Ap] T.M. ApOSTOL, Mathematical Analysis, Addison-Wesley, 1957 [Du] P .L. DUREN, Univalent Functions (Die Grundlehren der mathema- tischen Wissenschaften 259), Springer- 'lerlag, 1983. [HLP] G.H. HARDY, J.E. LITTLEWOOD, AND G. POLYA, Inequalities, 2nd Ed., Cambridge University Press, 1952. [Mi] D.S. MITRINOVIC, Analytic Inequalities (Die Grundlehren der math- ematischen Wissenschaften 165), 1970. [Pom] CH. POMMERENKE, Univalent Functions, Vandenhoeck and Ruprecht: Gottingen, 1975. [Po] S. PONNUSAMY, Foundations of Complex Analysis, Narosa Publish- ing House, 1995. [Roc] R. T. ROCKAFELLAR, Convex Analysis, Princeton University Press, 1970. [Za] A.C. ZAANEN, Continuity, Integration and Fourier Theory, Springer- Verlag, 1989. BOOKS FOR FURTHER READING: [Au] J.P. AUBIN, Applied Functional Analysis, John Wiley & Sons, 1979. [BaN] G. BACHMAN, AND L. NARICI, Functional Analysis, Academic Press, 1966. [Ba] B. BOLLOBAS, Linear Analysis, Cambridge University Press, 1990. 449 
450 BIBLIOGRAPHY [Da] M.M. DAY, Normed Linear Spaces, Springer-Verlag, 1973. [GoP] C. GOFFMAN, AND G. PEDRICK, First Course in Functional Anal- ysis, Prentice Hall of India Pvt. Ltd., 1996. [Hal] P.R. HALMOS, Introduction to Hilbert Space, Chelsea Publishing Company, 1951. [Ha2] P.R. HALMOS, Finite Dimensional Vector Spaces, Springer-Verlag, 1958. [HeS] E. HEWITT, AND K. STROMBERG, Real and Abstract Analysis, Springer-Verlag, 1969. [LuS] L.A. LUSTERNIK, AND V.J. SOBOLEV, Elements of Functional Anal- ysis, Hindustan Publishing Corporation, New Delhi, 1971. [Pe] GERT K. PEDERSEN, Analysis Now, Springer-Verlag, 1989. [Ru] W. RUDIN, Functional Analysis, McGraw-Hill Book Company Inc., 1973. [Si] G.F. SIMMONS, Introduction to Topology and Modern Analysis, McGraw-Hill Publishing Company Inc., 1963. [TaL] A.E. TAYLOR, AND D.C. LAY, Introduction to Functional Analysis, John Wiley & Sons, 1980. [Yo] K. YOSIDA, Functional Analysis, Springer-Verlag, 1980. [Zi] R.J. ZIMMER, Essential Results of Functional Analysis, University of Chicago Press, 1990. 
Index B(X), 78 BV[a, b], 1 C(X, Y), 80 C[a, b], 79, 80 C k ( I), 18 C k [a, b], 334 C 1 (I), 18 Coo(I), 18 Cc[a, b], 79, 80 C[a, b], 21 CF(X), 163 CF[a, b], 1 C[a, b], 21 Cn\[a, b] := C 1 [a, b], 21 L(V), 37 L(V, W), 37 L(X), 255 L(X, Y), 255 L2[a, b], 180 X., 142 C U {oo }, 65 en, 29 F, 1 JF'1, 29 N, 13 IR, 1 IR+ 13 , IRn 29 , IRt , 13 Z,12 A 2[a, b], 179 C, 167 Coo, 157, 167 Co, 167 451 IP 76 , IP(n), 76 1 00 76 , loo(n), 76 B(X), 142, 255 B(X, Y), 142, 255 B(X,]F), 142 P(IR), 31 1m T, 38 KerT, 38 co(S), 44 null (T), 38 rank (T), 38 Pn(]F), 41 absolutely convergent, 66, 158 absolutely summable, 77 additive identity, 9 addi ti ve inverse, 9 adjoint operator, 329, 436 algebra, 264 Banach, 265 commutative, 265 commutative Banach, 265 noncommutative, 265 normed, 265 algebraic dual, 321 AM-GM inequality, 52 arcwise connected, 109 arithmetic mean, 359 attractive fixed point, 208 Baire Categories, 290 Baire Category Theorem, 291 
452 Banach algebra, 265 commutative, 265 Banach space, 156 Banach Theorem, 297 basis, 32 complete orthonormal, 399, 400 Hamel, 399 Schauder, 253 Bernstein polynomial, 250 Bessel's inequality, 402, 404 bijective, 11 bilinear, 345 bounded essentially, 181 bounded linear functionals, 256 bounded linear operator, 255, 259 canonical map, 287 Cantor set, 92 Cantor's Theorem, 291 cardinal number, 14 cardinality, 14 Cauchy sequence, 120, 156 Cauchy-Schwarz inequality, 68 chordal metric, 65 closed, 90 closed ball, 82 closed convex hull, 44 Closed Graph Theorem, 301 closed mapping, 111 closed operator, 266, 299 closed subspace, 354 codomain, 10 commutative algebra, 265 commutative Banach algebra, 265 compact, 104, 105 relatively, 105 sequential, 105 complete, 124 complete normed space, 156 complete orthonormal, 399 complete orthonormal basis, 399, 400 completion, 127, 279 metric space, 127 INDEX normed space, 279 complex Fourier series, 416 complex number, 5 conjugate, 6 inverse, 10 modulus, 7 complex plane, 7 concave function, 43 conjugate, 6 conjugate bilinear, 345 conjugate index, 67 conjugate operator, 329 conjugate space, 264 connected, 108 arcwise, 109 continuous, 94 uniformly, 105 continuous extension, 307 continuous operator, 259 continuously differentiable, 18 contraction, 200 contraction mapping, 200 contractive map, 210 convergence pointwise, 172 uniform, 172 convergent, 65 absolutely, 66, 158 strongly, 262 uniformly, 261 convergent series, 66 convex function, 43 convex functional, 256 convex hull, 44 convex set, 43 coordinate spaces, 143 coordinate vector, 33 countable, 14 covering, 103 dense subset, 101 derivative, 20 diameter, 61 differentiable, 15 differential operator, 268 
INDEX 453 dimension, 32 Dirichlet, 27, 97 Dirichlet kernel, 417 disconnected, 108 discrete metric, 62 distance, 145 Chordal, 65 mean-square, 353 divergent, 66 divergent series, 66 domain, 10 dual space, 142, 264 duality map, 434 differentiable, 20, 21 functional, 221 convex, 256 fundamental sequence, 120 geometric mean, 359 graph, 299 equivalence relation, 4 equivalent metrics, 87, 99, 100 essential supremum, 181 essentially bounded, 181 Euclidean, 62 Euclidean metric, 75 \ extension, 307 Holder's inequality, 68 Hahn-Banach Theorem, 310 Hamel basis, 399 harmonic mean, 359 Hausdorff space, 93 Hilbert space, 353 separable, 401 Hilbert space(s) equivalence, 414 homeomorphism, 111 uniform, 111 hyperbolic metric, 80 factor space, 286 field, 9 finite covering, 103 finite dimension, 32 . first category, 290 fixed point, 197 attractive, 208 repulsive, 208 Fourier coefficients, 411, 416 Fourier expansion, 411 Fourier series, 411, 416 complex, 416 Frechet's metric, 67 Fredholm integral equation, 216 function, 10 characteristic, 97 concave, 43 convex, 43 differentiable, 15 function spaces, 143 function (s ) composite, 11 continuous, 20, 94 idempotent operator, 397 induced norm, 347 Inequality Minkowski's, 70 AM-GM, 52, 359 Bessel's, 404 Cauchy-Schwarz, 68, 360 Cauchy-Schwarz- Buniakowski, 348 Holder's, 68 Jensen's, 46 triangle, 48, 60, 144, 348 Young's, 56, 69 infinite dimension, 32 injective, 11 inner product, 343 semi, 344 weighted, 426 inner product space, 344 closed subspace, 354 complete, 353 complex, 344 real, 344 standard, 346, 357 
454 INDEX subspace, 354 interior, 82 interior point, 82 invertible, 275 isometric, 118, 364 isometric isomorphism, 364 isometrically isomorphic, 364 isometry, 108, 118, 201, 363 isomorphism, 364 isomorphic, 36, 364 isometrically, 364 isomorphism, 36, 364 iteration method, 199 nonexpansive, 108, 210 one-to-one, 11 onto, 11 open, 111 surjective, 11 maximal orthonormal, 399 maximum metric, 75 maximum norm, 162, 356 meager, 290 mean arithmetic, 359 geometric, 359 harmonic, 359 metric, 60 chordal, 65 discrete, 62 equivalent, 99 Euclidean, 75 Frechet's, 67 homogeneous, 146 invariant, 146 maximum, 75 mean-square, 353 natural, 75 trivial, 62 metric space, 60 bounded, 60, 61, 67 compact, 104, 105 complete, 124 completion, 127 discrete, 146 unbounded, 60, 77 metric subspace, 61 Minimum principle, 382 Minkowski's inequality, 70 modulo, 286 modulus, 7 multiplicative identity, 9 multiplicative inverse, 9 Jensen's inequality, 46 Kronecker symbol, 238, 370 Lebesgue, 183 Lebesgue integral, 22 limit, 19 limit point, 65, 90 linear functional extension of, 307 linear functionals, 42 linear operator, 36 linearly dependent, 32 linearly independent, 32 linearly ordered, 4 Lipschitz condition, 108 Lipschitz constant, 108, 200 Lipschitzian, 108 lower integral, 24 map duality, 434 Riesz, 434 mapping, 10 bijective, 11 closed, 111 contractive, 210 homeomorphic, 111 injective, 11 isometry, 118, 364 Lipschitz, 108 natural embedding, 323 natural metric, 75, 145, 353 neighbourhood, 82 nonexpansive map, 210 nonexpansive mapping, 108 
INDEX 455 nonreflexive, 327 norm, 144, 310, 347 convergent, 147 essential supremum, 181 induced, 347 maximum, 162 operator, 255 semi, 145 strict, 358 supremum, 173 uniform, 162, 255 norm dual, 264, 321 normed algebra, 265 normed space, 144, 310 complete, 156 completion, 279 nowhere dense, 290 null space, 38 nullity, 38 shift, 268 unbounded, 255 unitary, 437 Volterra integral, 206 operator norm, 255 orthogonal, 367, 370, 372 complement, 372 orthogonal projection, 390, 396 orthonormal, 370 complete, 399 maximal, 399 orthonormal basis, 400 complete, 400 orthonormalization, 377 one-to-one, 11 onto, 11 open ball, 82 open base, 84 open covering, 103 open mapping, 111 Open Mapping Theorem, 297 open set, 82 operator, 35, 221 additive, 35 adjoint, 329, 436 backward, 268 bounded, 255 bounded linear, 255, 259 closed, 266, 299 conjugate, 329 continuous, 259 differential, 268 homogeneous, 35 idempotent, 397 invertible, 275 left shift, 441, 442 linear, 36 right shift, 442 self-adjoint, 437 parallelogram rule, 243 Parseval identity, 407 Parseval Relation, 407 Parseval theorem, 407 partial ordering, 4 partially ordered set, 4 partition, 22 norm or mesh, 22 path connected, 109 Plancherel relation, 409 pointwise, 172 pointwise convergence, 172 power set, 14 pre-Hilbert space, 344 principal value, 8 projection, 233, 396 complementary, 397 orthogonal, 390, 396 Stereographic, 116 Projection theorem, 383, 390 proper subset, 3 pseudo-metric, 60, 133 punctured disc, 20 Pythagorean, 368 quotient map, 287 quotient space, 286 range, 11 range space, 38 
456 INDEX Schauder basis, 253 second category, 290 self-adjoint operator, 437 semi-inner product, 344 semi-inner product space, 344 seminorm, 145, 309 seminormed space, 309 separable, 101 sequence, 13 Cauchy, 120, 156 convergent, 65 fundamental, 120 sequence spaces, 143 sequential compactness, 105 series convergent, 66 divergent, 6, Fourier, 411 shift operator, 268 space algebraic dual, 321 Banach, 156 complete, 156 conjugate, 264 dual, 142, 264 factor, 286 Hausdorff, 93 Hilbert, 353 inner product, 344 metric, 60 nonreflexive, 327 pre-Hilbert, 344 quotient, 286 reflexive, 323 topological, 93 unitary, 345 spanning set, 32 square integrable, 179 square summable, 77 standard basis, 33 stereographic, 116 Stereographic projection, 116 strict norm, 358 strictly contractive, 210 strictly convex, 242 strong convergence, 262 sub covering, 103 submultiplicative, 264 subspace, 31 metric, 61 summable, 161 absolutely, 161 supnorm, 173 supremum norm, 173 surjective, 11 rank, 38 Rayleigh Theorem, 447 reflexive, 323 relative metric, 61 relatively compact, 105 repulsive fixed point, 208 residual set, 290 Etiemann-Lebesgue, 408 Riemann-Stieltjes, 22 Riemann-Stieltjes integral, 25 Etiesz map, 434 Etiesz Theorem, 430 rotund, 242 Theorem Baire Category, 291 Banach, 297 Bohman-Korovkin, 253 Bolzono- Weierstrass, 104 Cantor, 291 Cantor-Bernstein, 15 Closed Graph, 301 Hahn-Banach, 310 Heine-Borel, 104 Open Mapping, 297 Parseval, 407 Peano-Picard, 214 Projection, 383, 3O Pythagorean, 368, 372 Rank-Nullity, 39 Rayleigh, 447 Riemann-Lebesgue, 408 Etiesz, 430 Riesz Representation, 430 
INDEX 457 Riesz-Fisher, 403 Uniform Boundedness, 305 Uniform Convergence, 173 Weierstrass, 247 topological space, 93 topology, 93 Zariski, 93 totally ordered, 4 triangle inequality, 48, 60 trivial metric, 62 uniform convergence, 172, 262 uniform norm, 162, 255, 356 uniformly continuous, 105 unit vector, 144 unitary, 437 unitary space, 162, 345 upper bound, 5 upper integral, 24 Young's inequality, 54, 56, 69 unbounded operator, 255 uncountable, 14 Uniform Boundedness Principle, 305 Zariski topology, 93
un un n . Ions 0 ti n I I is s. ponnusamy Associate Professor Department of Mathematics Indian Institute of Technology, Madras Chennai-600 036, India Foundations of Functional Analysis provides fundamental concepts about the theory, application and various methods involving functional analysis for students, teachers, scientists and engineers. Divided into three parts it covers: · Basic facts of linear algebra and real analysis · Normed spaces, contraction mappings, linear operators between normed spaces and fundamental results on these topics · Hilbert spaces and the representation of continuous linear functional with applications In this self-contained book, all the concepts, results and their consequences are motivated and illustrated by numerous examples in each chapter with carefully chosen exercises. a - rnation I ISBN 1-84265-079-3 9 791842 650799