Front cover
Title page
Date-line
Dedication
Contents
Preface
Chapter 1. Elementary Spectral Theory
1.2. The Spectrum and the Spectral Radius
1.3. The Gelfand Representation
1.4. Compact and Fredholm Operators
Exercises
Addenda
Chapter 2. C$^\ast$-Algebras and Hilbert Space Operators
2.2. Positive Elements of C$^\ast$-Algebras
2.3. Operators and Sesquilinear Forms
2.4. Compact Hilbert Space Operators
2.5. The Spectral Theorem
Exercises
Addenda
Chapter 3. Ideals and Positive Functionals
3.2. Hereditary C$^\ast$-Subalgebras
3.3. Positive Linear Functionals
3.4. The Gelfand-Naimark Representation
3.5. Toeplitz Operators
Exercises
Addenda
Chapter 4. Von Neumann Algebras
4.2. The Weak and Ultraweak Topologies
4.3. The Kaplansky Density Theorem
4.4. Abelian Von Neumann Algebras
Exercises
Addenda
Chapter 5. Representations of C$^\ast$-Algebras
5.2. The Transitivity Theorem
5.3. Left Ideals of C$^\ast$-Algebras
5.4. Primitive Ideals
5.5. Extensions and Restrictions of Representations
5.6. Liminal and Postliminal C$^\ast$-Algebras
Exercises
Addenda
Chapter 6. Direct Limits and Tensor Products
6.2. Uniformly Hyperfinite Algebras
6.3. Tensor Products of C$^\ast$-Algebras
6.4. Minimality of the Spatial C$^\ast$-Norm
6.5. Nuclear C$^\ast$-Algebras and Short Exact Sequences
Exercises
Addenda
Chapter 7. K-Theory of C$^\ast$-Algebras
7.2. The K-Theory of AF-Algebras
7.3. Three Fundamental Results in K-Theory
7.4. Stability
7.5. Bott Periodicity
Exercises
Addenda
Appendix
Notes
References
Notation Index
Subject Index
Back cover

Автор: Murphy G.J.  

Теги: mathematics  

ISBN: 0-12-511360-9

Год: 1990

Текст
                    C* -ALGEBRAS
AND
OPERATOR
THEORY
Gerard J. Murphy


C*-ALGEBRAS AND OPERATOR 1HEORY 
@ C*-ALGEBRAS AND OPERATOR THEORY Gerard J. Murphy Mathematics Deportment University College Cork, Ireland ACADEMIC PRESS, INC. Harcourt Brace Jovanovich, Publishers Boston San Diego New York London Sydney Tokyo Toronto 
This book is printed on acid-free paper. @ Copyright @ 1990 by Academic Press, Inc. All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopy, recording, or any information storage and retrieval system, without permission in writing from the publisher. ACADEMIC PRESS, INC. 1250 Sixth Avenue, San Diego, CA 92101 United Kingdom Edition published by ACADEMIC PRESS LIMITED 24-28 Oval Road, London NW1 7DX Library of Congress Cataloging-in-Publication Data Murphy, Gerard J. C. -algebras and operator theory / Gerard J. Murphy. p. cm. Includes bibliographical references. ISBN 0-12-511360-9 (alk. paper) 1. C. -algebras. 2. Operator theory. I. Title. QA326.M87 1990 512' .55-dc20 90-524 CIP Printed in the United States of America 90 91 92 93 9 8 7 6 5 4 3 2 1 
For my family Mary, Alison, Adele, Neil
Contents Preface 9 Chapter 1. Elementary Spectral Theory 1.1. Banach Algebras 1 1.2. The Spectrum and the Spectral Radius 5 1.3. The Gelfand Representation 13 1.4. Compact and Fredholm Operators 18 Exercises 30 Addenda 34 Chapter 2. C*-Algebras and Hilbert Space Operators 2.1. C*-Algebras 35 2.2. Positive Elements of C*-Algebras 44 2.3. Operators and Sesquilinear Forms 48 2.4. Compact Hilbert Space Operators 53 2.5. The Spectral Theorem 66 Exercises 73 Addenda 75 Chapter 3. Ideals and Positive Functionals 3.1. Ideals in C*-Algebras 77 3.2. Hereditary C*-Subalgebras 83 3.3. Positive Linear Functionals 87 3.4. The Gelfand-Naimark Representation 93 3.5. Toeplitz Operators 96 Exercises 107 Addenda 110 vn
Vlll Contents Chapter 4. Von Neumann Algebras 4.1. The Double Commutant Theorem 112 4.2. The Weak and Ultraweak Topologies 124 4.3. The Kaplansky Density Theorem 129 4.4. Abelian Von Neumann Algebras 133 Exercises 136 Addenda 138 Chapter 5. Representations of C*-Algebras 5.1. Irreducible Representations and Pure States 140 5.2. The Transitivity Theorem 149 5.3. Left Ideals of C*-Algebras 153 5.4. Primitive Ideals 156 5.5. Extensions and Restrictions of Representations 162 5.6. Liminal and Postliminal C*-Algebras 167 Exercises 171 Addenda 172 Chapter 6. Direct Limits and Tensor Products 6.1. Direct Limits of C*-Algebras 173 6.2. Uniformly Hyperfinite Algebras 178 6.3. Tensor Products of C*-Algebras 184 6.4. Minimality of the Spatial C*-Norm 196 6.5. Nuclear C*-Algebras and Short Exact Sequences 210 Exercises 213 Addenda 216 Chapter 7. K-Theory of C*-Algebras 7.1. Elements of K-Theory 217 7.2. The K-Theory of AF-Algebras 221 7.3. Three Fundamental Results in K-Theory 229 7.4. Stability 241 7.5. Bott Periodicity 245 Exercises 262 Addenda 264 Appendix 267 Notes 277 References 279 Notation Index 281 Subject Index 283
Preface This is an introductory textbook to a vast subject, which although more than fifty years old is still extremely active and rapidly expanding, and coming to have an increasingly greater impact on other areas of math- ematics, as well as having applications to theoretical physics. I have at- tempted to give a leisurely and accessible exposition of the core material of the subject, and to cover a number of topics (the theory of C*-tensor products and K-theory) having a high contemporary profile. There was no intention to be encyclopedic, and many important topics had to be omitted in order to keep to a moderate size. This book is aimed at the beginning graduate student and the special- ist in another area who wishes to know the basics of this subject. The reader is assumed to have a good background in real and complex analysis, point set topology, measure theory, and elementary general functional anal- ysis. Thus, such results as the Hahn-Banach extension theorem, the uni- form boundedness principle, the Stone-Weierstrass theorem, and the Riesz- Kakutani theorem are assumed known. However, the theory of locally convex spaces is not presupposed, and the relevant material including the Krein-Milman theorem and the separation theorem are developed in a brief appendix. The book is arranged so that the appendix is not used until Chapter 4, and the first three chapters can, if desired, form the basis of a short course. The background material for the book is covered by the following textbooks: [Coh], [Kel], [Rud 1], and [Rud 2]. Each chapter concludes with a list of exercises arranged roughly accord- ing to the order in which the relevant item appeared in the chapter, and statements of additional results related to, and extending, the material in the text. The symbols N, Z, R, R+, and C refer, respectively, to the sets of non- negative integers, integers, real numbers, non-negative real numbers, and complex ntUl1bers. Other notation is explained as needed. . IX 
x Preface The reader who has finished this book and wants direction for further study may refer to the Notes section where some books are recommended. I am indebted to many authors of books on operator theory and oper- ator algebras. Section 7.5 of this book is based on the approach of J. Cuntz to K-theory. I should like to thank my colleagues Trevor West and Martin Mathieu for reading preliminary drafts of some of the earlier chapters. Gerard J. Murphy 
CHAPTER 1 Elementary Spectral Theory In this chapter we cover the basic results of spectral theory. The most important of these are the non-emptiness of the spectrum, Beurling's spec- tral radius formula, and the Gelfand representation theory for commutative Banach algebras. We also introduce compact and Fredholm operators and analyse their elementary theory. Important concepts here are the essential spectrum and the Fredholm index. Throughout this book the ground field for all vector spaces and alge- bras is the complex field C, unless the contrary is explicitly indicated in a particular context. 1.1. Banach Algebras We begin by setting up the basic vocabulary needed to discuss Banach algebras and by giving some examples. An algebra is a vector space A together with a bilinear map A 2 -+ A, (a, b)  ab, such that a( be) = (ab)e ( a, b, e E A). A subalgebra of A is a vector subspace B such that b, b' E B => bb' E B. Endowed with the multiplication got by restriction, B is itself an algebra. A norm 11.11 on A is said to be submultiplicative if lIabll < lIalillbll (a, b E A). In this case the pair (A, 11.11) is called a normed algebra. If A admits a unit 1 (a1 = 1a = a, for all a E A) and 11111 = 1, we say that A is a unital normed algebra. 1 
2 1. Elementary Spectral Theory If A is a normed algebra, then it is evident from the inequality lIab - a'b'lI < lIalillb - b'lI + lIa - a'lllIb'lI that the multiplication operation ( a, b)  ab is jointly continuous. A complete normed algebra is called a Banach algebra. A complete unital normed algebra is called a unital Banach algebra. A subalgebra of a normed algebra is obviously itself a normed alge- bra with the norm got by restriction. The closure of a subalgebra is a subalgebra. A closed subalgebra of a Banach algebra is a Banach algebra. 1.1.1. Ezample. If 5 is a set, /,00(5), the set of all bounded complex- valued functions on 5, is a unital Banach algebra where the operations are defined pointwise: (f + g) (x) = f(x) + g(x) (fg)(x) = f(x)g(x) (Af)(x) = Af(x), and the norm is the sup-norm Ilfll oo = sup If(x)l. xES 1.1.2. Ezample. If f! is a topological space, the set Cb(f!) of all bounded continuous complex-valued functions on f! is a closed subalgebra of /,oo(f!). Thus, Cb(f!) is a unital Banach algebra. If f! is compact, C(f!), the set of continuous functions from f! to C, is of course equal to Cb(f!). 1.1.3. Ezample. If f! is a locally compact Hausdorff space, we say that a continuous functionf from f! to C vanishes at infinity, if for each positive number £ the set {w E f! Ilf(w)1 > £} is compact. We denote the set of such functions by C o ( f!). It is a closed subalgebra of Cb(f!), and therefore, a Banach algebra. It is unital if and only if f! is compact, and in this case Co(f!) = C(f!). The algebra Co(f!) is one of the most important examples of a Banach algebra, and we shall see it used constantly in C*-algebra theory (the functional calculus). 1.1.4. Ezample. If (f!, Jl) is a measure space, the set L 00 (f2, Jl) of ( classes of) essentially bounded complex-valued measurable functions on f! is a unital Banach algebra with the usual (pointwise-defined) operations and the essential supremum norm f  Ilflloo. 
1.1. Banach Algebras 3 1.1.5. Eample. If f! is a measurable space, let Boc>(f!) denote the set of all bounded complex-valued measurable functions on f!. Then Boc>(f!) is a closed subalgebra of fOC>(f!), so it is a unital Banach algebra. This example will be used in connection with the spectral theorem in Chapter 2. 1.1.6. Eample. The set A of all continuous functions on the closed unit disc D in the plane which are analytic on the interior of D is a closed subalgebra of C(D), so A is a unital Banach algebra, called the di3c algebra. This is the motivating example in the theory of function algebras, where many aspects of the theory of analytic functions are extended to a Banach algebraic setting. All of the above examples are of course abelian-that is, ab = ba for all elements a and b-but the following examples are not, in general. 1.1.7. Eample. If X is a normed vector space, denote by B(X) the set of all bounded linear maps from X to itself (the operator3 on X). It is routine to show that B(X) is a normed algebra with the pointwise-defined operations for addition and scalar multiplication, multiplication given by ( u, v)  u 0 v, and norm the operator norm: lIu II = sup lIu( x )11 = sup lIu( x )11. x#O IIxll IIxlll If X is a Banach space, B(X) is complete and is therefore a Banach algebra. 1.1.8. Eample. The algebra Mn(e) of n x n-matrices with entries in e is identified with B( en). It is therefore a unital Banach algebra. Recall that an upper triangular matrix is one of the form All Al2 Al n o A22 A2n o 0 A33 A3n o 0 0 Ann (all entries below the main diagonal are zero). These matrices form a sub algebra of Mn(e). We shall be seeing many more examples of Banach algebras as we proceed. Most often these will be non-abelian, but in the first three sections of this chapter we shall be principally concerned with the abelian case. If (B).)).EA is a family of subalgebras of an algebra A, then n).EAB). is a subalgebra, also. Hence, for any subset S of A, there is a smallest subal- gebra B of A containing S (namely, the intersection of all the subalgebras 
4 1. Elementary Spectral Theory containing S). This algebra is called the sub algebra of A generated by S. If S is the singleton set {a}, then B is the linear span of all powers an (n = 1,2,...) of a. If A is a normed algebra, the closed algebra G gated by a set S is the smallest closed subalgebra containing S. It is plain that G = B, where B is the sub algebra generated by S. If A = G(T), where T is the unit circle, and if z: T -+ C is the inclusion function, then the closed algebra generated by z and its conjugate z is G(T) itself (immediate from the Stone-Weierstrass theorem). A left (respectively, right) ideal in an algebra A is a vector subspace I of A such that a E A and bEl => ab E I (respectively, ba E I). An ideal in A is a vector subspace that is simultaneously a left and a right ideal in A. Obviously, 0 and A are ideals in A, called the trivial ideals. A maximal ideal in A is a proper ideal (that is, it is not A) that is not contained in any other proper ideal in A. Maximal left ideals are defined similarly. An ideal I is modular if there is an element u in A such that a - au and a - ua are in I for all a E A. It follows easily from Zorn's lemma that every proper modular ideal is contained in a maximal ideal. If w is an element of a locally compact Hausdorff space n, and Mw = {f E Co(n) I f(w) = OJ, then Mw is a modular ideal in the algebra Co(n). This is so because there is an element u E Co(n) such that u(w) = 1, and hence, f - uf E MWJ for all f E Go(n). Since MWJ is of co dimension one in Co(n) (as M EB eu = Co(n)), it is a maximal ideal. If I is an ideal of A, then AI I is an algebra with the multiplication given by (a + 1)( b + I) = ab + I. If I is modular, then AI I is unital (if a - au, a - ua E I for all a E A, then u + I is the unit). Conversely, if AI I is unital then I is modular. If A is unital, then obviously all its ideals are modular, and therefore, A posesses maximal ideals. If (I..\)..\EA is a family of ideals of an algebra A, then n..\EAI..\ is an ideal of A. Hence, if S C A, there is a smallest ideal I of A containing S. We call I the ideal generated by S. If A is a normed algebra, then the closure of an ideal is an ideal. The closed ideal J generated by a set S is the smallest closed ideal containing S. It is clear that J is the closure of the ideal generated by S. 
1.2. The Spectrum and the Spectral Radius 5 1.1.1. Theorem. If I is a closed ideal in a normed algebra A, then AI I is a normed algebra when endowed with the quotient norm lIa + III = inf lIa + bll. bEl Proof. Let £ > 0 and suppose that a, b belong to A. Then £ + lIa + III > lIa + a'il and £ + lib + III > lib + b'lI for some a', b' E I. Hence, (£ + lIa + 111)(£ + lib + III) > lIa + a'lIlIb + b'lI > lIab + ell, where e = a'b+ ab' + a'b' E I. Thus, (e + lIa + 111)(£ + IIb+ III) > lIab + III. Letting £ -+ 0, we get lIa + III lib + III > Ilab + III; that is, the quotient norm is submultiplicative. 0 A homomorphi3m from an algebra A to an algebra B is a linear map I.p: A -+ B such that <pC ab) = I.p( a )I.p( b) for all a, b E A. Its kernel ker( <p) is an ideal in A and its image <p(A) is a subalgebra of B. We say I.p is unital if A and B are unital and <p(1) = 1. If I is an ideal in A, the quotient map 7r: A -+ AI I is a homomorphism. If <p,,,p are continuous homomorphisms from a normed algebra A to a normed algebra B, then <p = "p if <p and "p are equal on a set S that generates A as a normed algebra (that is, A is the closed algebra generated by S). This follows from the observation that the set {a E A I <pC a) = "p( a)} is a closed subalgebra of A. If A is the disc algebra and A ED, the function A -+ e, f  f(A), is a continuous homomorphism. Moreover, every non-zero continuous homo- morphism from A to C is of this form. This follows from the fact that the closed subalgebra generated by the unit and the inclusion function z:D -+ e is A. We show this: If f E A and 0 < r < 1, define fr E C(D) by fr(A) = f(rA). By uniform continuity of I on D, we have limrl- III - Irlloo = o. Since Ir is extendable to an analytic function on the open disc of center 0 and radius 1/r, it is the uniform limit on D of its Taylor series. Thus, fr is the uniform limit of polynomial functions on D, and therefore, so is f. 1.2. The Spectrum and the Spectral Radius Let e[z] denote the algebra of all polynomials in an indeterminate z with complex coefficients. If a is an element of a unital algebra A and p E C[z] is the polynomial p = Ao + AIZ1 + ... + AnZ n , 
6 1. Elementary Spectral Theory we set p(a) = Ao1 + AlaI +... + Ana n . The map C[z] -+ A, p  p(a), is a unital homomorphism. We say that a E A is invertible if there is an element b in A such that ab = ba = 1. In this case b is unique and written a-I. The set Inv(A) = {a E A I a is invertible} is a group under multiplication. We define the 3pectrum of an element a to be the set a(a) = a A(a) = {A Eel Al - a ft: Inv(A)}. We shall henceforth find it convenient to write Al simply as A. 1.2.1. Ezample. Let A = C(f2), where f! is a compact Hausdorff space. Then a(f) = f(f!) for all f E A. 1.2.2. Ezample. Let A = £oo(S), where S is a non-empty set. Then a(f) = (f(S))- (the closure in C) for all f E A. 1.2.3. Ezample. Let A be the algebra of upper triangular n x n-matrices. If a E A, say All A 1 2 Al n o A22 A2n o 0 Ann a= it is elementary that a(a) = {All, A22'...' Ann}. Similarly, if A = Mn(C) and a E A, then a(a) is the set of eigenvalues of a. Thus, one thinks of the spectrum as simultaneously a generalisation of the range of a function and the set of eigenvalues of a finite square matrix. 1.2.1. Remark. If a, b are elements of a unital algebra A, then 1 - ab is invertible if and only if 1-ba is invertible. This follows from the observation that if 1 - ab has inverse c, then 1 - ba has inverse 1 + bca. A consequence of this equivalence is that a(ab) \ {O} = a(ba) \ {O} for all a, b E A. 
1.2. The Spectrum and the Spectral Radius 7 1.2.1. Theorem. Let a be an element of a unital algebra A. If O'(a) is non-empty and p E C[z], then O'(p( a)) = p( 0'( a)). Proof. We may suppose that p is not constant. If f.-l E e, there are elements Ao, . . . , An in C, where Ao =I- 0, such that p - f.-l = Ao(z - AI)... (z - An), and therefore, p( a) - f.-l = Ao (a - AI) . . . (a - An). It is clear that p( a) - f.-l is invertible if and only if a - AI, . . . , a - An are. It follows that f.-l E O'(p( a)) if and only if f.-l = p( A) for some A E 0'( a), and therefore, O'(p( a)) = p( 0'( a)). 0 The spectral mapping property for polynomials is generalised to con- tinuous functions in Chapter 2, but only for certain elements in certain algebras. There is a version of Theorem 1.2.1 for analytic functions and Banach algebras (see [Tak, Proposition 2.8], for example). We shall not need this, however. 1.2.2. Theorem. Let A be a unital Banach algebra and a an element of A such that "all < 1. Then 1 - a E Inv(A) and 00 (1- a)-1 = Lan. n=O Proof. Since E=o Ilanll < E=o lIall n = (1 - Il a ll)-1 < +00, the series E=o an is convergent, to b say, in A, and since (1 - a)(l + . . . + an) = 1 - a n + 1 converges to (1 - a)b = b(l - a) and to 1 as n  00, the element b is the inverse of 1 - a. 0 The series in Theorem 1.2.2 is called the Neumann series for (1- a)-I. 1.2.3. Theorem. If A is a unital Banach algebra, then Inv(A) is open in A, and the map Inv(A)  A, a  a-I , is differentiable. Proof. Suppose that a E Inv(A) and lib-ail < "a-III-I. Then Ilba- 1 -111 < lib - alllla- 1 " < 1, so ba- 1 E Inv(A), and therefore, b E Inv(A). Thus, Inv(A) is open in A. 
8 1. Elementary Spectral Theory If b E A and Ilbll < 1, then 1 + b E Inv(A) and 00 00 11(1 + b)-1 -1 + bll = II L(-l)nb n -1 + bll = II L(-l) nbn ll n=O n=2 00 < L IIbli n = IIbIl 2 /(1 -lIbll)-1 . n=2 Let a E Inv(A) and suppose that Ilell < !lIa- 1 11- 1 . Then lIa- 1 ell < 1/2 < 1, so ( with b = a-I e), 11(1 + a- 1 e)-1 - 1 + a- 1 ell < lIa-1eIl2/(1 -lIa-1ell)-1 < 21Ia- 1e Il 2 , since 1 -lIa-1ell > 1/2. Now define u to be the linear operator on A given by u(b) = -a- 1 ba- 1 . Then, II (a + e) -1 - a-I - u ( e) II = II (1 + a-I e) -1 a-I - a-I + a-I ea -111 < 11(1 + a- 1 e)-1 -1 + a- 1 elllla- 1 11 < 2(lIa- 1 1131IeIl 2 ). Consequently, lim Ilea + C)-I - a-I - ue c)1I = 0, c--+O " e" and therefore, the map a: b  b- 1 is differentiable at b = a with derivative a'(a) = u. 0 The algebra e[z] is a normed algebra where the norm is defined by setting Ilpll = sup Ip(A)I. 111 Observe that Inv(C[z]) = e \ {OJ, so the polynomials Pn = 1 + z/n are not invertible. But limn--+ooPn = 1, which shows that Inv(e[z]) is not open in C[z]. Thus, the norm on C[z] is not complete. 1.2.4. Lemma. Let A be a unital Banach algebra and let a E A. The spectrum a( a) of a is a closed subset of the disc in the plane of centre the origin and radius lIall, and the map C\a(a)-+A, A(a-A)-I, is differentiable. Proof. If IAI > lIall, then IIA- 1 all < 1, so 1 - A- 1 a is invertible, and therefore, so is A-a. Hence, A rt. a(a). Thus, A E a(a) => IAI < Iiali. The set a(a) is closed, that is, C \ a(a) is open, because Inv(A) is open in A. Differentiability of the map A  (a - A)-1 follows from Theorem 1.2.3. 0 The following result can be thought of as the fundamental theorem of Banach algebras. 
1.2. The Spectrum and the Spectral Radius 9 1.2.5. Theorem (Gelfand). If a is an element of a unital Banach algebra A, then the spectrum 0'( a) of a is non-empty. Proof. Suppose that a( a) = 0 and we shall obtain a contradiction. If IAI > 211all, then IIA- 1a ll < t, and therefore, l-IIA- 1a ll > t. Hence, 00 11(1- A- 1 a)-1 -111 = II L(A- 1 a)nll n=1 IIA -lall 1 < 1 _ 11,\ -lall < 211,\ - all < 1. Consequently, 11(1 - A -la)-111 < 2, and therefore, II(a - A)-III = IIA- 1 (1- A- 1 a)-111 < 2/IAI < lIall- 1 (a i= 0 since a(a) = 0). Moreover, since the map A  (a - A)-1 is contin- uous, it is bounded on the (compact) disc 211a1lD. Thus, we have shown that this map is bounded on all of C; that is, there is a positive number M such that II( a - A)-III < M (A E C). If TEA *, the function A  T( (a - A) -1) is entire, and bounded by MIITII, so by Liouville's theorem in complex analysis, it is constant. In particular, T( a-I) = T( (a - 1) -1). Because this is true for all TEA *, we have a-I = (a - 1)-1, so a = a-I, which is a contradiction. 0 It is easy to see that there are algebras in which not all elements have non-empty spectrum. For example, if e( z) denotes the field of quotients of e[z], then C(z) is an algebra, and the spectrum of z in this algebra is empty. 1.2.6. Theorem (Gelfand-Mazur). If A is a unital Banach algebra in which every non-zero element is invertible, then A = Cl. Proof. This is immediate from Theorem 1.2.5. 0 If a is an element of a unital Banach algebra A, its spectral radius is defined to be r(a) = sup IAI. .;\EO'( a) By Remark 1.2.1, r(ab) = r(ba) for all a,b EA. 1.2.4. Eample. If A = C(n), where n is a compact Hausdorff space, then r(f) = Ilflloo (f E A). 1.2.5. Eample. Let A = M 2 (C) and a=( ). Then lIall = 1, but r( a) = 0, since a 2 = o. 
10 1. Elementary Spectral Theory 1.2.7. Theorem (Beurling). lfa is an element afa unital Banach algebra A, then r( a) == inf lIa n IIl/n == lim lIa n Ill/n. n>l n-+oo Proof. If A E a(a), then An E a(a n ), so IAnl < lIanll, and therefore, rea) < infnl lIanll l / n < liminf n -+ oo Ilanll l / n . Let  be the open disc in e centered at 0 and of radius 1/ r( a) (we use the usual convention that 1/0 == +00). If A E , then 1 - Aa E Inv(A). If r E A*, then the map f:   C, A  r((l - Aa)-l), is analytic, so there are unique complex numbers An such that 00 f(A) == L An An (A E ). n=O However, if IAI < l/llall( < l/r( a)), then IIAal1 < 1, so 00 (1 - Aa) -1 == LAn an , n=O and therefore, 00 f(A) == L Anr(a n ). n=O It follows that An == r( an) for all n > o. Hence, the sequence (r( an)A n) converges to 0 for each A E , and therefore a fortiori, it is bounded. Since this is true for each r E A*, it follows from the principle of uniform boundedness that (A na n ) is a bounded sequence. Hence, there is a positive number M (depending on A, of course) such that II A n an II < M for all n > 0, and therefore, Ilanll l / n < Ml/n/IAI (if A f= 0). Consequently, lim SUPn-+oo lIanll l / n < l/IAI. We have thus shown that if r( a) < IA -11, then lim SUPn-+oo lIanll l / n < IA -11. It follows that lim SUPn-+oo Ilanll l / n < r( a), and since r(a) < liminf n -+ oo Ilanll l / n , therefore rea) == lim n -+ oo Ilanll l / n . 0 1.2.6. Ezample. Let A be the set of Cl-functions on the interval [0,1]. This is an algebra when endowed with the pointwise-defined operations, and a submultiplicative norm on A is given by Ilfll == II fll 00 + Ilf' II 00 (f E A). It is elementary that A is complete under this norm, and therefore, A is a Banach algebra. Let x: [0, 1]  e be the inclusion, so x E A. Clearly, IIxnll == 1 + n for all n, so r(x) == lim(l + n)l/n == 1 < 2 = Ilxll. Recall that if 1< is a non-empty compact set in C, its complement C \ 1< admits exactly one unbounded component, and that the bounded components of e \ 1< are called the holes of I{. 
1.2. The Spectrum and the Spectral Radius 11 1.2.8. Theorem. Let B be a closed subalgebra of a unital Banach algebra A, containing the unit of A. (1) The set Inv(B) is a clopen subset of B n Inv(A). (2) For each b E B, O'A(b) C O'B(b) and 80'B(b) C 80'A(b). (3) lfb E B and O"A(b) has no holes, then O"A(b) = O"B(b). Proof. Clearly Inv(B) is an open set in B n Inv(A). To see that it is also closed, let (b n ) be a sequence in Inv(B) converging to a point b E B n Inv(A). Then (b;;l) converges to b- 1 in A, so b- 1 E B, which implies that b E Inv(B). Hence, Inv(B) is clop en in B n Inv(A). If b E B, the inclusion 0" A ( b) C 0" B( b) is immediate from the inclusion Inv(B) C Inv(A). If A E 80" B (b), then there is a sequence (An) in e \ 0" B (b) converging to A. Hence, b - An E Inv( B), and b - A  Inv( B), so b - A  Inv( A), by Condition (1). Also, b - An E Inv(A), so An E C \ O'A(b). Therefore, A E 80" A ( b). This proves Condition (2). If b E Band 0" A (b) has no holes, then C \ 0" A (b) is connected. Since C \ O'B(b) is a clop en subset of C \ O"A(b) by Conditions (1) and (2), it follows that C \ 0" A(b) = C \ O"B(b), and therefore, 0" A(b) = O"B(b). 0 1.2.7. Ezample. Let G = G(T) and let A be the disc algebra. If f E A, let c.p(f) be its restriction to T. One easily checks that the map c.p: A  G, f  c.p(f), is an isometric homomorphism onto the closed subalgebra B of C generated by the unit and the inclusion z:T  C (the equation 1Ic.p(f)lloo = Ilflioo is given by the maximum modulus principle). Clearly, O"B(Z) = O"A(Z) = D, and O"c(z) = T. Let a be an element of a unital Banach algebra A. Since 00 00 L Ila n In!!1 < L lIalln In! < 00, n=O n=O the series 2::'=0 an In! is convergent in A. We denote its sum by ea. In proving the next theorem, we shall use some elementary results concerning differentiation. Suppose that f, 9 are differentiable maps from R to A with derivatives f', g', respectively. Then f 9 is differentiable and (fg)' = fg' + f'g. (To prove this, just mimic the proof of the scalar-valued case.) If f' = 0, then f is constant. We prove this: If rEA *, then the function R  C, t  r(f(t)), is differentiable with zero derivative, and therefore, r(f(t)) = r(f(O)) for all t. Since r was arbitrary, this implies that f(t) = f(O). 
12 1. Elementary Spectral Theory 1.2.9. Theorem. Let A be a unital Banach algebra. (1) If a E A and f: R -+ A is differentiable, f(O) = 1, and f'(t) = af(t) for all t E R, then J(t) = eta for all t E R. (2) If a E A, then e a is invertible with inverse e- a , and if a, b are commut- ing elements of A, then e a + b = eae b . Proof. First we observe that if f: R -+ A is defined by f(t) = eta, then J(t) = E  0 tnan/nI, so differentiating term by term we get f'(t) = af(t). Now suppose J, 9 are any pair of differentiable maps from R to A such that f'(t) = af(t) and g'(t) = ag(t) and J(O) = g(O) = 1. Then the map h: R -+ A, t  J(t)g( -t), is differentiable with zero derivative (apply the product rule for differentiation). Hence, h(t) = 1 for all t E R. Applying this to the map t  eta, we get etae- ta = 1; in particular, eae- a = 1. It follows that if f: R -+ A is differentiable, f(O) = 1, and J'(t) = aJ(t) for all t, then f(t) = eta (set g(t) = eta and get f(t)e- ta = 1, so f(t) = eta). Now suppose that a and b are commuting elements of A and set f(t) = etae tb . Then f(O) = 1 and f'(t) = etabe tb + aetae tb (by the product rule) = (a + b )f(t). Hence, J(t) = et(a+b) for all t E R, so, in particular, e a + b = f(l) = eae b . 0 We shall see later that not every invertible element is of the form ea. If an algebra is non-unital we can adjoin a unit to it. This is very helpful in many cases, and we shall frequently make use of it, but it does not reduce the theory to the unital case. There are situations where adjoining a unit is unnatural, such as when one is studying the group algebra £1 (G) of a locally compact group G (see the addenda section of this chapter for the definition of this algebra). If A is an alg«:bra, we set A = A E8 e as a vector space. We define a multiplication on A making it a unital algebra by setting (a, A)( b, j-t) = (ab + Ab + j-ta, Aj-t). The unit is (0,1). The algebra A is called the unitization of A. The map A -+ A, a  (a, 0), is an injective homomorphism, which we use to identify A as an ideal of A. We then write a + A for (a, A). The map A -+ C, a + A  A, is a unital homomorphism with kernel A, called the canonical homomor- phism. If A is abelian, so is A. 
1.3. The Gelfand Representation 13 If A is a normed algebra, we make A into a normed algebra by setting lIa + All = lIall + I A I. Observe that A is a closed subalgebra of A, and that A is a Banach algebra if A is one. If A is a non-unital Banach algebra, then for a E A we set 0' A (a) = 0' A(a), and r(a) = sUPAEO'A(a) IAI. Note that 0 is an element of O'A(a) in this case. 1.3. The Gelfand Representation The idea of this section is to represent an abelian Banach algebra as an algebra of continuous functions on a locally compact Hausdorff space. This is an extremely useful way of looking at these algebras, but in the case of the more "complicated" algebras, the picture it presents may be of limited accuracy. We begin by proving some results on ideals and multiplicative linear functionals. 1.3.1. Theorem. Let I be a modular ideal of a Banach algebra A. If I is proper, so is its closure I. If I is maximal, then it is closed. Proof. Let u be an element of A such that a - au and a - ua are in I for all a E A. If b E I and lIu - bll < 1, then the element v = 1 - u + b is invertible in A. If a E A, then av = a - au + ab E I, so A = Av C I. This contradicts the assumption that I is proper, and shows that lIu - bll > 1 for all b E I. It follows that u fI. I, so I is proper. If I is maximal, then I = I, as 1 is a proper ideal containing I. 0 1.3.1. Remark. If L is a left ideal of a Banach algebra A, it is modular if there is an element u in A such that a - au E L for all a E A, and in this case its closure is a proper left ideal. Moreover, if L is a modular maximal left ideal, it is closed. The proofs are the same as for Theorem 1.3.1. 1.3.2. Lemma. If I is a modular maximal ideal of a unital abelian algebra A, then AI I is a field. Proof. The algebra AI I is unital and abelian, with unit u + I say. If J is an ideal of AI I and 7r is the quotient map from A to AI I, then 7r- 1 (J) is an ideal of A containing I. Hence, 7r- 1 (J) = A or I, by maximality of I. Therefore, J = AI I or o. Thus, AI I and 0 are the only ideals of AI I. Now suppose that 7r( a) is a non-zero element of AI I. Then J = 7r( a)( AI I) is a non-zero ideal of AI I, and therefore, J = AI I. Hence, there is an element b of A such that (a + I)(b + I) = u + I, so a + I is invertible. This shows that AI I is a field. 0 
14 1. Elementary Spectral Theory Note that if cp: A -+ B is a homomorphism between algebras A and B and B is unital, then cp: A -+ B, a + ,,\  cp( a) +"\, (a E A, ,,\ E C) is the unique unital homomorphism extending cp. If cp: A -+ B is a unital homomorphism between unital algebras, then cp(Inv(A)) C Inv(B), so a(cp(a)) C a(a) (a E A). A character on an abelian algebra A is a non-zero homomorphism T: A -+ e. We denote by n(A) the set of characters on A. 1.3.3. Theorem. Let A be a unital abelian Banach algebra. (1) If T E n( A), then IITII = 1. (2) The set n(A) is non-empty, and the map T  ker(T) defines a bijection from n(A) onto the set of all maximal ideals of A. Proof. If T E n(A) and a E A, then T(a) E a(a), so IT(a)1 < r(a) < Iiali. Hence, IITU < 1. Also, T(l) = 1, since T(l) = T(I)2 and T(l) f= o. Hence, liT II = 1. Let I denote the closed ideal ker( T). This is proper, since T f= 0, and I + C1 = A, since a - T( a) E I for all a E A. It follows that I is a maximal ideal of A. If Tl, T2 E n( A) and ker( Tl) = ker( T2), then for each a E A we have Tl(a - T2(a)) = 0, so Tl(a) = T2(a). Thus, Tl = T2. If I is an arbitrary maximal ideal of A, then I is closed by Theorem 1.3.1 and AI I is a unital Banach algebra in which every non-zero element is in- vertible, by Lemma 1.3.2. Hence, by Theorem 1.2.6 AI I = C(l + I). It follows that A = IE8C1. Define T:A -+ C by T(a+"\) ="\, (a E I, ,,\ E C). Then T is a character and ker( T) = I. Thus, we have shown that the map T  ker( T) is a bijection from the characters onto the maximal ideals of A. We have seen already that A admits maximal ideals (since it is unital). Therefore, n(A) i= 0. 0 1.3.4. Theorem. Let A be an abelian Banach algebra. (1) If A is unital, then a(a) = {T(a) IT E n(A)} (a E A). (2) If A is non-tmital, then a(a) = {T(a) IT E n(A)} U {OJ (a E A). Proof. If A is unital and a is an element of A whose spectrum contains "\, then the ideal I = (a - "\)A is proper, so I is contained in a maximal ideal 
1.3. The Gelfand Representation 15 ker(r), where r E O(A). Hence, r(a) = A. This shows that the inclusion a( a) C {r( a) IrE O( A)} holds, and the reverse inc!usion is clear. Now suppose that A is non-unital, and let roo: A --+ e be the canonical homomorphism. Then O(A) = {f IrE O(A)}U{r oo }, where f is the unique character on A extending the character r on A. Hence, by Condition (1), a(a) = aA(a) = {r(a) IrE O(A)} = {r(a) IrE n(A)} U {OJ for each a E A. 0 If A is an abelian Banach algebra, it follows from Theorem 1.3.4 that O(A) is contained in the closed unit ball of A*. We endow O(A) with the relative weak* topology, and call the topological space O(A) the character space, or spectrum, of A. 1.3.5. Theorem. If A is an abelian Banach algebra, then O(A) is a locally compact Hausdorff space. If A is unital, then O(A) is compact. Proof. It is easily checked that O(A) U {OJ is weak* closed in the closed unit ball S of A*. Since S is weak* compact (Banach-Alaoglu theorem), O(A) U {OJ is weak* compact, and therefore, O(A) is locally compact. If A is unital, then O(A) is weak* closed in S and thus compact. 0 Note that O(A) may be empty. This is the case for A = 0, for example. Suppose that A is an abelian Banach algebra for which the space O(A) is non-empty. If a E A, we define the function a. by a.: O(A) --+ C, r t-+ r(a). Clearly the topology on O(A) is the smallest one making all of the functions a continuous. The set {r E O(A) Ilr(a)1 > £} is weak* closed in the closed unit ball of A* for each £ > 0, and weak* compact by the Banach-Alaoglu theorem. Hence, a E C o (f2(A)). We call a the Gelfand transform of a. Although the following result is very important, its proof is easy, be- cause we have already done most of the work needed to demonstrate it. 1.3.6. Theorem (Gelfand Representation). Suppose that A is an abelian Banach algebra and that O( A) is non-empty. Then the map A --+ Co(O(A)), a t-+ a, is a norm-decreasing homomorphism, and r(a) = lIali oo (a E A). If A is unital, a(a) = a(O(A)), and if A is non-unital, u(a) = a(O(A))U{O}, for each a E A. 
16 1. Elementary Spectral Theory Proof. By Theorem 1.3.4 the spectrum 0-( a) is the range of a, together with {OJ if A is non-unital. Hence, r(a) = lIall oo , which implies that the map a  a is norm-decreasing. That this map is a homomorphism is easily checked. 0 The kernel of the Gelfand representation is called the radical of the algebra A. It consists of the elements a such that r( a) = o. It therefore contains the nilpotent elements. If the radical is zero, A is said to be 3emi3imple. In a general algebra an element whose spectrum consists of the set {OJ is said to be qua3inilpotent. Let a, b be commuting elements of an arbitrary Banach algebra A. Then r(a + b) < r(a) + r(b), and r(ab) < r(a)r(b). To see this, we may suppose that A is unital and abelian (if necessary, adjoin a unit and restrict to the closed subalgebra generated by 1, a, and b). Then r(a + b) = lI(a + b)"lIoo < lIali ce + IIbli oo = rea) + r(b) by Theorem 1.3.6. Similarly, r(ab) = lI(ab)"lIoo < lIalloollblloo = r(a)r(b). Direct proofs of the first of these inequalities (that is, where the Gelfand representation is not invoked) tend to be messy. The spectral radius is neither subadditive nor submultiplicative in gen- eral: Let A = M 2 (C) and suppose a = ( ) and b=( ). Then rea) = r(b) = 0, since a and b have square zero, but r(a + b) r(ab) = 1. The interpretation of the character space as a sort of generalised spec- trum is motivated by the following result. 1.3.7. Theorem. Let A be a unital Banach algebra generated by 1 and an element a. Then A is abelian and the map a: f2(A)  o-(a), T  T(a), is a homeomorphism. Proof. It is clear that A is abelian and that a is a continuous bijection, and because f2( A) and 0-( a) are compact Hausdorff spaces, a is therefore a homeomorphism. 0 To illustrate this, consider the disc algebra A. If z is its canonical generator, then since o-(z) = D, we have f2(A) = D by Theorem 1.3.7. In this case if f E A, then j(A) = f(A), so the Gelfand transform is the identity map. We now present an interesting application of the preceding results to a problem in classical analysis. 
1.3. The Gelfand Representation 17 1.3.1. Ezample. We denote by f 1 (Z) the set of all complex-valued func- tions f on Z such that E=-oo If(n)1 is finite. This is a Banach space when endowed with the pointwise-defined operations and the norm 00 IIfllt = 2: If(n)l. n=-oo If f, 9 E /,1 (Z) we define their convolution f * g: Z  C by the formula 00 (f * g)(m) = 2: f(m - n)g(n). (1) n=-oo If f E ll(Z), it is bounded, so the sum in Eq. 1 exists. To see that f * 9 E ll(Z) observe that 00 00 00 2: I(f * g)(n)1 = 2: I 2: fen - m)g(m)1 n=-oo n=-oo m=-oo 00 00 < 2: 2: If(n - m)llg(m)1 n=-oo m=-oo 00 00 = 2: 2: If(n - m)llg(m)1 m=-oo n=-oo mf;oo (lg(m)1 n oo If(n - m)l) 00 2: Ig(m)llIfllt = IIf1l111g111. m=-oo Thus, f * 9 E ll(Z) and IIf * gilt < IIfllt IIg1l1. It is now a straightforward exercise to show that /,l(Z) is an abelian unital Banach algebra with mul- tiplication given by (f, g)  f * g. The characteristic function of the set {O} is the unit, and if w is the characteristic function of the set {I}, then f = E _ -oo f(n)w n for all f E ll(Z). For z E T, define a character Tz on f 1 (Z) by setting 00 Tz(f) = 2: f(n)zn. n=-oo We then have a map T  n(f 1 (Z)), z  Tz, 
18 1. Elementary Spectral Theory which, it is easy to check, is a bijection. In fact, this is a homeomorphism, and to see this we need only show continuity, since the domain and range are compact and Hausdorff. Continuity is shown if we show the function T  e, z  Tz(f), is continuous when f E £l(Z), and this follows from the observation that Tz(f) is the uniform limit in z of the continuous functions ElnlN f(n )zn (N = 1,2,. · .), since E  -00 If(n)znl = IIfll1 < 00. We identify Q(i 1 (Z)) with T using the above homeomorphism. Thus, the Gelfand transform j of f E £l(Z) is a continuous function on T such that 00 j(z) = L f(n)zn. n=-oo It is readily verified that the numbers f( n) are the Fourier coefficients of j, f ( n) =  {21f j ( e it)e -i nt dt. 27r Jo Thus £1 (Z)", the set of all Gelfand transforms, is the set of all functions h E G(T) whose Fourier series is absolutely convergent. One can show that not every function in C(T) has such a Fourier series. A well-known theorem of Wiener states that if a continuous function on T has an absoluely convergent Fourier series and never vanishes, then its reciprocal has such a Fourier series. The proof of this is easy, using what we know about the algebra £1 (Z): Let h be a continuous function that never vanishes and that has ab- solutely convergent Fourier series, so h = j for some f E £l(Z). Because j(z) i= 0 for all z E T, it follows from Theorem 1.3.4 that 0 rt. a(f). Thus, f is invertible in £1 (Z), with inverse 9 say. Then 9 = 1/ h, so 1/ h has absolutely convergent Fourier series. This proof is due to Gelfand. We shall resume our study of Banach algebras in Chapter 2, but now we turn to single operator theory. 1.4. Compact and Fredholm Operators This section is concerned with the elementary spectral theory of oper- ators. We begin with the simplest non-trivial class of operators, the com- pact ones, a class that plays an important and fundamental role in operator theory. These operators behave much like operators on finite-dimensional vector spaces, and for this reason they are relatively easy to analyse. A linear map u: X  Y between Banach spaces X and Y is compact if u( S) is relatively compact in Y, where S is the closed unit ball of X. Equivalently, u( S) is totally bounded. In this case u( S) is bounded, and therefore, u is bounded. 
1.4. Compact and Fredholm Operators 19 1.4.1. Remark. Note that the range of a compact operator is separable. This is immediate from the fact that a compact metric space is separable, and that the closure of the image of the ball under a compact operator is compact. The theory of compact operators arose out of the analysis of linear integral equations. The following example illustrates the connection. 1.4.1. Ezample. Let [ = [0,1] and let X be the Banach space C(I), where the norm is the supremum norm. If k E C(I 2 ), define u E B(X) by setting u(f)(s) = 1 1 k(s, t)f(t) dt (f E x, s E I). We show that u(f) EX. Observe first that lu(f)( s) - u(f)( s')1 = III (k( s, t) - k( s', t))f(t) dtl < 1 1 I k ( s, t) - k( s' , t) II f ( t ) I dt < sup Ik(s, t) - k(s', t)llIfllooo tEl Now k is uniformly continuous because [2 is compact, so if c > 0, there exists 8 > 0 such that max{ls - s'l, It - t'l} < 8 => Ik(s, t) - k(s', t')1 < c. Hence, Is - s'l < 8 => lu(f)(s) - u(f)(s')1 < £11/1100. (1) Thus, u(f) is continuous, that is, u(f) EX, but more is true, for it is immediate from Inequality (1) that u(S) is equicontinuous, where S is the closed unit ball of X. Also, u( S) is pointwise-bounded, that is, sUPfES lu(f)(s)1 < 00, since I u (f) ( s ) I < 1 1 I k ( s, t) f ( t ) 1 dt < II k II 00 II f II 00 . By the Arzela-Ascoli theorem [Rud 2, Theorem A5] the set u(S) is totally bounded. Therefore, u is a compact operator on X. The function k is called the kernel of the operator u, and u is called an integral operator. A similar example is obtained if we define v E B(X) by v(f)(s) = s s' 10 f(t)dt. If s,s' E I and f E X, then Iv(f)(s)-v(f)(s')1 = I Is f(t)dtl < Is - s'lllflloo. Hence, v(S) is equicontinuous and pointwise-bounded, so by the Arzela-Ascoli theorem again, v(S) is totally bounded; that is, v is compact. Observe that v has no eigenvalues (it will follow from Theorem 1.4.11 that v is quasi nilpotent ). That v(f) = 0 => f = 0 is elementary. Suppose 
20 1. Elementary Spectral Theory then that A E e \ {OJ and f E X and v(f) = Af. Then f(O) = 0 and (by differentiation) f'(t) = Jlf(t), where Jl = 1/ A. Consequently, f(t) = f(O)e llt = 0 for all t, so f = o. The operator v is called the Volterra integral operator on X. If X, Yare Banach spaces, we denote by B(X, Y) the vector space of all bounded linear maps from X to Y. This is a Banach space when endowed with the operator norm. The set of all compact operators from X to Y is denoted by K(X, Y). The proof of the following is a routine exercise. 1.4.1. Theorem. Let X and Y be Banach spaces and U E B(X, Y). Then the following conditions are equivalent: (1) U is compact; (2) For each bounded set S in X, the set u( S) is relatively compact in Y; (3) For each bounded sequence (xn) in X, the sequence (u(xn)) admits a subsequence that converges in Y. It follows easily from Theorem 1.4.1 that [{(X, Y) is a vector subspace of B(X, Y). Also, if X'  X  Y  Y' are bounded linear maps between Banach spaces and u is compact, then wu and uv are compact. Hence K(X) = K(X, X) is an ideal in B(X). 1.4.2. Theorem. If X is a Banach space, then K(X) = B(X) if and only if X is finite-dimensional. Proof. If S denotes the closed unit ball of X, then K(X) = B(X) <=> id x is compact <=> S is compact <=> X is finite-dimensional. 0 1.4.3. Theorem. If X, Y are Banach spaces, then K(X, Y) is a closed vector space of B(X, Y). Proof. We show that if a sequence (un) in K(X, Y) converges to an operator u in B(X, Y), then u is compact. Let S denote the closed unit ball of X and let c > o. Choose an integer N such that IluN - ull < £/3. Since UN(S) is totally bounded, there are elements Xl,. .., X n E S, such that for each X in S, the inequality IluN(X) - UN(X j)1I < £/3 holds for some index j. Hence, lIu(X) - u(Xj)1I < lIu(x) - uN(x)1I + IIUN(X) - uN(xj)1I + lIuN(Xj) - u(xj)1I < c/3 + c/3 + c/3 = c. Thus, u(S) is totally bounded, and therefore, u E K(X, Y). 0 Recall that a linear map u: X  Y is of finite rank if u(X) is finite- dimensional and that rank(u) = dim(u(X)). 
1.4. Compact and Fredholm Operators 21 If X and Y are Banach spaces and u E B(X, Y) is of finite rank, then u E K(X, Y). This is immediate from the fact that the closed unit ball of the finite-dimensional space u(X) is compact. It follows from this remark and Theorem 1.4.3 that norm-limits of finite-rank operators are compact, and it is natural to ask whether the con- verse is true. This is the case for Hilbert spaces, as we shall see in the next chapter, but it is not true for arbitrary Banach spaces. P. Enflo [Enf] has given an example of a Banach space for which there are compact operators that are not norm-limits of finite-rank operators. If u: X -+ Y is a bounded linear map between Banach spaces, we define its transpose u* E B(Y*,X*) by u*(r) = r 0 u. 1.4.4. Theorem. Let X, Y be Banach spaces and let u E K(X, Y). Then u* E K(Y*,X*). Proof. Let S be the closed unit ball of X and let e > O. Since u( S) is totally bounded, there exist elements x I, . . . , X n in S, such that if xES, then lIu(x )-u(xi)1I < e/3 for some index i. Define v E B(Y*, C n ) by setting v( r) = (ru( Xl), . . . , ru( x n )). Since the rank of v is finite, v is compact, and therefore v(T) is totally bounded, where T is the closed unit ball of Y*. Hence, there exist functionals rl,..., r m in T, such that if r E T, then IIv( r) - v( rj )11 < e/3 for some index j. Observe that Ilv(r)-v(rj)lI= m!iX lu*(r)(xi)-u*(rj)(xi)l. 1In Now suppose that xES. Then lIu(x) - u(xi)1I < e/3 for some index i, and lu*(r)(xi) - u*(rj)(xi)1 < e/3. Hence, lu*(r)(x) - u*(rj)(x)1 < lu*(r)(x) - u*(r)(xi)1 + lu*(r)(xi) - u*(rj)(xi)1 + lu*(rj)(xi) - u*(rj)(x)1 < e/3 + e/3 + e/3 = e. It follows that lIu*(r) - u*(rj)1I < e, so u*(T) is totally bounded and there- fore u* is compact. 0 A linear map u: X -+ Y between Banach spaces is bounded below if there is a positive number 8 such that lIu(x)1I > 811xll (x E X). Note that in this case u(X) is necessarily closed, for if (u(x n )) is a Cauchy sequence in u(X), then (x n ) is a Cauchy sequence in X and therefore converges to some element x EX, because X is complete. Hence, the sequence (u( X n)) converges to u(x) by continuity of u. Thus, u(X) is complete and therefore closed in Y. Observe that every invertible linear map is bounded below, as is every isometric linear map. It is easily checked that u: X -+ Y is not bounded below if and only if there is a sequence of unit vectors (x n ) in X such that lim n -. oo u(x n ) = o. These remarks will be used in the following theorem. 
22 1. Elementary Spectral Theory 1.4.5. Theorem. Let u be a compact operator on a Banach space X and suppose that A E e \ {O}. (1) The space ker( u - A) is nni te-dimensional. (2) The space (u - A)(X) is closed and nnite-codimensional in X. (In fact, the codimension of (u - A)(X) in X is the dimension of ker( u* - A).) Proof. Let Z = ker(u - A). Then u(Z) C Z, and the restriction Uz of u to Z is in K( Z). Since Uz = A idz and A i= 0, the map id z is compact. Hence, Z is finite-dimensional by Theorem 1.4.2. Because Z is finite-dimensional, there is a closed vector space Y in X such that Z EB Y = X. Observe that (u - A)X = (u - A)Y, so to show that (u - A)X is closed in X it suffices to show that the restriction ( u - A)y: Y ---+ X is bounded below. Suppose otherwise, and we shall obtain a contradiction. There is a sequence (x n ) of unit vectors in Y such that limn-+oc> Ilu(x n ) - AX n II = O. Using the compactness of u and going to a subsequence if necessary, we may suppose that (u( x n)) is convergent. It follows from the equation x n = A -1 ( u( X n) - ( u - A)( x n)) that the sequence (xn) is convergent, to x say, and, since Y is closed in X, it contains x. Obviously, u( x) = Ax, so x E Y n ker( u - A) and therefore x = o. How- ever, x is the limit of unit vectors and is therefore itself a unit vector, a contradiction. This shows that (u - A)y is bounded below. Now let W = X/(u - A)(X). To show that (u - A)(X) is finite- co dimensional in X, we have to show W is finite-dimensional, and we do this by showing W* is finite-dimensional. Let 7r: X ---+ W be the quo- tient map. It is clear that the image of 7r* is contained in the kernel of u* - A. In fact these spaces are equal. For suppose that ()" E ker( u* - A). Then ()" annihilates (u - A)( X) and therefore induces a bounded linear functional r: W ---+ e such that ()" = r 0 7r = 7r*( r). Since u* is com- pact by Theorem 1.4.4, ker( u* - A) is finite-dimensional by the first part of this proof. Thus, 7r* has finite-dimensional range, and clearly 7r* is in- jective, so W* is finite-dimensional, and therefore dim(W) = dim(W*) = dim(7r*(W*)) = dim(ker(u* - A)). 0 If u: X ---+ X is a linear map on a vector space X, then the sequence of spaces (ker( un)) is clearly increasing. If ker( un) i= ker( u n+ 1 ) for all n EN, we say that u has infinite a3cent and set ascent( u) = +00. Otherwise we say u has finite a3cent and we define ascent ( u) to be the least p such that ker( uP) = ker( u P + 1 ). In this case, ker( uP) = ker( un) for all n > p. The sequence of spaces (un(X)) is decreasing. We say that u has infinite de3cent, and we set descent ( u) = +00, if un(X) i= u n + 1 (X) for all n E N. Otherwise, we say that u has finite de3cent and we define descent(u) to be the least pEN such that uP(X) = u p + 1 (X). In this case uP(X) = un(X) for all n > p. We recall now a theorem of F. Riesz from elementary functional anal- 
1.4. Compact and Fredholm Operators 23 ysis [Rud 2, Lemma 4.22]: If Y is a proper closed vector subspace of a normed vector space X and c > 0, then there exists a unit vector x E X such that Ilx + YII > 1 - c. This simple result plays a key role in the theory of compact operators. (The result as stated here is a slight reformulation of Lemma 4.22 of [Rud 2].) 1.4.6. Theorem. Let u be a compact operator on a Banach space X and suppose that ,,\ E e \ {O}. Then u - ,,\ has finite ascent and descent. Proof. Suppose the ascent is infinite, and we deduce a contradiction. If N n = ker( u - ,,\)n, then N n-1 is a proper subspace of N n, and therefore, by the theorem of Riesz discussed earlier, there is a unit vector X n E N n such that IIxn + N n - 1 11 > 1/2. If m < n, then U(Xn) - u(xm) = "\xn + (u - "\)(xn) - (u - "\)(xm) - "\xm = "\xn - z, where z E N n - 1 . Hence, lIu(xn) - u(xm)1I = lI"\x n - zll = 1,,\llIx n -,,\ -1 zll > 1,,\1/2 > o. It follows that (u(xn)) has no convergent subsequence, contra- dicting the compactness of u. Consequently, ascent ( u) < +00. The proof that u - ,,\ has finite descent is completely analogous and is left as an exercise. 0 We shall have more to say about compact operators presently. One can give direct proofs of these later results, but the details are tedious and a little messy, whereas when one uses the homomorphism property of the Fredholm index, which we are now going to introduce, they drop out very nicely. The index and the essential spectrum, which we shall also introduce, are indispensible items in the operator theorist's tool-kit. Nevertheless, many of the proofs in Fredholm theory are elementary (although often neither trivial nor obvious). Let X, Y be Banach spaces and u E B(X, V). We say u is Fredholm if ker( u) is finite-dimensional and u( X) is finite-codimensional in Y. We define the nullity of u to be dim(ker( u)) and denote it by nul( u). The defect of u is the co dimension of u(X) in Y, and is denoted by def(u). The index of u is defined to be ind(u) = nul(u) - def(u). The index is a very simple prototype of the application of algebraic topological methods to this subject. The connecting homomorphism in the K-theory of Banach algebras (to be introduced in Chapter 7) can be thought of as a generalised Fredholm index. Note that because there is a finite-dimensional (and therefore closed) vector subspace Z of Y, such that u(X) EB Z = Y, it is a consequence of the following theorem that u(X) is closed in Y. 
24 1. Elementary Spectral Theory 1.4.7. Theorem. Let X, Y be Banach spaces and u E B(X, Y). Suppose that there is a closed vector subspace Z of Y such that u(X) EB Z = Y. Then u(X) is closed in Y. Proof. The bounded linear map X/ker(u)  Y, x +ker(u)  u(x), has the same range as u and is injective, so we may suppose without loss of generality that u is injective. The map v:XEBZY, (x,z)u(x)+z, is a continuous linear isomorphism between Banach spaces, so by the open mapping theorem, v-I is also continuous. If x E X, then IIxll = IIv-Iu(x)11 < IIv-Illllu(x)lI, so Ilu(x)11 > IIv- I II-1I1 x ll. Thus, u is bounded below, and therefore u(X) is closed in Y. 0 The following theorem is a fundamental result of Fredholm theory. 1.4.8. Theorem. Let X  Y  Z be Fredholm linear maps between Banach spaces X, Y, Z. Then vu is Fredholm and ind(vu) = ind(v) + ind(u). Proof. Set Y 2 = ker( v) nu( X) and choose suitable closed vector subspaces Y I , Y 3 , Y 4 of Y, such that u(X) = Y 2 EB Y 3 , ker(v) = Y 1 EB Y 2 , and Y = Y I EB u(X) EB Y 4 . Note that Y I , Y 2 , Y 4 are finite-dimensional. The map ker(vu)  Y 2 , X  u(x), is surjective and it has the same kernel as u, so the kernel of vu is finite- dimensional and nul(vu) = nul(u) + dim(Y 2 ). Since v(Y) = v(Y 3 ) EB v(Y 4 ) and v(Y 3 ) = vu(X), therefore v(Y) = vu(X) EB V(Y4). Choose a finite-dimensional vector subspace Z' of Z such that v(Y) EB Z' = Z, so Z = vu(X) EB v(Y 4 ) EB Z'. Because v(Y 4 ) EB z' is finite-dimensional, vu(X) is finite-codimensional in Z. Therefore, vu is a Fredholm operator. The map Y4  v(Y 4 ), y  v(y), is a linear isomorphism, so dim(Y 4 ) = dim(v(Y 4 )). Hence, def(vu) = dim(Y 4 ) + dim(Z') = dim(Y 4 ) + def(v). Consequently, nul(vu) + def(u) + def(v) = nul(u) +dim(Y 4 ) +nul(v)+def(v) = nul(u) +nul(v) +def(vu), and therefore, ind( vu) = nul( vu) - def( vu) = nul( u) + nul( v) - def( u) - def( v) = ind( u) + ind( v). 0 We give an immediate easy application of the index: 
1.4. Compact and Fredholm Operators 25 1.4.9. Theorem. Let u be a compact operator on a Banach space X, and let -X E C \ {o}. (1) The operator u - -X is Fredholm of index zero. (2) H p denotes the (finite) ascent of u - -X, then X = ker(u - -X)P E8 (u - -X)P(X). Proof. That u--X is Fredholm follows from Theorem 1.4.5, and the ascent and descent of u - -X are finite by Theorem 1.4.6. If we suppose that m,n are integers greater than max{ascent(u - -X),descent(u - A)}, then we have nul(u - -X)m = nul(u - -x)n and def(u - -x)m = def(u - -x)n, so ind((u - -x)m) = ind((u - -x)n), and therefore m ind(u - -X) = n ind(u --X) by Theorem 1.4.8. It follows that ind( u - -X) = o. Thus, Condition (1) is proved. If x E ker(u - -X)P n (u - -X)P(X), then there is an element y E X such that x = (u - -X)P(y) and (u - -X)2p(y) = O. Since ker(u - -X)P = ker(u - -X)2 p , it follows that (u - -X)P(y) = 0; that is, x = O. Moreover, since nul( u - -X)P = def( u - -X)P, because ind( u - -X)P = 0, it follows that X = ker(u - -X)P E8 (u - -X)P(X). 0 1.4.10. Corollary (Fredholm Alternative). The operator u - -X is injective if and only if it is surjective. Proof. Since the index of u - -X is zero, the nullity is zero if and only if the defect is zero; that is, u - -X is injective if and only if it is surjective. 0 1.4.2. Remark. If Y, Z are complementary vector subspaces of a vector space X, and u, v are linear maps on Y, Z, respectively, we denote by u EB v the linear map on X given by (u E8 v)(y + z) = u(y) + v(z) (y E Y, z E Z). Clearly, u E8 v is invertible if and only if u and v are invertible. If X is a Banach space and w E B(X), we write a(w) for aB(X)(w). If Y, Z are closed complementary vector subspaces of X, and if u E B(Y), v E B(Z), w E B(X), and w = u EB v, then a(w) = a(u) U a(v), by the preceding observation. 1.4.11. Theorem. Let u be a compact operator on a Banach space X. Then a( u) is countable, and each non-zero point of a( u) is an eigenvalue of u and an isolated point of a( u). Proof. If -X is a non-zero point of a( u), then by the Fredholm alternative, Corollary 1.4.10, u - -X is not injective, and therefore -X is an eigenvalue of u. The operator u - -X has finite ascent, p say, and by Theorem 1.4.9 
26 1. Elementary Spectral Theory we can write X = Y EI1 Z, where Y = ker(u - A)P and Z = (u - A)P(X). The spaces Y, Z are closed and invariant for u (that is, u(Y) C Y and u( Z) C Z). Hence, u - A = (uy - A idy) EI1 (uz - A idz), where Uy, Uz are the restrictions of u to Y, Z, respectively. Since (uy - A idy)P = 0, the spectrum 0'( u y) is the singleton set {A}. Also, the operator u z is compact and ker( u z - A idz)P = 0, so (u Z - A idz)P is invertible (as it is injective and Fredholm of index zero), and therefore Uz - A id z is invertible. Hence, A ft 0'( U z). This implies that 0'( u) \ {A} = 0'( u z), so A is an isolated point of 0'( u) because 0'( u z) is closed in 0'( u). Countability of 0'( u) follows by elementary topology. 0 1.4.2. Eample. Let us interpret our resullts now in terms of integral equations. Let I = [0,1] and suppose k E G(I 2 ). Consider the integral equation 1 1 k( 8, t)f(t) dt - >.f(8) = g( 8). Here A is a non-zero scalar, 9 E G(l) is a known function, and f E G(l) is the unknown. If u is the compact integral operator corresponding to the kernel k, as in Example 1.4.1, then we can rewrite our equation as (u - A)(f) = g. The non-zero spectrum of u is of the form {An I 1 < n < N}, where N is an integer or 00. If A i= An for all n, then the integral equation has a unique solution: f = (u - A )-1 (g). If on the other hand A = An say, then the homogeneous equation 1 1 k( 8, t) f ( t) dt - >.J ( 8) = 0 has a non-zero solution by the Fredholm alternative (Corollary 1.4.10), and by Theorem 1.4.5 the solution set is finite-dimensional. Observe that if N = 00, then lim n ..... oo An = 0 by Theorem 1.4.11. 1.4.3. Eample. One should not be misled by Theorem 1.4.11-the spec- tral behaviour of compact operators is not typical of all operators. To illustrate this, let H be a separable Hilbert space with an orthonormal basis (en)=l. If (An) is a bounded sequence of scalars, define u E B(H) by setting u(x) = 2: :'- 1 AnGnen when x = 2::'=1 Gne n . We call u the diagonal operator with diagonal (An) with respect to the basis (en). It is readily verified that II u II = sUPn I An I, and that u is invertible if and only if inf n IAnl > 0, and in this case u- 1 is the diagonal operator with respect to (en) with diagonal (A;;l). These observations imply that O'(u) is the closure of the set {An In = 1,2,.. .}. 
1.4. Compact and Fredholm Operators 27 Suppose that a non-empty compact set K in C is given and choose a dense sequence (An) in K. If u is the corresponding diagonal operator, then o-(u) = K. Thus, the spectrum is an arbitrary non-empty compact set in general. We need to consider now a few elementary results (some of which extend what we said in Remark 1.4.2). These will be used immediately for the proof of Theorem 1.4.15. A linear map p: X -+ X on a vector space X is idempotent if p2 = p. In this case X = p(X) EI1 ker(p), since ker(p) = (1 - p)(X). In the reverse direction, if X = y EI1 Z, where Y and Z are vector subspaces of X, then there is a unique idempotent p on X such that p(X) = Y and ker(p) = Z. We call p the projection of X on Y along Z. 1.4.12. Theorem. Let Y, Z be closed complementary vector subspaces of a Banach space X. Tben tbe projection p of X on Y along Z is bounded. Proof. Let (xn) be a sequence in X converging to 0 and suppose that (p(x n )) converges to a point y of X. By the closed graph theorem, p will have been shown to be bounded if we show that y = O. Now y E Y, since p(xn) E Y and Y is closed in X, and -y E Z, since X n - p(xn) E Z and -y = limn--+oo(x n - p(x n )). Hence, y E Y n Z, and therefore y = o. 0 1.4.13. Corollary. Let v E B(Y) and w E B(Z), and suppose that u = v EB w. Then u E B(X). Proof. We have to show u is continuous. Let p be the projection of X onto Y along Z. Suppose that (Xn) is a sequence in X converging to a point x. Then (u(xn)) = (vp(x n ) + w(l - p)(xn)) converges to u(x) vp(x) + w(l - p)(x) by continuity of v, w, and p. 0 1.4.14. Theorem. Let u: X -+ Y be a linear map between normed vec- tor spaces X and Y and suppose that X is finite-dimensional. Then u is bounded. Proof. Define a new norm on X by setting IIX II' = max( Ilx II, Ilu( x) II). Then 11.11' is equivalent to the original norm on X (because all norms on finite-dimensional vector spaces are equivalent). This shows that u is bounded. 0 If u is a linear map between vector spaces, a pseudo-inverse of u is a linear map v: Y -+ X such that uvu = u. Observe that uv and vu are idempotents, that ker(vu) = ker(u), and that uv(Y) = u(X). 
28 1. Elementary Spectral Theory 1.4.15. Theorem. Let X, Y be Banach spaces and let u E B(X, Y) be a Fredholm operator. Then u admits a pseudo-inverse v that is Fredholm and is such that 1 - uv and 1 - vu are of finite rank. Moreover, if ind( u) = 0, we may choose v to be invertible. Proof. Choose a closed vector subspace Xl of X such that ker(u)E8X I = X and a finite-dimensional vector subspace Y I of Y such that u(X) E8 Y I = Y. The restriction UI:X I -+ u(X), X  u(x), is a continuous linear isomorphism, so by the open mapping theorem its inverse VI: u(X) -+ Xl is also continuous. Let v: Y -+ X be the linear map defined as follows: On u(X), v = VI; and on Y I v= { O, ifind(u)#0 w, if ind( u) = 0, where W is a linear isomorphism of Y 1 onto ker( u) (such an isomorphism exists if ind( u) = 0). It is easily checked that v is continuous and that uvu = u. Now ker(v) C Y I and Xl C v(Y), so v has finite nullity and defect and is therefore a Fredholm operator. Because (1 - vu)( X) = ker( vu) = ker( u), the idempotent 1 - vu is of finite rank. Also, uv(Y) = u(X), so (1 - uv)(Y) C (1 - uv)(Y I ), and therefore 1 - uv is of finite rank. If we now suppose that ind(u) = 0, then v(Y) = X, and ker(v) C Y I , so ker( v) = o. Hence, v is invertible. 0 The following characterisation of Fredholm operators is extremely use- ful. Note incidentally that all operators on a finite-dimensional vector space are Fredholm, so that in this case Fredholm theory is degenerate. Thus, we shall be interested only in infinite-dimensional spaces for these operators. 1.4.16. Theorem (Atkinson). Let X be an infinite-dimensional Banach space and let u E B(X). Then u is Fredholm if and only if u + K(X) is invertible in the quotient algebra B(X)/ K(X). Proof. Let 7r be the quotient homomorphism from B(X) to B(X)/ K(X). If u is Fredholm, then by Theorem 1.4.15 there is a Fredholm operator v in B(X) such that 1 - vu and 1- uv are of finite rank and therefore compact. Hence, 0 = 7r(1- uv) = 1- 7r(u)7r(v) and 0 = 7r(1- vu) = 1- 7r(v)7r(u), so 7r( u) is invertible in B(X)/ K(X). Conversely, suppose 7r( u) is invertible, with inverse 7r( v). Then uv = l+WI andvu = 1+w2, wherewI,w2 E K(X). Clearlyker(u) C ker(1+w2)' and ker(l + W2) is finite-dimensional by Theorem 1.4.5, so nul(u) < +00. Also, (l+WI)(X) = uv(X) C u(X), and (l+WI)(X) has finite codimension in X by Theorem 1.4.5. Consequently, def(u) < 00. Thus, u is Fredholm.D 
1.4. Compact and Fredholm Operators 29 1.4.17. Theorem. Let X be an infinite-dimensional Banach space and let cl) denote the set of Fredholm operators on X. Then q> is open in B(X) and the index function ind: cl)  Z, u  ind( u), is continuous. Proof. If, as usual, 7r denotes the quotient homomorphism from B(X) to B(X)/K(X), then cl) = 7r- I (Inv(B(X)/K(X))) by the Atkinson character- isation, Theorem 1.4.16. By Theorem 1.2.3, the set of invertible elements in B(X)/ K(X) is open, and therefore cl) is open in B(X) by continuity of 7r. Let u E cl) and choose v E cl), a pseudo-inverse of u, such that 1-vu and 1 - uv E K(X) (this is possible by Theorem 1.4.15). Suppose that w E <P and Ilu - wI! < IIvll- I . Then lIuv - wv II < 1, so s = 1 + wv - uv is invertible in B(X) by Theorem 1.2.2. Now u + wvu = uvu + su, so wvu = su (as u = uvu) and therefore ind(w) + ind(v) + ind(u) = ind(s) + ind(u). But ind( s) = 0, because s is invertible, so ind( w) = - ind( v). Thus, the index map is locally constant and therefore continuous. 0 1.4.18. Theorem. Let X be an infinite-dimensional Banach space, and suppose that w E K(X), that u E B(X), and that u is Fredbolm. Then ind(u + w) = ind(u). Proof. By Theorem 1.4.17, the function a: [0, 1]  Z, t  ind(u + tw), is continuous, and therefore 0[0, 1] is connected in the discrete space Z. Hence, 0[0,1] is a singleton set, so ind(u) = 0(0) = 0(1) = ind(u + w). 0 1.4.3. Remark. Let u be a Fredholm operator on an infinite-dimensional Banach space X. If u is the sum of an invertible operator and a compact operator, then by Theorem 1.4.18 ind( u) = 0, since invertible operators are of course of index zero. The converse is also true; that is, if ind( u) = 0, then u is the sum of an invertible operator and a compact operator. For by Theorem 1.4.15 there is an invertible pseudo-inverse v for u, and if we denote by 7r the quotient map from B(X) to B(X)/ K(X), the equation u = uvu implies that 7r( u) = 7r( U )7r( V )7r( u), and since 7r( u) is invertible by Theorem 1.4.16, it follows that 7r( u) = 7r( v-I). Hence, u - V-I is compact, so u is the sum of an invertible and a compact operator. Incidentally, it is easy to give examples of operators that are of index zero and not invertible (for instance, if p is a finite-rank non-zero idempotent, then 1-p is Fredholm of index zero, and non-invertible). 
30 1. Elementary Spectral Theory Again suppose X to be an infinite-dimensional Banach space and sup- pose that u E B(X). We define the e33ential 3pectrum of u to be O"e(u) = {-X Eel u - -X is not Fredholm}. Let C denote the quotient algebra B(X)/ K(X). This algebra is called the Calkin algebra on X. If 7r is the quotient map from B(X) to C, it is clear from the Atkinson characterisation (Theorem 1.4.16) that O"e( u) = O"c( 7ru). Thus, 0" e (u) is a non-empty compact set. Obviously, 0" e (u) C 0"( u). 1.4.4. Ezample. Suppose that H is a Hilbert space with an orthonormal basis (en)  1. The unilateral 3hift on this basis is the operator u in B(H) such that u(e n ) = e n +l for all n. Observe that nul(u) = 0 and def(u) = 1, so u is a Fredholm operator and ind( u) = -1. If instead we suppose that (fn)nEZ is an orthonormal basis for H, the bilateral 3hift on this basis is the operator v such that v(f n) = f n+l for all n E Z. This operator is invertible, so ind( v) = o. Hence, u and v are not similar (two elements a, b of a unital algebra are 3imilar if there is an invertible element c such that a = c- 1 bc). It follows from Theorem 1.4.16 that if 7r: B(H) -+ B(H)/ K(H) is the quotient homomorphism, then 7r( u) is invertible. It is natural to ask if one can write 7r( u) = 7r( w) for some invertible operator w in B( H). If this were the case, then ind( u) = ind( w), since u - w E I«H). This is, however, impossible, since ind(u) = -1, and ind(w) = o. An interesting consequence is that 7r( u) provides an example of an invertible element that cannot be written as an exponential, for if 7r( u) = e W for some w in the Calkin algebra, then w = 7r(w') for some w' E B(H), and therefore 7r(u) = e 7r (w') = 7r(e W '). But e W ' is invertible in B(H), which contradicts what we have just shown. Thus, 7r( u) has no logarithm in the Calkin algebra. We shall have more to say about shifts in the next chapter. We shall see further examples and applications concerning compact and Fredholm operators in later chapters. We turn in Chapter 2 to the case where the algebras have involutions and the operators have adjoints. This is the self-adjoint theory, and it is in this setting that some of the deepest results concerning algebras and operators have been proved. 1. Exercises 1. Let (AA).xEA denote a family of Banach algebras. The direct 3um A = ffi,\A,\ is the set of all (a A ) E I1.xAA such that lI(a.x)11 = sup'\lIaAII is finite. Show that this is a Banach algebra under the pointwise-defined operations (a,\) + (b A ) = (a,\ + b,\) Il( a,\) = (Ila,\) (a,\)(b,\) = (a,\b,\), 
1. Exercises 31 and norm given by (a.\)  lI(a.\)II. Show that A is unital or abelian if this is the case for all of the algebras A.\. The re3tricted 3um B = EBo AA is the set of all elements (a A ) E A such that for each £ > 0 there exists a finite subset F of A for which lIaAIl < c if A E A \ F. Show that B is a closed ideal in A. 2. Let A be a Banach algebra and f! a non-empty set. Denote by .eOO(f!, A) the set of all bounded maps f from f! to A. Show that .eOO(f!, A) is a Banach algebra with the pointwise-defined operations and the sup-norm IIfll = sup{lIf(w)1I I w E f!}. If f! is a compact Hausdorff space, show that the set C(f!, A) of all continuous functions from f! to A is a closed subalgebra of .eOO(f!, A). 3. Give an example of a unital non-abelian Banach algebra A in which 0 and A are the only closed ideals. 4. Give an example of a non-modular maximal ideal in an abelian Banach algebra. (If A is the disc algebra, let Ao = {f E A I f(O) = OJ. Then Ao is a closed subalgebra of A and admits an ideal of the type required.) 5. Let A be a unital abelian Banach algebra. ( a) Show that a( a + b) C a( a) + a( b) and a( ab) C a( a )a( b) for all a, b E A. Show that this is not true for all Banach algebras. (b) Show that if A contains an idempotent e (that is, e = e 2 ) other than 0 and 1, then f!(A) is disconnected. ( c) Let al,..., an generate A as a Banach algebra. Show that f!( A) is homeomorphic to a compact subset of c n . More precisely, set O'(al,... ,an) = {(r(al)'... ,r(a n )) IrE f!(A)}. Show that the canon- ical map from f!( A) to a( aI, . . . , an) is a homeomorphism. 6. Let A be a unital Banach algebra. (a) If a is invertible in A, show that a(a- l ) = {A- l I A E a(a)}. (b) For any element a E A, show that r(a n ) = (r(a))n. ( c) If A is abelian, show that the Gelfand representation is isometric if and only if lIa 2 11 = lIal1 2 for all a E A. 7. Let A be a Banach algebra. Show that the spectral radius function r: A  R is upper semi-continuous. (One can show that r is not in general continuous [Hal, Problem 104].) 8. Show that if B is a maximal abelian sub algebra of a unital Banach algebra A, then B is closed and contains the unit. Show that a A(b) = aB(b) for all b E B. 
32 1. Elementary Spectral Theory 9. Let (f!, J.L) be a measure space. Show that the linear span of the idem- potents is dense in LOO(f!, J.L). Show that the spectrum of the Banach alge- bra LOO(O" J.L) is totally disconnected, by showing that if A is an arbitrary abelian Banach algebra in which the idempotents have dense linear span, its spectrum f!(A) is totally disconnected. 10. Let A = C 1 [0, 1], as in Example 1.2.6. Let x: [0, 1] --+ C be the inclusion. Show that x generates A as a Banach algebra. If t E [0,1], show that Tt belongs to f!(A), where Tt is defined by Tt(f) = f(t), and show that the map [0,1] --+ f!(A), t t-+ Tt, is a homeomorphism. Deduce that r(f) = IIflloo (f E A). Show that the Gelfand representation is not surjective for this example. 11. Let A be a unital Banach algebra and set (( a) = inf Ilabli IIbll=l (a E A). We say that an element a of A is a left topological zero divisor if there is a sequence of unit vectors (an) of A such that lim n --+ oo aa n = o. Equivalently, ((a) = o. (a) Show that left topological zero divisors are not invertible. (b) Show that I((a) - (b)1 < Iia - bll for all a, b E A. Hence, ( is a continuous function. ( c) If a is a boundary point of the set Inv( A) in A, show that there is a sequence of invertible elements (v n ) converging to a such that lim n --+ oo IIV n -111- 1 = O. Using the continuity of (, deduce that ((a) = o. Thus, boundary points of Inv( A) are left topological zero divisors. In particular, if ,,\ is a boundary point of the spectrum of an element a of A, then ,,\ - a is a left topological zero divisor. (d) Let f! be a compact Hausdorff space and let A = C(f!). Show that in this case the topological zero divisors are precisely the non-invertible elements (if f is non-invertible, then 0 is a boundary point of the spectrum of f f). ( e) Give an example of a unital Banach algebra and a non-invertible ele- ment that is not a left topological zero divisor. 12. A derivation on an algebra A is a linear map d: A --+ A such that d( ab) = adb + d( a )b. Show that the Leibnitz formula, dn(ab) =  (;)dr(a)dn-r(b) (n = 1, 2, . . .), holds. 
1. Exercises 33 13. Suppose that d is a bounded derivation on a unital Banach algebra A and A E C \ {OJ such that da = Aa. Show that a is nilpotent, that is, that an = 0 for some positive integer n (use the boundedness of o:(d)). 14. Suppose that d is a bounded derivation on a unital Banach algebra A, and that a E A and a:-a = o. Show that da is quasinilpotent. (Hint: Show that d n + 1 (an) = 0 and hence, d n ( an) = n!( da)n.) For a E A, the map b t-+ [a, b] = ab-ba is a bounded derivation on A. Therefore, the Kleinecke- Shirokov theorem holds: If [a, [a, b]] = 0, then [a, b] is quasinilpotent. 15. Let H be a Hilbert space with an orthonormal basis (en)=l' and let u be an operator in B(H) diagonal with respect to (en) with diagonal the sequence (An). Show that u is compact if and only if lim n --+ oo An = O. 16. Let X be a Banach space. If p E B(X) is a compact idempotent, show that its rank is finite. 17. Let u: X -+ Y be a compact operator between Banach spaces. Show that if the range of u is closed, then it is finite-dimensional. (Hint: Show that the well-defined operator X/ker(u) -+ u(X), x + ker(u)...... u(x), is an invertible compact operator.) 18. Let X, Y be Banach spaces and suppose that u E B(X, Y) has compact transpose u*. Show that u is compact using the fact that u** is compact. 19. Let u: X -+ Y and u': X' -+ Y' be bounded operators between Banach spaces. Show that the linear map u EB u': X EB X' -+ Y EB Y', (x, x') ...... (u(x), u'(x')), is bounded with norm max{llull, Ilu'II}. Show that if u and u' are Fredholm operators, so is u EB u', and ind( u EI1 u') = ind( u) + ind( u'). 20. If X is an infinite-dimensional Banach space and u E B(X), show that n 0' (u + v) = 0"( u) \ {A Eel u - A is Fredholm of index zero}. vEK(X) 
34 1. Elementary Spectral Theory 1. Addenda Let G be a locally compact abelian group. If J.L is Haar measure on G, we write L1(G) for L1(G,J.L). If I,g E L1(G), then there is an element 1 * 9 E £l(G) such that (f*g)(x) = J f(x-y)g(y)dp,(y) for almost all x in G. The product 1 * 9 is the convolution of f and g. Under the multiplication operation given by (f, g)  f * g, £1 (G) is an abelian Banach algebra, called the group algebra of G. It has a unit if and only if G is discrete. Let G be the dual group of G, that is, the set of continuous homo- morphisms, from G to T. This is endowed with a suitable topology making it a locally compact group. For fELl (G) and, E G, define J('Y) = J f(x h(x) dp,(x). Then the function 1 A T--y: L (G) -+ C, f  f(,), is a character on L 1 (G), and all characters on L 1 (G) are of this form. The map G-+f2(L 1 (G)), ,T--y, is a homeomorphism. Reference: [Cnw 2]. Suppose u is a non-zero compact operator on an infinite-dimensional Banach space X. Then there is a non-trivial closed vector subspace Y of X such that v(Y) C Y for all operators v E B(X) commuting with u. This is a special case of a theorem of Lomonosov [TL, Theorem 7.15]. Let X be an infinite-dimensional Banach space. An operator u E B(X) is a Rie3zoperator if its essential spectrum is the zeroset, O'e(u) = {OJ. The spectral theory of these operators is similar to that of compact operators. Obviously, the sum of a quasinilpotent operator and a compact operator is a Riesz operator. The converse is true for Hilbert spaces and is also known for some other Banach spaces. For certain Banach algebras and certain of their closed ideals, one can develop a Fredholm theory that is analogous to the classical Fredholm theory relative to B(X) and ]{(X) for Banach spaces X. References: [BMSW], [Wes]. 
CHAPTER 2 C*-Algebras and Hilbert Space Operators In this chapter we commence our study of C*-algebras and of opera- tors on Hilbert spaces. Hilbert spaces are very well-behaved compared with general Banach spaces, and the same is even more true of C*-algebras as compared with general Banach algebras. The main results of this chapter are a theorem of Gelfand, which asserts that (up to isomorphism) all abelian C*-algebras are of the form Co(n), where n is a locally compact Hausdorff space, and the spectral theorem. This theorem enables us to "synthesize" a normal operator from linear combinations of projections where the coef- ficients lie in the spect rum. It is a very powerlul result. 2.1. C*-Algebras We begin by defining a number of concepts that make sense in any algebra with an involution. An involution on an algebra A is a conjugate-linear map a  a* on A, such that a** = a and (ab)* = b*a* for all a,b E A. The pair (A,*) is called an involutive algebra, or a *-algebra. If S is a subset of A, we set S* = {a* I a E S}, and if S* = S we say S is 3 elf-adjoint. A self-adjoint subalgebra B of A is a *-3ubalgebra of A and is a *-algebra when endowed with the involution got by restriction. Because the intersection of a family of *-subalgebras of A is itself one, there is for every subset S of A a smallest *-algebra B of A containing S, called the *-algebra generated by S. If I is self-adjoint ideal of A, then the quotient algebra AI I is a *-algebra with the involution given by (a + 1)* = a* + I (a E A). We define an involution on A extending that of A by setting (a, ,,\)* = (a* , .x). Thus, A is a *-algebra, and A is a self-adjoint ideal in A. 35 
36 2. C*-Algebras and Hilbert Space Operators An element a in A is self-adjoint or hermitian if a = a*. For each a E A there exist unique hermitian elements b, c E A such that a = b + ic (b = t (a + a*) and c = ii (a - a*)). The elements a* a and aa* are hermitian. The set of hermitian elements of A is denoted by A..a. We say a is normal if a* a = aa*. In this case the *-algebra it generates is abelian and is in fact the linear span of all am a*n, where m, n E Nand n + m > o. An element p is a projection if p = p* = p2 . If A is unital, then 1* = 1 (1* = (11*)* 1). If a E Inv(A), then (a*)-I = (a- I )*. Hence, for any a E A, a(a*) = a(a)* = { E C I A E a(a)}. An element u in A is a unitary if u*u = uu* = 1. If u*u = 1, then u is an isometry, and if uu* = 1, then u is a co-isometry. If c.p: A -+ B is a homomorphism of *-algebras A and Band c.p preserves adjoints, that is, c.p(a*) = (c.p(a))* (a E A), then c.p is a *-homomorphism. If in addition c.p is a bijection, it is a *-isomorphism. If c.p: A -+ B is a *- homomorphism, then ker( c.p) is a self-adjoint ideal in A and c.p( A) is a *-subalgebra of B. An automorphism of a *-algebra A is a *-isomorphism c.p: A -+ A. If A is unital and u is a unitary in A, then Ad u: A -+ A, * a  uau , is an automorphism of A. Such automorphisms are called inner. We say elements a, b of A are unitarily equivalent if there exists a unitary u of A such that b = uau*. Since the unitaries form a group, this is an equivalence relation on A. Note that a(a) = a(b) if a and b are unitarily equivalent. A Banach *-algebra is a *-algebra A together with a complete submul- tiplicative norm such that Ila* II = lIall (a E A). If, in addition, A has a unit such that 11111 = 1, we call A a unital Banach *-algebra. A C*-algebra is a Banach *-algebra such that Ila*all = lI a ll 2 (a E A). (1) A closed *-subalgebra of a C*-algebra is obviously also a C*-algebra. We shall therefore call a closed *-subalgebra of a C*-algebra a C*-subalgebra. If a C*-algebra has a unit 1, then automatically 11111 = 1, because 11111 = 111*111 = 11 1 11 2 . Similarly, if p is a non-zero projection, then Ilpll = 1. If u is a unitary of A, then Ilull = 1, since IIull 2 = lIu*ull = 11111 = 1. Hence, a(u) C T, for if A E a(u), then A-I E a(u- I ) = a(u*), so IAI and lA-II < 1; that is, IAI = 1. The seemingly mild requirement on a C*-algebra in Eq. (1) is in fact very strong-far more is known about the nature and structure of these 
2.1. C*-Algebras 37 algebras than perhaps of any other non-trivial class of algebras. Because of the existence of the involution, C*-algebra theory can be thought of as "infinite-dimensional real analysis." For instance, the study of linear func- tionals on C*-algebras (and of traces, cf. Section 6.2) is "non-commutative measure theory." 2.1.1. Ezample. The scalar field C is a unital C*-algebra with involution given by complex conjugation A ....-..+ A. 2.1.2. Ezample. If 0 is a locally ompact Hausdorff space, then Co(O) is a C* -algebra with involution f ....-..+ f. Similarly, all of the following algebras are C*-algebras with involution given by f  f: (a) £00(5) where 5 is a set; (b) Loo(n, J.L) where (n, 1") is a measure space; (c) Cb(O) where 0 is a topological space; (d) Boo(O) where 0 is a measurable space. 2.1.3. Ezample. If H is a Hilbert space, then B(H) is a C*-algebra. We shall see that every C*-algebra can be thought of as a C*-subalgebra of some B(H) (Gelfand-Naimark theorem). We defer to Section 2.3 a fuller consideration of this example. 2.1.4. Ezample. If (AA)AEA is a family of C*-algebras, then the direct sum EBAA.x is a C*-algebra with the pointwise-defined involution, and the restricted sum EBo A.x is a closed self-adjoint ideal (cf. Exercise 1.1). 2.1.5. Ezample. If 0 is a non-empty set and A is a C*-algebra, then £00(0, A) is a C*-algebra with the pointwise-defined involution. This of course generalises Example 2.1.2 ( a). If 0 is a locally compact Hausdorff space, we say a continuous function f: 0  A vanishes at infinity if, for each £ > 0, the set {w E 0 Illf(w)11 > £} is compact. Denote by Co(n, A) the set of all such functions. This is a C*-subalgebra of £00(0, A). The following easy result has a surprising and important corollary: 2.1.1. Theorem. If a is a self-adjoint element of a C*-algebra A, then r(a) = lIali. Proof. Clearly, lIa 2 11 = lIall 2 , and therefore by induction lIa 2n II = lI a ll 2n , so r(a) = lim n --+ oo lIa n ll 1 / n = lim n --+ oo lIa 2n IIl/2n = lIali. 0 2.1.2. Corollary. There is at most one norm on a *-algebra making it a C*-algebra. 
38 2. C*-Algebras and Hilbert Space Operators Proof. If 11.111 and 11.112 are norms on a *-algebra A making it a C*-algebra, then lIall; = Ila*allj = r(a*a) = sup IAI AEO'(a* a) so II a I It = II a 112 . (j = 1, 2), o 2.1.3. Lemma. Let A be a Banach algebra endowed with an involution such that IIal1 2 < Ila*all (a E A). Then A is a C*-algebra. Proof. The inequalities II a l1 2 < Ila* all < Ila* II II all imply that lIall < Ila* II for all a. Hence, lIall = Ila*lI, and therefore lIall 2 = lIa*all. 0 We associate to each C*-algebra A a certain unital C*-algebra M(A) which contains A as an ideal. This algebra is of great importance in more advanced aspects of the theory, especially in certain approaches to K-theory. A double centraliser for a C*-algebra A is a pair (L, R) of bounded linear maps on A, such that for all a, b E A L(ab) = L(a)b, R(ab) = aR(b) and R(a)b = aL(b). For example, if c E A and Le, Re are the linear maps on A defined by Le(a) = ea and Re(a) = ae, then (Le, Re) is a double centraliser on A. It is easily checked that for all e E A lIeli = sup Ilebll = sup Ilbell, IIblll IIblll and therefore IILell = liRe II = lIeli. 2.1.4. Lemma. If (L, R) is a double centraliser on a C*-algebra A, then IILII = IIRII. Proof. Since IlaL(b)11 = IIR(a)bll < IIRlillallllbll, we have IIL(b)11 = sup IlaL(b)11 < IIRlillbll, lIalll and therefore II L II < II RII. Also, II R( a )bll = II aLe b) II < II Lilli a 1111 bll implies IIR(a)11 = sup IIR(a)bll < IILllllall, IIblll and therefore IIRII < IILII. Thus, IILII = IIRII. o If A is a C* -algebra, we denote the set of its double centralisers by M(A). We define the norm of the double centraliser (L, R) to be IILII = IIRII. It is easy to check M(A) is a closed vector subs pace of B(A) EB B(A). 
2.1. C*-Algebras 39 If (L1' R 1 ) and (L 2 , R 2 ) E M( A), we define their product to be (L 1 ,R 1 )(L 2 ,R 2 ) = (L 1 L 2 ,R 2 R 1 ). Straightforward computations show that this product is again a double centraliser of A and that M(A) is an algebra under this multiplication. If L:A  A, define L*:A  A by setting L*(a) = (L(a*))*. Then L* is linear and the map L  L* is an isometric conjugate-linear map from B(A) to itself such that L** = Land (L 1 L 2 )* = LiLi. If (L, R) is a double centraliser on A, so is (L, R)* = (R*, L*). It is easily verified that the map (L, R)  (L, R)* is an involution on M(A). 2.1.5. Theorem. If A is a C*-algebra, then M(A) is a C*-algebra under the multiplication, involution, and norm defined above. Proof. The only thing that is not completely straightforward that has to be checked is that if T = (L, R) is a double centraliser, then IIT*TII IITII2. If II all < 1, then IIL(a)1I2 = II(L(a))* L(a)11 = IIL*(a*)L(a)11 Ila* R* L(a)11 < IIR* LII = IIT*TII, so IITII2 = sup IIL(a)1I2 < IIT*TII < IITII2, lIalll and therefore IIT*TII = IIT112. o The algebra M(A) is called multiplier algebra of A. The map A  M(A), a  (La, Ra), is an isometric *-homomorphism, and therefore we can, and do, identify A as a C*-subalgebra of M(A). In fact A is an ideal of M(A). Note that M(A) is unital (the double centraliser (idA,id A ) is the unit), so A = M(A) if and only if A is unital. We have already seen in Chapter 1 that for every Banach algebra A, its unitisation A is a Banach algebra with the norm II( a, "\)11 = lIall + 1"\1. If A is a Banach *-algebra, then so is A with this norm. However, if A is a C*-algebra, there is a problem here, since this norm does not make A a C*-algebra in general. For instance, if A = C and (a,"\) = (-2, 1), we have lI(a, "\)11 2 = 9, but lI(a, "\)*(a, "\)11 = 1. We can, however, endow A with a norm making it a C*-algebra: 2.1.6. Theorem. If A i a C*-algebra, then there is a (necessarily unique) norm on its unitisation A making it into a C*-algebra, and extending the norm of A. 
40 2. C*-Algebras and Hilbert Space Operators Proof. Uniqueness of the norm is given by Corollary 2.1.2. The proof of existence falls into two cases, depending on whether A is unital or non- uni tal. Suppose first that A has a unit e. Then the map <p from A to the direct sum of the C* -algebras A and C defined by <p( a, ,\) = (a + '\e, ,\) is a *-isomorphism. Hence, one gets a norm on A making it a C*-algebra by setting lI(a, '\)11 = 1I<p(a, '\)11. Now suppose A has no unit. If 1 is the unit of M(A), then AnC1 = O. The map <p from A onto the C*-subalgebra A EB C1 of M(A) defined by setting <p( a, ,\) = a + ,\ 1 is a *-isomorphism, so we get a norm on A making it a C*-algebra by setting lI(a, '\)11 = 1I<p(a, '\)11. 0 If A is a C*-algebra, we shall always understand the norm of A to be the one making it a C*-algebra. Note that when A is non-unital, M(A) is in general very much bigger than A. For instance, it is shown in Section 3.1 that if A = C o (f2), where f2 is a locally compact Hausdorff space, then M(A) = Cb(f2). If <p: A  B is a *-homomorphism between *-algebras A and B, then it extends uniquely to a unital *-homomorphism cj;: A  B. 2.1.7. Theorem. A *-homomorphism <p: A  B from a Banach *-algebra A to a C*-algebra B is necessarily norm-decreasing. Proof. We may suppose that A, B, and <p are unital (by going to A, B, and cj; if necessary). If a E A, then a(<pa) C a(a), so lI<pa11 2 = 11<p(a)*<p(a)1I = II <p ( a * a) II = r ( <p ( a * a )) < r ( a * a) < II a * a II < II a 11 2 . Hence, II <p ( a ) II < II a II. 0 2.1.8. Theorem. If a is a hermitian element of a C*-algebra A, then a( a) C R. Proof. We may suppose that A is unital. Since e ia is unitary, a( e ia ) C T. If,\ E a(a) and b = 2::'=1 in(a_,\)n-1 In! then eia_e iA = (e i (a-A)_l)e i '\ = (a - ,\ )be iA . Since b commutes with a, and since a - ,\ is non-invertible, e ia - e iA is non-invertible. Hence, e iA E T, and therefore ,\ E R. Thus, a(a) C R. 0 2.1.9. Theorem. If r is a character on a C*-algebra A, then it preserves adjoints. Proof. If a E A, then a = b + ic where b, c are hermitian elements of A. The numbers r(b) and Tee) are real because they are in a(b) and a(e) respectively, so r(a*) = r(b - ic) = r(b) - ir(e) = (r(b) + ir(e))-= r(a)-.O The character space of a unital abelian Banach algebra is non-empty, so this is true in particular for unital abelian C* -algebras. However, there are non-unital, non-zero, abelian Banach algebras for which the character 
2.1. C*-Algebras 41 space is empty. Fortunately, this cannot happen in the case of C*-algebras. Let A be a non-unital, non-zero, abelian C*-algebra. Then A contains a non-zero hermitian element, a say. Since r(a) = lIall by Theorem 2.1.1, it follows that there is a character r on A such that Ir(a)1 = Iiall f= o. Hence, the restriction of r to A is a non-zero homomorphism from A to C, that is, a character on A. We shall now completely determine the abelian C*-algebras. This re- sult can be thought of as a preliminary form of the spectral theorem. It allows us to construct the functional calculus, a very useful tool in the analysis of non-abelian C*-algebras. 2.1.10. Theorem (Gelfand). If A is a non-zero abelian C *-alge bra, then the Gelfand representation c.p: A  Co(n(A)), " a  a, is an isometric *-isomorphism. Proof. That c.p is a norm-decreasing homomorphism, such that 11c.p(a)1I = r( a), is given by Theorem 1.3.6. If r E n( A), then <pC a*)( r) = r( a*) = r( a)- = c.p( a )*( r), so c.p is a *-homomorphism. Moreover, <p is isomet- ric, since 1Ic.p(a)1I2 = 1I<p(a)*c.p(a)11 = 1I<p(a*a)1I = r(a*a) = Ila*all = Il a 11 2 . Clearly, then, <peA) is a closed *-subalgebra of C o (!1) separating the points of n(A), and having the property that for any r E !1(A) there is an element a E A such that c.p(a)(r) f= O. The Stone-Weierstrass theorem implies, therefore, that <p(A) = C o (f2(A)). 0 Let S be a subset of a C*-algebra A. The C*-algebra generated by S is the smallest C*-subalgebra of A containing S. If S = {a}, we denote by C*(a) the C*-subalgebra generated by S. If a is a normal, then C*(a) is abelian. Similarly, if A is unital and a normal, then the C*-subalgebra generated by 1 and a is abelian. Observe that r( a) = lIall if a is a normal element of a C*-algebra (apply Theorem 2.1.10 to C*(a)). The following result is important. 2.1.11. Theorem. Let B be a C*-subalgebra of a unital C*-algebra A containing the unit of A. Then O'B(b) = O'A(b) (b E B). Proof. First suppose that b is a hermitian element of B. Since in this case 0' A(b) is contained in R, it has no holes, and therefore by Theorem 1.2.8, O'A(b) = O'B(b). Therefore, b is invertible in B if and only if it is invertible in A. 
42 2. C*-Algebras and Hilbert Space Operators Now suppose that b is an arbitrary element of B, that is invertible in A, so there is an element a E A such that ba = ab = 1. Then a*b* = b*a* = 1, so bb*a*a = 1 => bb* is invertible in A and therefore in B. Hence, there is an element c E B such that bb*c = 1. Consequently, b*c = a, so a E B, which implies that b is invertible in B. Thus, for any element of B, its invertibility in A is equivalent to its invertibility in B. The theorem follows. 0 If A is a unital C*-algebra and a E Asa, then e ia is a unitary, but not all unitaries are of this form-it will be seen later that the Calkin algebra on a Hilbert space is a C*-algebra and the image of the unilateral shift in this algebra provides an example of a unitary that has no logarithm (cf. Ex- ample 1.4.4). Using Theorem 2.1.10, we can give some useful conditions that ensure a unitary doe3 have a logarithm. 2.1.12. Theorem. Let u be a unitary in a unital C*-algebra A. If 0'( u) =I T, then there exists a E Asa such that u = e ia . (If 111 - ull < 2, then O'(u) =I T.) Proof. By replacing u by Au for some A E T if necessary, we may sup- pose that -1 ft O'(u). Since u is normal, we may also suppose that A is abelian (replacing A by the C*-subalgebra generated by 1 and u if need be). Let cp: A  C(n) be the Gelfand representation, let f = cp(u), and as usual denote by In: C \ (-00, 0]  C the principal branch of the logarithm function. Then 9 = In 0 f is a well-defined element of Co(n), and e 9 = f. Since If(w)1 = 1 for all w E f2, the real part of 9 vanishes, so 9 = ih where h = Ii E Co(n). Let a = cp-l(h). Then a E Asa and u = e ia because cp( u) = e ih = eC;?( ia) = cp( e ia ). The parenthetical observation in the statement of the theorem follows from the equations 111 - ull = r(l - u) = sup{11 - All A E O'(u)}, which imply that -1 ft 0'( u) when 111 - u" < 2. o We are now going to set up the functional calculus, for which we need to make two easy observations: If (): n  f2' is a continuous map between compact Hausdorff spaces n and n', then the tran3po3e map ()t: C(n')  C(f2), f  fB, is a unital *-homomorphism. Moreover, if () is a homeomorphism, then B t is a *-isomorphism. Our second observation is that a *-isomorphism of C*-algebras is nec- essarily isometric. This is an immediate consequence of Theorem 2.1.7. 
2.1. C*-Algebras 43 2.1.13. Theorem. Let a be a normal element of a unital C*-algebra A, and suppose that z is the inclusion map of 0'( a) in C. Then there is a unique unital *-homomorphism c.p: C(O'(a))  A such that c.p(z) = a. Moreover, c.p is isometric and im( c.p) is the C*-subalgebra of A generated by 1 and a. Proof. Denote by B the (abelian) C*-algebra generated by 1 and a, and let 1/J: B  C(f2(B)) be the Gelfand representation. Then 1/J is a *-isomorphism by Theorem 2.1.10, and so is at: C(a(a))  C(f2(B)), since a:f2(B) --+ a(a) is a homeomorphism. Let c.p:C(a(a))  A be the com- position 1/J-l at, so c.p is a *-homomorphism. Then c.p( z) = a, since c.p( z) = 1/J- 1 (a t (z)) = 1/J- 1 (a) = a, and obviously c.p is unital. From the Stone- Weierstrass theorem, we know that C(a(a)) is generated by 1 and z; c.p is therefore the unique unital *-homomorphism from C(a(a)) to A such that c.p(z) = 1. It is clear that c.p is isometric and im( c.p) = B. 0 As in Theorem 2.1.13, let a be a normal element of a unital C*-algebra A, and let z be the inclusion map of C(a(a)) in C. We call the unique unital *-homomorphism c.p: C(a(a))  A such that c.p(z) = a the functional calculu3 at a. If p is a polynomial, then c.p(p) = p( a), so for f E C( a( a)) we may write f( a) for c.p( a). Note that f( a) is normal. Let B be the image of c.p, so B is the C*-algebra generated by 1 and a. If r E f2(B), then f(r(a)) = r(f(a)), since the maps f  f(r(a)) and f  r(f(a)) from C(a(a)) to Care *-homomorphisms agreeing on the generators 1 and z and hence are equal. 2.1.14. Theorem (Spectral Mapping). Let a be a normal element of a unital C*-algebra A, and let f E C(O'(a)). Then a(f(a)) = f(a(a)). Moreover, if 9 E C( a(f( a))), then (g 0 f)(a) = g(f(a)). Proof. Let B be the C* -subalgebra generated by 1 and a. Then a(f( a)) = {r(f(a)) IrE f2(B)} = {f(r(a)) IrE f2(B)} = f(a(a)). If C denotes the C*-subalgebra generated by 1 and f(a), then C C B and for any r E f2(B) its restriction rc is a character on C. We therefore have r((g 0 f)(a)) = g(f(r(a))) = g(rc(f(a))) = rc(g(f(a))) = r(g(f(a))). Hence, (g 0 f)(a) = g(f(a)). 0 We close this section by showing that if f2 is a compact Hausdorff space, then the character space of C(f2) is f2. 
44 2. C*-Algebras and Hilbert Space Operators 2.1.15. Theorem. Let n be a compact Hausdorff space, and for each w E 0 let 6w be the character on C(O) given by evalution at w; that is, 6w(f) = few). Then the map o -+ O(C(f2)), w  6w, is a homeomorphism. Proof. This map is continuous because if (W.x).xEA is a net in 0 converging to a point w, then lim.xEA f(w A ) = f(w) for all f E C(O), so the net (6 w .\) is weak* convergent to 6 w . The map is also injective, because if w,w' are distinct points of f2, then by Urysohn's lemma there is a function f E 0(0) such that f(w) = 0 and f(w') = 1, and therefore 6w -:F 6 W '. Now we show surjectivity of the map. Let r E f2(C(O)). Then M = ker(r) is a proper C*-algebra of C(f2). Also, M separates the points of f2, for if w, w' are distinct points of Q, then as we have just seen there is a function f E C(f2) such that f(w) -:F few'), so 9 = f - r(f) is a function in M such that g(w) -:F g(w'). It follows from the Stone-Weierstrass theorem that there is a point w E f2 such that f(w) = 0 for all f E M. Hence, (f - r(f))(w) = 0, so f(w) = r(f), for all f E C(f2). Therefore, r = 6 w . Thus, the map is a continuous bijection between compact Hausdorff spaces and therefore is a homeomorphism. 0 2.2. Positive Elements of C*-Algebras In this section we introduce a partial order relation on the hermitian elements of a C*-algebra. The principal results are the existence of a unique positive square root for each positive element and Theorem 2.2.4, which asserts that elements of the form a* a are positive. 2.2.1. Remark. Let A = C o (f2), where f2 is a locally compact Hausdorff space. Then Asa is the set of real-Valued functions in A and there is a natural partial order on Asa given by f < 9 if and only if f( w) < g( w) for all w E f2. An element f E A is positive, that is, f > 0, if and only if f is of the form f = 99 for some 9 E A, and in this case f has a unique positive square root in A, namely the function w t-+ J f( w ) . Note that if f = J we can also express the positivity condition in terms of the norm: If t E R, then f is positive if Ilf - tll < t, and in the reverse direction if IIfll < t and f > 0, then IIf - tll < t. We shall presently define a partial order on an arbitrary C*-algebra that generalises that of C o (f2), and we shall obtain similar, and many other, results. Let A be a unital algebra and B a subalgebra such that B + C1 = A. Then O'B(b) U {O} = 0' A(b) U {O} for all b E B. If B is non-unital, this is seen by observing that the map iJ -+ A, (b, A)  b + AI, is an isomorphism. 
2.2. Positive Elements of C*-Algebras 45 If B has a unit e not equal to the unit 1 of A, then for any b E B and ,,\ E C \ {O} invertibility of b + ,,\ in A is equivalent to invertibility of b + "\e in B, so 0' A ( b) = 0' B (b) U {O}. From these observations and Theorem 2.1.11, it is clear that for any C*-subalgebra B of a C*-algebra A we have lYB(b) U {OJ = O'A(b) U {OJ for all b E B. An element a of a C*-algebra A is positive if a is hermitian and 0'( a) C R +. We write a > 0 to mean that a is positive, and denote by A + the set of positive elements of A. By the preceding observation B+ = B n A+ for any C*-subalgebra B of A. If 5 is a non...empty set, then an element f E £00(5) is positive in the C*-algebra sense if and only if f( x) > 0 for all xES, because a(f) is the closure of the range of f. Hence, if f! is any locally compact Hausdorff space, then f E Co(f!) is positive if and only if f(w) > 0 for all w E f!. If a is a hermitian element of a C*-algebra A observe that C*( a) is the closure of the set of polynomials in a with zero constant term. 2.2.1. Theorem. Let A be a C*-algebra and a E A+. Then there exists a unique element b E A+ such that b 2 = a. Proof. That there exists b E C* ( a) such that b > 0 and b 2 = a follows from the Gelfand representation, since we may use it to identify C*(a) with Co(n), where n is the character space of C*(a), and then apply Re- mark 2.2.1. Suppose that e is another element of A + such that e 2 = a. As e commutes with a it must commute with b, since b is the limit of a sequence of polynomials in a. Let B be the (necessarily abelian) C*-subalgebra of A generated by band c, and let c.p: B  Co(f!) be the Gelfand representation of B. Then c.p(b) and c.p(e) are positive square roots of c.p(a) in Co(f!), so by another application of Remark 2.2.1, c.p(b) = c.p(e), and therefore b = e. 0 If A is a C*-algebra and a is a positive element, we denote by a 1 / 2 the unique positive element b such that b 2 = a. If e is a hermitian element, then e 2 is positive, and we set lei = (e 2 )1/2, e+ = !(Iel + e), and e- = !(Iel- e). Using the Gelfand representation of C*(e), it is easy to check that lel,e+ and e- are positive elements of A such that e = c+ - e- and e+e- = o. 2.2.2. Remark. If a is a hermitian element of the closed unit ball of a unital C*-algebra A, then 1 - a 2 E A+ and the elements u = a + i v 1 - a 2 and v = a - i V 1 - a 2 are unitaries such that a = !( u + v). Therefore, the unitaries linearly span A, a result that is frequently useful. 
46 2. C*-Algebras and Hilbert Space Operators 2.2.2. Lemma. Suppose that A is a unital C *-alge bra, a is a hermitian element of A and t E R. Then, a > 0 if lIa- tll < t. In the reverse direction, if lIall < t and a > 0, then lIa - tll < t. Proof. We may suppose that A is the (abelian) C*-subalgebra generated by 1 and a, so by the Gelfand representation A = C( a( a)). The result now follows from Remark 2.1.1. 0 It is immediate from Lemma 2.2.2 that A + is closed in A. 2.2.3. Lemma. The sum of two positive elements in a C*-algebra is a positive element. Proof. Let A be a C*-algebra and a, b positive elements. To show that a+b > 0 we may suppose that A is unital. By Lemma 2.2.2, Ila-llalill < lIall and IIb-lIblill < IIbll, so lIa+b-llall-llbllll < Ila-lIallll+llb-lIbllli < lIall+llbli. By Lemma 2.2.2 again, a + b > o. 0 2.2.4. Theorem. If a is an arbitrary element of a C*-algebra A, then a*a is posi ti ve. Proof. First we show that a = 0 if -a* a E A +. Since a( -aa*) \ {O} = a( -a*a) \ {OJ by Remark 1.2.1, -aa* E A+ because -a*a E A+. Write a = b + ie, where b, e E ABa. Then a*a + aa* = 2b 2 + 2e 2 , so a*a = 2b 2 + 2c 2 - aa* E A+. Hence, O'(a*a) = R+ n (-R+) = {O}, and therefore lIall 2 = lIa*all = r(a*a) = o. Now suppose a is an arbitrary element of A, and we shall show that a * a is posi ti ve. If b = a * a, then b is hermitian, and therefore we can write b = b+ - b-. If e = ab-, then -e*e = -b-a*ab- = -b-(b+ - b-)b- = (b-)3 E A +, so e = 0 by the first part of this proof. Hence, b- = 0, so a*a=b+EA+. 0 If A is a C*-algebra, we make ABa a poset by defining a < b to mean b - a E A +. The relation < is translation-invariant; that is, a < b => a + e < b + e for all a, b, e E ABa. Also, a < b => ta < tb for all t E R +, and a < b <=> -a > -b. Using Theorem 2.2.4 we can extend our definition of lal: for arbitrary a set I a I = (a * a) 1/2 . We summarise some elementary facts about A+ in the following result. 2.2.5. Theorem. Let A be a C*-algebra. (1) The set A+ is equal to {a*a I a E A}. (2) If a, b E ABa and e E A, then a < b => e*ae < e*be. (3) If 0 < a < b, then lIali < IIbli. (4) If A is unital and a, b are positive invertible elements, then a < b => o < b- I < a-I. 
2.2. Positive Elements of C*-Algebras 47 Proof. Conditions (1) and (2) are implied by Theorem 2.2.4 and the exist- ence of positive square roots for positive elements. To prove Condition (3) we may suppose that A is unital. The inequality b < IIbll is given by the Gelfand representation applied to the C*-algebra generated by 1 and b. Hence, a < II bll. Applying the Gelfand representation again, this time to the C*-algebra generated by 1 and a, we obtain the inequality lIall < IIbll. To prove Condition (4) we first observe that if c > 1, then c is invertible and c- I < 1. This is given by the Gelfand representation applied to the C*-subalgebra generated by 1 and c. Now a < b => 1 = a- I / 2 aa- I / 2 < a- I / 2 ba- I / 2 => (a- I / 2 ba- I / 2 )-1 < 1, that is, al/2b-Ial/2 < 1. Hence, b- I < (a l / 2 )-I(a l / 2 )-1 = a-I. 0 2.2.6. Theorem. If a, b are positive elements of a C*-algebra A, then the inequality a < b implies the inequality a l / 2 < b l / 2 . Proof. We show a 2 < b 2 => a < b and this will prove the theorem. We may suppose that A is unital. Let t > 0 and let c, d be the real and imaginary hermitian parts of the element (t + b + a)( t + b - a). Then c = t((t + b + a)(t + b - a)) + (t + b - a)(t + b + a)) = t 2 + 2tb + b 2 - a 2 > t 2 . Consequently, c is both invertible and positive. Since 1 + ic- I / 2 dc- I / 2 = c- I / 2 ( c + id)c- I / 2 is invertible, therefore c + id is invertible. It follows that t + b - a is left invertible, and therefore invertible, because it is hermitian. Consequently, -t ft O'(b - a). Hence, O'(b - a) C R+, so b - a is positive, that is, a < b. 0 It is not true that 0 < a < b => a 2 < b 2 in arbitrary C*-algebras. For example, take A = M 2 (C). This is a C*-algebra where the involution is given by ( a (3 ) * ( a 1 ) , 6 = P "$ . Let p and q be the projections p = (  ) and q = t ( ). Then p < p + q, but p2 = p 1:. (p + q)2 = P + q + pq + qp, since the matrix q + pq + qp = t (  ) has a negative eigenvalue. It can be shown that the implication 0 < a < b => a 2 < b 2 holds only in abelian C*-algebras [Ped, Proposition 1.3.9]. 
48 2. C*-Algebras and Hilbert Space Operators 2.3. Operators and Sesquilinear Forms In this section (and the next) we shall interpret and apply many of the ideas of Chapter 1 and the first two sections of this chapter in the context of operators on Hilbert spaces. We shall also prove the invaluable polar decomposition theorem. An important concern in the present section is the correspondence of operators and sesquilinear forms. This is interesting in its own right, but it also has wide applicability-for example, we shall use it in the proof of the spectral theorem. We begin by showing that operators on Hilbert spaces have adjoints. 2.3.1. Theorem. Let HI and H 2 be Hilbert spaces. (1) lEu E B(HI,H2)' then there is a unique element u* E B(H 2 ,H I ) such that (U(XI),X2) = (XI,U*(X2)) (Xl E HI, X2 E H 2 ). (2) The map u r-+ u* is conjugate-linear and u** = u. Also lIull = lIu*1I = lIu*ull l / 2 . Proof If u E B(HI' H 2 ) and X2 E H 2 , then the function HI -+ C, xl  (U(XI)' X2), is continuous and linear, so by the Riesz representation theorem for linear functionals on Hilbert spaces there is a unique element U*(X2) E HI such that (U(XI),X2) = (XI,U*(X2)) (Xl E HI). Moreover, IIU*(X2)1I = sup I(U(XI),X2)1 < lIu1l1lX211. IIxdl l The map u*: H 2 -+ HI, x2  U*(X2), is linear and lIu*1I < lIuli. Thus, u* satisfies Condition (1) (uniqueness of u is obvious). If Xl E HI and IIXIII < 1, then (u( Xl), u( Xl)) = (Xl, u*u( Xl)) < lIu*ulI, so lIull 2 = sup IIU(XI)1I2 < lIu*ull < lIull 2 . Uxtlll Hence, lIuli = lIu*ull l / 2 . The other assertions in Condition (2) of the theorem have routine verifications. 0 If u: HI -+ H 2 is a continuous linear map between Hilbert spaces, we call the map u*:H 2 --+ HI the adjointofu. Note that ker(u*) = (im(u))l., where im(u) is the range of u, and hence, (im(u*))- = ker(u)l.. If HI  H 2  H3 are continuous linear maps between Hilbert spaces, then (vu)* = u*v*. 
2.3. Operators and Sesquilinear Forms 49 If H is a Hilbert space, then B(H) is a C*-algebra under the involution u t-+ u*, where u* is the adjoint of u. It follows in particular that Mn(C) = B(cn) is a C*-algebra. Observe that the involution on Mn(C) is given by (Aij)ij = ('xji)ij. If H is a vector space, a map 0': H 2 -+ C is a sesquilinear form if it is linear in the first variable and conjugate-linear in the second. For such a form the polarisation identity 3 O'(x, y) = i L ikO'(x + iky, x + iky) k=O holds. Thus, sesquilinear forms 0' and 0" on H are equal if and only if 0'( x, x) = 0" (x, x) for all x E H. Sesquilinear forms are taken up in more detail later in this section. If H is a Hilbert space and u E B(H), then (x, y) t-+ (u(x), y) is a sesquilinear form on H. Hence, if u, v E B (H), then u = v if and only if (u(x), x) = (v(x), x) for all x E H. If u*u = id and uu* = id, we say u is a unitary operator. This is equivalent to u being isometric and surjective. Observe that u is isometric {:} u*u = ide 2.3.1. Ezample. Let (en)=l be an orthonormal basis for a Hilbert space H, and suppose that u is an operator diagonal with respect to (en), with diagonal sequence (A n ). Then u * is also diagonal with respect to (en) and its diagonal sequence is ('xn). This follows from the observation that (u*(en),e m ) = (en,u(e m )} = (en, Am em) = .xmb nm , where b nm is the Kro- necker delta symbol, which implies that u*( en) = 'xnen. Since all operators diagonal with respect to the same basis commute, uu* = u*u; that is, u is normal. 2.3.2. Ezample. Let (en) and H be as in the preceding example, but this time let u denote the unilateral shift on this basis, so u( en) = e n +1 for all n > 1. The adjoint u* is the backward shift: u*( en) = e n -1 if n > 1 and u*( e1) = o. It follows that u*u = 1. It is easily seen that u has no eigenvalues. In contrast, u* has very many, for if IAI < 1, then A is an eigenvalue: Set x = 2::=1 Ane n and observe that x E H because 2:=1 IA/ 2n < 00, and that x :F 0 and u*(x) = Ax. It follows from this, and the fact that IIu* II = I/ull = 1, that 0'( u) = 0'( u*) = D. Incidentally, if (fn)=l is an orthononnal basis for another Hilbert space K and v is the unilateral shift on (fn), so v(fn) = fn+1, then v = wuw*, where w: H -+ K is the unitary operator such that w( en) = fn for all n > 1. From the abstract point of view, the operators u and v are therefore the same, so one can speak of "the" unilateral shift. 
50 2. C*-Algebras and Hilbert Space Operators If K is a closed vector subspace of a Hilbert space H, we call the projection p of H on K along K.L the (orthogonal) projection on K. This is self-adjoint. If u E B(H), then 1< is invariant for u (that is, u(K) C K) if and only if pup = up. We say that K is reducing for u if both K and K.L are invariant for u. This is equivalent to p commuting with u, because K.L is invariant for u if and only if K is invariant for u * . The following result on projections will be used frequently and tacitly. 2.3.2. Theorem. Let p, q be projections on a Hilbert space H. Then the following conditions are equivalent: (1) p < q. (2) pq = p. (3) qp = p. (4) p(H) C q(H). ( 5 ) II p( x ) II < II q ( x ) II (x E H). (6) q - p is a projection. Proof. Equivalence of Conditions (2),(3), and (4) is clear, as are the implications (2) => (6) => (1). We show (1) => (5) => (2), and this will prove the theorem. If we assume Condition (1) holds, IIq(x)112-lIp(x)1I2 = ((q - p)(x),x) = II(q - p)I/2(x)112 > 0, so Condition (5) holds. If now we assume Condition (5) holds, IIp(l-q)(x)1I < lI(q - q2)(x)1I = 0, and therefore p = pq; that is, Condition (2) holds. 0 Let u: HI  H 2 be a continuous linear map between Hilbert spaces. Since (U(HI )).L = ker( u*), the operator u is Fredholm if and only if u(H I ) is closed in H 2 and the spaces ker( u) and ker( u*) are finite-dimensional. In this case ind( u) = dim(ker( u)) - dim(ker( u *)), and the adjoint of u is also Fredholm and such that ind( u*) = - ind( u). (To see that u* has closed range, recall from Theorem 1.4.15 that there is a continuous linear map v: H 2  HI such that u = uvu. Hence, u* = u*v*u*, so u*v* is an idempotent and u*(H 2 ) = u*v*(H I ). Thus, u*(H 2 ) is closed in HI.) An operator u on a Hilbert space H is normal if and only if lIu(x)1I = lIu*(x)1I (x E H), since ((uu* - u*u)(x),x) = lIu*(x)1I2 -lIu(x)1I2. Thus, ker( u) = ker( u *) if u is normal, and therefore a normal Fredholm operator has index zero. A continuous linear map u:H I  H 2 between Hilbert spaces H I ,H 2 is a partial i30metry if u is isometric on ker( u).L, that is, II u( x) II = II x II for all x E ker( u ).L . 2.3.3. Theorem. Let HI, H 2 be Hilbert spaces and u E B(HI' H 2 ). Then the following conditions are equivalent: (1) u = uu*u. 
2.3. Operators and Sesquilinear Forms 51 (2) u*u is a projection. (3) u u * is a projection. ( 4 ) u is a partial isometry. Proof. The implication (1) => (2) is obvious. To show the converse sup- pose that u*u is a projection. Then lIu( x )11 2 = (u(x), u( x)} = (u*u(x), x) = lIu*u(x )11 2 for all x E HI, so u(l - u*u) = 0, and therefore u = uu*u. To show that (2) => (3), suppose again that u*u is a projection. Then (uu*)3 = (uu*)2, so O'(uu*) C {O, I}. Hence, uu* is a projection by the functional calculus. Thus, (2) => (3), and clearly, then, (3) => (2) by symmetry. To show that (1) => (4), suppose that u = uu*u. Then u*u is the projection onto ker(u)-L, since u* = u*uu*, and ker(u)-L = (u*(H 2 ))- - u*u(H I ). Hence, if x E ker(u)-L, then lIu(x)112 = (u*u(x),x) = (x, x) = "x" 2 . Thus, u is a partial isometry, so (1) => (4). Finally, we show (4) => (2) (and this will prove the theorem). Suppose that u is a partial isometry. If p is the projection of HI on ker( u)-L and x E ker( u ) -L, then (u * u ( x ), x) = II u ( x ) 11 2 = (x, x) = (p( x ), x). If x E ker( u ), then (u*u(x), x) = 0 = (p(x), x). Thus, (u*u(x), x) = (p(x), x} for all x E HI. Hence, u*u = p, so (4) => (2). 0 Just as we can write a complex number as the product of a unitary (= number of modulus one) times a non-negative number, the following result asserts that we can write an operator as the product of a partial isometry times a positive operator. 2.3.4. Theorem (Polar Decomposition). Let v be a continuous linear operator on a Hilbert space H. Then there is a unique partial isometry u E B(H) such that v = ulvl and ker(u) = ker(v). Moreover, u*v = Ivl. Proof. If x E H, "I v 1 ( x ) II 2 = (I v 1 ( x ), I v 1 ( x ) ) = (I V 1 2 ( X ), x) = (v * v ( x ), x) = (v( x), v( x)} = IIv( x) 11 2 . Hence, the map uo: Ivl(H)  H, Ivl(x)  vex), is well-defined and isometric. It is also linear. Therefore, it has a unique linear isometric extension (also denoted uo) to (Ivl(H)) Define u in B(H) by setting u = { Uo, on Ivl(H) .1 0, on Ivl(H) . Then ulvl = v, and u is isometric on ker( u)-L, because ker( u) = 1'l)I(H).l. Thus, u is a partial isometry and ker(u) = ker(lvl). Now (u*v(x), Ivl(y)) = 
52 2. C*-Algebras and Hilbert Space Operators (v( x), v (y)) = (v*v( x), y) (Ivl( x), Ivl(y)) => (u*v( x), z) = (Ivl( x), z) for all z E Ivl(H), and therefore for all z E H. Thus, u*v = Ivl. It follows that ker( Ivl) = ker( v), so ker( u) = ker( v). Now suppose that w E B(H) is another partial isom etry su ch that v = wlvl and ker(w) = ker(v). Then w is equal to u on Ivl(H) and on .1 Ivl(H) = ker(v) = ker(w) = ker(u). Thus, w = u. 0 Before we turn to the correspondence between sesquilinear forms and operators, we present a very brief survey of the basic definitions and facts pertaining to sesquilinear forms, since these are not always covered in books on general functional analysis. The sesquilinear form u on a vector space H is said to be hermitian if O'(y, x) = u(x, y)- for all x, y E H. It follows from the polarisation identity that a sesquilinear form O' is hermitian if and only if O'( x, x) E R (x E H). A sesquilinear form O' is positive if O'( x, x) > 0 for all x E H. Thus, positive sesquilinear forms are hermitian. The inequality lu(x,y)1 < vl O'(x,x) vl u(y,y) (x, Y E H), which holds for any positive sesquilinear form O', is called t he Cau chy- Schwarz inequality. It implies that the function p: x  vi O'( x, x) is a semi-norm on H; that is, p satisfies the axioms of a norm except that the implication p( x) = 0 => x = 0 may not hold. A sesquilinear form O' on a normed vector space H is bounded if there is a positive ntUl1ber M such that 100(x, y)1 < Mllxllilyll (x,y E H). The norm IIO'II of O' is the infimum of all such numbers M. Obviously, 100(x, y)1 < 1I001I1Ixllllyli. A sesquilinear form is continuous if and only if it is bounded. The proofs of these facts are elementary and are the same as for the corresponding results on inner products. 2.3.5. Theorem. If u is an operator on a Hilbert space H, then the sesquilinear form O'u: H 2  C, (x, y)  (u(x), y), is hermitian if and only if u is hermitian, and positive if and only if u is positive. 
2.4. Compact Hilbert Space Operators 53 Proof. We show only the implication, au is positive => u is positive, since the other assertions are easy exercises (if u is positive, use the existence of a positive square root for u to show the converse of the result we are now going to prove). Suppose that au is positive. Then it is hermitian and therefore u is hermitian. To see that a( u) C R +, we show that u - A is invertible if A < o. In this case if x E H, then lI(u - A)(x)112 = ((u - A)(X), (u - A)(X)) = lIu(x)1I2 + IAI211xll 2 - 2A(U(X),x) > IA1211 x 1l 2 . Thus, lI(u - A)(x)11 > IAlllxll, so u - A is bounded below. Hence, (u - A)(H) is closed in Hand ker(u - A) = o. Therefore, (u - A)(H) = ker(u* _ ).1 = ker( u - A).1 = 0.1 = H. Hence, u - A is invertible. 0 By the preceding theorem, if u is a operator on a Hilbert space H, then u is hermitian if and only if (u(x), x) E R (x E H), and u is positive if and only if (u(x), x) > 0 (x E H). 2.3.6. Theorem. Let a be a bounded sesquilinear form on a Hilbert space H. Then there is a unique operator u on H such that a(x,y) = (u(x),y) Moreover, Ilull = lIali. (x,y E H). Proof. Uniqueness of u is obvious. For each y E H, the function H  C, x  a(x, y), is continuous and linear, so by the Riesz representation theorem there is a unique element v(y) E H such that a(x, y) = (x, v(y)) (x E H). Also, IIv(y)1I = sup la(x, y)1 < Ilallllyli. IIxlll The map v: H  H, y  v(y), is linear and Ilvll < lIali. If u = v*, then a(x,y) = (u(x),y) (x,y E H), and also the inequality la(x,y)1 < lIullllxlllly11 which holds for all x,y, implies that lIall < lIuli. Hence, lIall = lIuli. 0 2.4. Compact Hilbert Space Operators In this section we analyse two closely related classes of compact oper- ators, the Hilbert-Schmidt and the trace-class operators. Some of the de- tails are a little technical, but the results are useful to us for the analysis of von Neumann algebras, as well as being important in applications and having intrinsic interest. We begin by looking at general compact operators 
54 2. C*-Algebras and Hilbert Space Operators on a Hilbert space and we strengthen some of the results of Section 1.4 in this case. We shall need to view Hilbert spaces as dual spaces. Let H be a Hilbert space and H* = H as an additive group, but define a new scalar multiplication on H* by setting A.x = x, and a new inner product by setting (x, y)* = (y, x). Then H* is a Hilbert space, and obviously the norm induced by the new inner product is the same as that induced by the old one. If x E H, define v(x) E (H*)* by setting v(x)(y) = (y,x)* = (x, y). It is a direct consequence of the Riesz representation theorem that the map v:H  (H*)*, x  vex), is an isometric linear isomorphism, which we use to identify these Banach spaces. The weak* topology on H is called the weak topology. A net (XA)AEA converges to a point x in H in the weak topology if and only if (x, y) = limA(x.\, y) (y E H). Consequently, the weak topology is weaker than the norm topology, and a bounded linear map between Hilbert spaces is necessarily weakly continuous. The importance to us of the weak topology is the fact that the closed unit ball of H is weakly compact (Banach-Alaoglu theorem). 2.4.1. Theorem. Let u: HI  H 2 be a compact linear map between Hilbert spaces HI and H 2 . Then the image of the closed unit ball of HI under u is compact. Proof. Let S be the closed unit ball of HI. It is weakly compact, and u is weakly continuous, so u(S) is weakly compact and therefore weakly closed. Hence, u(S) is norm-closed, since the weak topology is weaker than the norm topology. Since u is a compact operator, this implies that u(S) is norm-compact. 0 2.4.2. Theorem. Let u be a compact operator on a Hilbert space H. Then both lul and u* are compact. Proof. Suppose that u has polar decomposition u = wlul say. Then lul = w*u, so lul is compact, and u* = lulw*, so u* is compact. 0 2.4.3. Corollary. If H is any Hilbert space, then K(H) is self-adjoint. Thus, ]{(H) is a C*-algebra, since (as we saw in Chapter 1) K(H) is a closed ideal in B(H). An operator u on a Hilbert space H is diagonalisabZe if H admits an orthonormal basis consisting of eigenvectors of u. Diagonalisable operators are necessarily normal, but not all normal operators are diagonalisable. For instance, the bilateral shift is normal (it is a unitary), but it has no eigenvalues. 
2.4. Compact Hilbert Space Operators 55 2.4.4. Theorem. If u is a compact normal operator on a Hilbert space H, then it is diagonalisable. Proof. By Zorn's lemma there is a maximal orthonormal set E of eigen- vectors of u. If [{ is the closed linear span of E, then H = K EB K.l, and K reduces u. The restriction UK.L: I{.l  [{.l is compact and normal. An eigenvector of UK.L is one for u also, so by maximality of E, the operator UK.l has no eigenvectors, and therefore 0'( UK.L) = {o} by Theorem 1.4.11. Hence, IIUK.L II = r(uK.L) (by normality) = 0, so K.l = o. Thus, I{ = H and E is an orthonormal basis of eigenvectors of u, so U is diagonalisable.D If H is a Hilbert space, we denote by F(H) the set of finite-rank operators on H. It is easy to check that F(H) is a self-adjoint ideal of B(H). 2.4.5. Theorem. If H is a Hilbert space, then F(H) is dense in [«H). Proof. Since F(H)- and ]{(H) are both self-adjoint, it suffices to show that if u is a hermitian element of [{(H), then U E F(H)-. Let E be an orthonormal basis of H consisting of eigenvectors of u, and let c > O. By Theorem 1.4.11 the set S of eigenvalues ,,\ of U such that 1,,\1 > c is finite. From Theorem 1.4.5 it is therefore clear that the set S' of elements of E corresponding to elements of S is finite. Now define a finite-rank diagonal operator v on H by setting v(x) = "\x if xES' and ,,\ is the eigenvalue corresponding to x, and setting v(x) = 0 if x E E \ S'. It is easily checked that Ilv - ull < sUP-XEu(u)\S 1,,\1 < c. This shows that U E F(H)-. 0 If x, yare elements of a Hilbert space H we define the operator x 0 y on H by (x 0 y)(z) = (z, y)x. Clearly, IIx0yli = IIxlillyll. The rank of x0y is one if x and yare non-zero. If x, x', y, y' E Hand U E B(H), then the following equalities are readily verified: (x 0 x')(y 0 y') = (y, x')(x 0 y') (x0Y)*=Y0 x u(x0y)=u(x)0Y (x 0 y)u = x 0 u*(y). The operator x 0 x is a rank-one projection if and only if (x, x) = 1, that is, x is a unit vector. Conversely, every rank-one projection is of the form x 0 x for some unit vector x. Indeed, if el,. . . , en is an orthonormal set, in H, then the operator 2::j=1 ej 0 ej is the orthogonal projection of H onto the vector subspace Cel + . . . + Ce n . 
56 2. C*-Algebras and Hilbert Space Operators If u E B(H) is a rank-one operator and x a non-zero element of its range, then u = x 0 y for some y E H. For if z E H, then u(z) = r(z)x for some scalar r(z) E C. It is readily verified that the map z  r(z) is a bounded linear functional on H, and therefore, by the Riesz representation theorem, there exists y E H such that r( z) = (z, y) for all z E H. Therefore, u = x 0 y. 2.4.6. Theorem. If H is a Hilbert space, then F(H) is linearly spanned by the rank-one projections. Proof. Let u E F( H) and we shall show it is a linear combination of rank- one projections. The real and imaginary parts of u are in F(H), since F(H) is self-adjoint, so we may suppose that u is hermitian. Now u = u+ - u-, and by the polar decomposition lul E F(H), so u+ and u- belong to F(H). Hence, we may assume that u > O. The range u(H) is finite-dimensional, and therefore it is a Hilbert space with an orthonormal basis, el,..., en say. Let P = E j 1 ei 0 ei, so P is the projection of H onto u(H). Then u = pu = u 1 / 2 pu 1 / 2 => u = Ej=l xi 0 xi' where xi = u 1 / 2 (ei). Now xi = Aifj for some unit vector fi and scalar Ai, so u = Ej=l I A il 2 fi 0 fi, and since the operators fi 0 fi are rank-one projections we are done. 0 2.4.7. Theorem. If H is a Hilbert space and I a non-zero ideal in B(H), then I contains F(H). Proof. Let u be a non-zero operator in I. Then for some x E H we have u( x) i= o. If p is a rank-one projection, then p = y 0 y for some unit vector y E H, and clearly there exists v E B(H) such that vu(x) = y (take v = (y 0 u(x))/lIu(x)1I2, for instance). Hence, p = vu(x 0 x)u*v*, so P E I as u E I. Thus, I contains all the rank-one projections and therefore by Theorem 2.4.6 it contains F(H). 0 If u: H -+ H' is a unitary between Hilbert spaces Hand H', then the map Ad u: J«H)  K(H'), v  uvu*, is a *-isomorphism. In fact, all *-isomorphisms between I«H) and I«H') are obtained in this way: 2.4.8. Theorem. Let Hand H' be Hilbert spaces and suppose that the map cp: K(H) -+ K(H') is a *-isomorphism. Then there exists a unitary u: H -+ H' such that c.p = Ad u. Proof. Let E be an orthonormal basis for H, and for e E E let Pe = e 0 e. Then Pe is a rank-one projection and peI«H)pe = Cpe. Hence, qe = CP(Pe) is a projection on H' such that qeI«H')qe = c.p(PeI«H)Pe) = c.p(Cpe) = Cqe. It is easily inferred from this that qe is also of rank one. Thus, we 
2.4. Compact Hilbert Space Operators 57 may write qe = e Q9 e for a unit vector e in H'. If e, 1 are distinct elements of E, then (}, e)e Q9} = qeqf = CP(PePf) = (I, e)ep( e Q9 I) = 0, and therefore e and ! are orthogonal. Thus, E = {e leE E} is an orthonormal set in H'. We claim it is maximal; that is, it is an orthonormal basis for H'. For if we suppose the contrary, then there is a unit vector x of H' orthogonal to E. Reasoning as above, but this time using ep-l instead of ep, there is an element y of H such that ep-l(x Q9 x) = y Q9 y, and the set E U {y} is orthonormal in H. This contradicts the fact that E is an orthonormal basis. This argument therefore shows that E is an orthonormal basis for H' as claimed. For e, 1 E E let qef = ep( e Q9 I). Then qee = qe, and qefqgh = (g, f)qeh (e,f,g,hEE), sInce (e Q9 f)(g Q9 h) = (g, f)e Q9 h. Because qef = qeqef, the range of qef is Ceo Hence, qef can be written in the form e Q9 y for some unit vector y E H'. Since qfe = q:f = y Q9 e, and qfe has range C}, w have y = XefJ for some scalar Aef of modulus one. Thus, qef = Aefe Q9 f. Since qeg = qefqfg, Aege Q9 9 = Aef( e Q9 })Afg(} Q9 g) = AefA fge Q9 g. Therefore, Aeg = AefAfg. Observe also that X eg = Age, since q:g = qge. Thus, if we fix an element, f say, in E and set J-Le = Aef for all e E E, we get Aeg = J-Lefi,g. Let u: H  H' be the unitary such that u( e) = J-Le e for all e E E. Then Adu(e Q9 g) = u(e) Q9 u(g) = J-Lee Q9 J-Lg9 = Aege Q9 9 = ep(e Q9 g). From this it follows that Ad u and ep are equal at x Q9 y for x, y in the linear span of E, and hence for all x,y in H, since E has closed linear span H. Thus, Ad u and ep are equal on all the rank-one operators on H, and since these have closed linear span K(H), we have Adu = cpo 0 We make a few observations now which we shall need in the proof of the next theorem, and which are also of independent interest. Let 0 be a locally compact Hausdorff space. For W E 0, denote by Tw the character on CoCO) given by evaluation at w: Tw(f) = f(w). If WI, . . . , W n are distinct points of 0, then T WI , . . . , T W n are linearly indepen- dent. For if Al T Wl + . . . + An T W n = 0 and we fix i, then by Urysohn's lemma we may choose f E CoCO) such that f(Wi) = 1 and f(wj) = 0 for j :F i. Hence, 0 = Ej=1 Ajf(wj) = Ai. It follows that if CoCO) is finite-dimensional, then 0 is finite. 
58 2. C*-Algebras and Hilbert Space Operators From this observation we show that the projections linearly span an abelian finite-dimensional C*-algebra. We may suppose the algebra is of the form Co(r!) by the Gelfand representation. Then r! is finite and therefore discrete, so the characteristic functions of the singleton sets span Co(r!). Suppose now that A is an arbitrary finite-dimensional C*-algebra. It is linearly spanned by its self-adjoint elements, and they in turn are linear combinations of projections by what we have just shown, so it follows that A is the linear span of its projections. If p is a finite-rank projection on a Hilbert space H, then the C* -algebra A = pB(H)p is finite-dimensional. To see this, write p = Ej=l ej 0 ej, where el,..., en E H. If u E B(H), then n n pup = L (ej 0 ej)u(ek 0 ek) = L (u(ek), ej)ej 0 eke j,k=l j,k=l Hence, A is in the linear span of the operators ej 0 ek (j, k = 1, . . . , n), and therefore dim(A) < 00. A closed vector subspace !{ of H is invariant for a subset A C B(H) if it is invariant for every operator in A. If A is a C*-subalgebra of B(H), it is said to be irreducible, or to act irreducibly on H, if the only closed vector subspaces of H that are invariant for A are 0 and H. The concept of irre- ducibility is of great importance in the representation theory of C*-algebras which we shall be taking up in Chapter 5. The following theorem gives a nice connection between irreducibility and the ideal of compact operators, and will be needed in succeeding chapters. 2.4.9. Theorem. Let A be a C*-algebra acting irreducibly on a Hilbert space H and having non-zero intersection with I«H). Then K(H) C A. Proof. The intersection A n !{(H) is a non-zero self-adjoint set, so it contains a non-zero self-adjoint element, u say. Now r( u) = II u II > 0, so a( u) contains non-zero elements. Hence, by Theorem 1.4.11 u admits a non-zero eigenvalue, A say. By the same theorem, the non-zero points of a( u) are isolated, so if f is the characteristic function of {A} on a( u), then f is continuous, and p = f( u) is a projection in A. Moreover, p is non-zero because f is non-zero. If z is the inclusion function of a( u) in C, then (z - A)f = 0, so (u - A)p = 0, and therefore p(H) C ker(u - A). By Theorem 1.4.5 the space ker( u - A) is finite-dimensional, so p is therefore of finite rank. Let q be a non-zero projection in A of minimal finite rank. Then the C*-algebra qAq is finite-dimensional; therefore, it is the linear span of its projections, by the remarks preceding this theorem. However, the minimal rank assumption on q implies that the only projections in qAq can be 0 
2.4. Compact Hilbert Space Operators 59 and q, so qAq = Cq. Now let y be a non-zero element of q(H). If K is the closure of the set of vectors u(y) (u E A), then K is a vector subspace of H invariant for A, and is non-zero since it contains y = q(y). It follows from the irreducibility of A that K = H. Hence, if x is an arbitrary element of q(H), then x = lim n -+ oo un(y) for some sequence (un) in A. Therefore, x = lim n -+ oo qunq(y), because x = q( x) and y = q(y). But qunq = Anq for some An E C, because qAq = Cq, so x E Cy. This shows that q(H) = Cy, and therefore q = y 0 y. Now suppose that x is an arbitrary unit vector of H. As before, there are operators Un E A such that x = lim n -+ oo un(y), so x 0 x = lim un(y) 0 un(y) = lim un(y 0 y)u = lim unqu. n-+oo n-+oo n-+oo Hence, x0x E A. Therefore, all rank-one projections are in A, so F(H) C A by Theorem 2.4.6, and therefore I«(H) C A, by Theorem 2.4.5. 0 Before we introduce the Hilbert-Schmidt operators, it is convenient to make a few observations about summable families. Let (X..\)..\EA be a family of elements of a Banach space X. Let A' denote the set of all non-empty finite subsets of A, and for each F E A', set x F = E..\EF X..\. Then (x F )FEA' is a net where F < G in A' if F C G. We say (X..\)..\EA is 3ummable to an element x E X if the net (XF )FEA' converges to x, and in this case we write x = E..\EA X..\. If all x..\ are in R+, then the family (X..\)..\EA is summable if and only if sup F E..\EF x..\ < +00, and in this case L x..\ = sup L X..\. ..\EA FEA' ..\EF We thus can use the right-hand side of this expression to define E..\EA x..\ whether (X..\)..\EA is summable or not, provided all x..\ are in R+. Let u be an operator on a Hilbert space H, and suppose that E is an orthonormal basis for H. We define the Hilbert-Schmidt norm of u to be IIul12 = (L Ilu(x)112)1/2. xEE This definition is independent of the choice of basis. To see this let E' be another orthonormal basis for H. Then for each finite non-empty set F of E, L Ilu(x)112 = L L l(u(x),y}12 xEF xEF yEE' = L L l(u(x),y}12 yEE' xEF < L lIu*(y )/1 2 , yEE' 
60 2. C*-Algebras and Hilbert Space Operators so L lIu(x)1I2 < L lIu*(y)1I2. xEE yEE' By symmetry, therefore, L lIu(x)1I2 = L lIu*(x)1I2 = L lIu(y)1I2. xEE xEE yEE' This shows not only that the expression for lIull2 is independent of the choice of basis, but also that II u * 112 = II U 112 . An operator u is a Hilbert-Schmidt operator if lIull2 < +00. We denote the class of all Hilbert-Schmidt operators on H by L 2 (H). 2.4.1. Ezample. Let (en)=l be an orthonormal basis for a Hilbert space H and let u be an operator on H diagonal with respect to (en), with diagonal sequence (An). Then u is a Hilbert-S chmidt operator if and only if E  llAnl 2 < 00, since lI u ll2 = v E - l I A nI 2 . More generally, if u is an arbitrary operator in B(H) and (an,m) is its matrix with respect to the basis (en), so that an,m = (u( em), en), then from the definition 00 00 lIull2 = L L la n ,mI 2 , m=l n=l and, therefore, u is Hilbert-Schmidt if and only if Em En la n ,ml 2 < 00. 2.4.2. Ezample. Let L 2 (T) and L2(T2) denote the Lebesgue L 2 -spaces of T and T 2 with the usual measures, normalised arc length m (that is, m is the Haar measure of T), and the corresponding product measure m x m. By elementary measure theory, C(T) and G(T 2 ) are L2-dense in L 2 (T) and L 2 (T 2 ), respectively. Define en E G(T) by en(A) = An, and e nm E G(T 2 ) by enm(A,J.L) = AnJ.Lm. These sequences are orthonormal in the corresponding L 2 -spaces. By the Stone-Weierstrass theorem, the sup- norm closed linear span of (en) in G(T) is G(T) itself, since this closed span is a C*-subalgebra separating the points of T and containing the constants. By similar reasoning the sup-norm closed linear span of (e nm ) is G(T 2 ). Thus, (en) and (e nm ) have L 2 -norm dense linear span in, and are therefore orthonormal bases of, L 2 (T) and L 2 (T 2 ), respectively. Let k be an element of L 2 (T 2 ). Then for almost all A E T,  Ik(A,)f()ldn1 < 00, 
2.4. Compact Hilbert Space Operators 61 SInce J J Ik(A,Jl)f(Jl)\ d(m x m)(A,Jl) < (J J Ik( A, JlW d(m x m )(A, Jl ))1/2( J J If(Jl W d(m x m )(A, Jl ))1/2 = IIk1l 2 11f1l2. Define the integral operator u = Uk on L 2 (T) by (Uf)(A) = J k(A,Jl)f(Jl)dmJl for almost all A. That u(f) E L 2 (T) follows by another application of the Cauchy-Schwarz inequality, J I(Uf)(A)12 dmA = J I J k(A,Jl)f(Jl) dmJll 2 dmA < J(J Ik(A, JlW dmJl)(J If(Jl)1 2 dmJl) dmA = IIkllllfll. Hence, u is bounded with norm lIull < II k 1l 2 . Now we compute lIull2. From the definition, Ilull = L lIu(e n )1I2 nEZ = L I(u(e n ), e m )1 2 n,mEZ = L I J(U(en))(A)em(A) dmAI 2 n,mEZ = L I J J k(A, Jl)en(Jl)em(A) dmJl dmAI 2 n,mEZ = L l(k,e m ,-n}12. n,mEZ Thus, lIull2 = II k 1l 2 , and therefore u is a Hilbert-Schmidt operator. 2.4.10. Theorem. Let u, v be operators on a Hilbert space H, and A E C. Then (1) lIu + Vll2 < IIul12 + II v ll2 and IIAul12 = IAlllul12; (2) Ilull < lI u ll2; (3) IIuvl12 < lIullllvl12 and IIuvll2 < lIul12l1vll. 
62 2. C*-Algebras and Hilbert Space Operators Proof. If F is is any finite set of orthonormal vectors of H, then II: I/u(x) + v(x)1/2 < 1I:(l/u(x)1/ + I/v(x)I/)2 V xEF V xEF < I I: 1/ u( x ) 1/2 + I I: 1/ v( x ) 1/2. V xEF V xEF It follows that lIu + Vll2 < lIull2 + IIVI/2. The equality IIAull2 = IAII/ull2 is trivial. If x is a unit vector of H, there is an orthonormal basis E containing x. Hence, lIu(x)1I2 < EyEE l/u(y)1/2 = I/ul/, so I/ul/ < I/UI/2. If E is an arbitrary orthonormal basis of H, then I/uvll = I: lIuv( X )1/2 < lIu 1/2 I: I/v( X )11 2 = lIu 112I1vll. xEE xEE Hence, IIuvll2 < IIullllvl/2. Therefore, IIuvll2 = IIv*u*II2 < IIv*IIIIu*II2 - lIull2l1vll. 0 2.4.11. Corollary. The set L 2 (H) is a self-adjoint ideal of B(H), and a normed *-algebra (that is, a normed algebra with an isometric involution), where the norm is given by u  I/UIl2. Note that if x, y E H, then I/x 0 yl/2 = I/xl/I/yll, so x 0 Y E L 2 (H). Hence, F(H) C L 2 (H). 2.4.12. Lemma. Let Ul, U2 be Hilbert-Schmidt operators on a Hilbert space H. If E is an orthonormal basis of H and v = U;U2, then the family ((v(x), X))xEE is absolutely summable, that is, EXEE I{v(x), x)1 < +00, and 3 I: (v(x), x) = t I: i k llu2 + ikull/. xEE k=O Proof. If F is a finite non-empty subset of E, then I: I{v(x), x)1 = I: I{U2(X), Ul(X)) I xEF xEF < I: IIU2(X)I/I/ U l(X)1I xEF < II: I/ U 2(X)1/2 II: I/ U l(X)1/2. V xEF V xEF 
2.4. Compact Hilbert Space Operators 63 Hence, ({v( x), x) )xEE is absolutely summable. Also, 3 (v(x), x) = (U2(X), U1(X)) = t L i k llu2(X) + iku1(x)1I2 k=O by the polarisation identity, so 3 3 L (v(x), x) = t L i k L lIe u2 + iku1)(X )11 2 = t L i k llU2 + iku111, xEE k=O xEE k=O which is the required result. o If u is an operator on a Hilbert space H, we define its trace-cla33 norm to be lIulh = l"uI1/211. If E is an orthonormal basis of H, then Ilulh = L (Iul(x), x). xEE If IIul11 < +00, we call u a trace-cla33 operator. The connection between trace-class operators and Hilbert-Schmidt operators is given in the follow- ing result. 2.4.13. Theorem. Let v be an operator on a Hilbert space H. The following conditions are equivalent: (1) v is trace-class. (2) Ivl is trace-class. (3) Iv1 1 / 2 is a Hilbert-Schmidt operator. (4) There exist Hilbert-Schmidt operators Ul, U2 on H such that v = U1 U2. Proof. The implications (1) => (2) => (3) => (4) are easy (for (3) => (4) use the polar decomposition of v), so we prove (4) => (1) only. Assume that v = U1 U2, where U1, U2 E L 2 (H). If v = wlv I is the polar decomposition of v, then Ivl = w*v = (W*U1)U2. If E is any orthonormal basis of H, then by the "polarisation identity" of the preceding lemma, Ex E E { I v I ( x ), x) < + 00, so II vIII < + 00. 0 It is clear from Theorem 2.4.13 that if v is a trace-class operator and u is an arbitrary operator on H, then uv and vu are also trace-class operators. We define the trace of a trace-class operator v to be tr(v) = L(v(x),x), xEE where E is any orthonormal basis of H. By Lemma 2.4.12 the definition of tr is independent of the choice of orthonormal basis. 
64 2. C*-Algebras and Hilbert Space Operators 2.4.14. Theorem. Let u and v be operators on a Hilbert space H. Then tr(uv) = tr(vu) if either (1) u and v are both Hilbert-Schmidt operators, or (2) v is trace-class. Proof. In Case (1), 3 tr(uv) = t L ikllv + iku*lI k=O 3 = t L ikll(v + iku*)*II k=O 3 = t L ikllu + ikv*ll k=O = tr(vu). In Case (2) v = UI U2 for some U1, U2 E L 2 (H), so tr( uv) = tr(( UU1 )U2) = tr(u2(uuI)) (by Case (1)) = tr(uI(u2U)) (same reason) = tr(vu). 0 There are similar results for the trace-class norm as for the Hilbert- Sclunidt norm, but the proofs require more work: 2.4.15. Theorem. Let u, v be operators on a Hilbert space H and A E C. (1) lIu + VIII < lIulh + Ilvlh and IIAulh = IAlli u lll. (2) lIull < Ilulh = lIu*III. (3) Iluvlh < Ilullllvlll and lIuvlll < Ilulhllvll. Proof. BeginningwithCondition(2)wehavellulh = Illull/211 > Illul l / 2 112 = Illulll = lIull. If u = wlul is the polar decomposition of u, then uu* = wluI 2 w*, so lu*1 2 = (wlulw*)2, and therefore lu*1 = wlulw*. Hence, lIu*lh = tr(lu*1) = tr(wlulw*) = tr(w*u) = tr(lul) = Ilulh. This proves Condition (2). Next, we show that Condition (3) holds. Let vu = w'lvul be the polar decomposition of vu and w" = w'*vw. Then Ivul = w'*vu = w"lul. Hence, Ivul 2 = lulw"*w"lul < lu1 2 11w"112 < lu1 2 11v1l 2 , so Ivul < lulllvil by Theorem 2.2.6. Consequently, if E is an orthonormal basis for H, Ilvulll = L (Ivul(x), x) xEE < L (Iul(x), x)llvll xEE = Il u llll1 v ll. 
2.4. Compact Hilbert Space Operators 65 Also, Iluvlh = l'v*u*lIl < IIvllllulh. This proves Condition (3). Finally, we show Condition (1). The equality II-Xulh = I-Xillulh follows from the corresponding statement for the norm 11.112. Suppose that u and v are trace-class operators, and let u = wlul, v = w'lvl, and u + v = wIt lu + vi be the respective polar decompositions. Then lu + vi = w"*(u + v) = w"*wlul + w"*w'lvl. If E is an orthonormal basis of H, Ilu+vlll = L(lu+vl(x),x) xEE = I L (w"*wlul(x), x) + L (w"*w'lvl(x), x)1 xEE xEE < L 1(l u l l / 2 (x), luI 1 / 2 W*W"(x))1 + L 1(l v I 1 / 2 (X), IvI 1 / 2 W'*W"(x)) xEE xEE < (L Ill u l l / 2 (x)1I 2 )1/2(L II' u l l / 2 w*w"(x)1I2)1/2 xEE xEE + (L Ill v I 1 / 2 (x)1I 2 )1/2(L II'vI I / 2 w'*w"(x)1I2)1/2 xEE xEE = lIull/211IuI1/2w*w"112 + IIvll/211Ivll/2w'*w"1I2 < lIull/21Iull/2 + IIvll/2I1vll/2 = lIulh + IIvlll' so lIu + vlh < Ilulh + Il v lll. o If H is a Hilbert space, we denote the set of trace-class operators on H by L 1 (H). From the preceding theorem it is clear that Ll(H) is a self- adjoint ideal of B(H), and the function u  lIulh is a norm on L 1 (H) making it a normed *-algebra. 2.4.16. Theorem. Let H be a Hilbert space. The function tr: L 1 ( H) --+ C, u....... tr( u ), is linear, and I tr( vu)1 < IIvllllu Ih (v E B(H), u E L 1 (H)). Proof. Linearity of the trace is clear. To show the inequality let u = wfuf be the polar decomposition of u and let E be an orthonormal basis of H. 
66 2. C*-Algebras and Hilbert Space Operators Then I tr(vu)1 = I L (vu(x), x)1 xEE = 1 L(l u l l / 2 (x), lul l / 2 W*v*(x))1 xEE < L IIlul l / 2 (x)llllIul l / 2 w*v*(x)11 xEE < (L IIl u l l / 2 (X)1I 2 )1/2(L IIlul l / 2 w*v*(x)1I2)1/2 xEE xEE = lIull/211Iull/2w*v* 112 < II u II  /2111 U 11 /211211 V II = lIulhllvll, so 1 tr(vu)1 < lIulllllvll. 0 If x,y E H, then IIx 0 ylh = Ilxllllyll and tr(x 0 y) - (x,y). The inclusions F(H) C L l (H) C L 2 (H) hold. 2.4.17. Theorem. Let H be a Hilbert space. Then for i = 1,2, the ideal Li(H) is contained in K(H), and F(H) is dense in Li(H) in the norm 1I.lli. Proof. An easy exercise. 0 2.5. The Spectral Theorem The normal operators form one of the best understood and most tract- able of classes of operators. The principal reason for this is the spectral theorem, a powerful structure theorem that answers many (not all) ques- tions about these operators. In this section we actually prove a more general result than the spectral theorem for normal operators (Theorem 2.5.6), and we get this extra useful generality without any increase in difficulty of the proofs. Indeed, the more general situation illustrates nicely the connection between spectral measures and representations of abelian C*-algebras. Let n be a compact Hausdorff space and H a Hilbert space. A spectral measure E relative to (n, H) is a map from the a-algebra of all Borel sets of n to the set of projections in B(H) such that (1) E(0) = 0, E(O) = 1; (2) E(5 1 n 52) = E(5 1 )E(5 2 ) for all Borel sets 5 1 ,5 2 of n; (3) for all x, y E H, the function Ex,,: 5  (E(5)x, y), is a regular Borel complex measure on O. Denote by M(O) the Banach space of all regular Borel complex mea- sures on n, and by Bex>(O) the C*-algebra of all bounded Borel-measurable complex-valued functions on O. 
2.5. The Spectral Theorem 67 2.5.1. Ezample. Let n be a compact Hausdorff space and let j.L be a positive regular Borel measure on n. Define Mcp E B(L 2 (n,j.L)) by Mcp(/) = c.p 1 (I E L 2 (n,j.L)). That Mcp is bounded is given by IIM'P(f)II = f Icp(w)f(wW dJlw < Ilcpll;" f If(w)12 dJlw, which implies that IIMcpl1 < 11c.p1100. The operator Mcp is called a multiplica- tion operator. The map Loo(n, j.L) -t B(L 2 (n, j.L)), c.p  Mcp, is a *-homomorphism of C*-algebras. In particular, the adjoint of Mcp is MtjJ, and Mcp is normal. In fact, these operators are typical of all normal operators (see Section 4.4). If S is a Borel set of n, then X s (the characteristic function of S) is a projection in Loo(n, j.L), so E(S) = M xs is a projection in B(L2(n, j.L)). The map E:S  E(S) is a spectral measure relative to the pair (n,L 2 (n,j.L)). Since the multiplication operators are a very important class we linger with this example a little longer to show that if c.p E Loo(n, j.L}, then IIMcp II = 11c.p1100. For if this is false, then there exists a positive ntUl1ber £ such that IIc.plloo - £ > IIMcpl1 and, therefore, there is a Borel set S of n such that j.L(S) > 0 and 1<p(w)1 > IIMcpli + £ for all w E S. Since j.L is regular, j.L(S) = sup{j.L(I<) I I{ is compact and K C S}, so we may suppose that S is compact. Then j.L(S) < 00, again by regularity of J-l. However, IIMcpIl2J-l(S) > IIMcp(xs)ll = f Icp(w)xs(wW dJlw > f(IIM'P1l +€?Xs(w)dJlw = (liMcpli + £)2 J-l(S), and therefore after dividing by j.L(S), we get IIMcpl1 > IIMcpl1 + €, a contra- diction. This shows that IIMcpl1 = 11c.p1100 as claimed. This result means that the map c.p  Mcp, is in fact an isometric *-isomorphism of Loo(n, J-l) onto a C*-subalgebra of B(L 2 (n, J-l)). We there- fore have a(Mcp) = a(c.p) (the spectrum of c.p in Loo(o., J-l)). 
68 2. C*-Algebras and Hilbert Space Operators 2.5.1. Lemma. Let f! be a compact Hausdorff space, let H be a Hilbert space, and suppose that JLx,y E M(f!) for all x, y E H. Suppose also that for each Borel set S of f! the function as: H 2 -+ C, (x, y)  JLx,,1(S), is a sesquilinear form. Then for each f E Boo(f!) the function U f: H 2 ..... C, (x, y) 1-+ J f dJ.Lx,1/, is a sesquilinear form. Proof. Suppose first that I is simple, so we can write I = Ej=l AjXSj' where Sl,. . . , Sn are pairwise disjoint Borel sets of f!, and AI,..., An are complex numbers. Then n n J f dJ.Lx,1/ = L Aj J XS j dJ.Lx,1/ = L AjJ.Lx,1/( Sj). j=l j=l The set of sesquilinear forms on H is a vector space with the pointwise- defined operations, and we have just shown that a/is a linear combination of the a Sj , so a/is a sesquilinear form. Now suppose that I is an arbitrary element of Boo(f!). Then I is the uniform limit of a sequence (In), where each In is a simple function in Boo(f!). Hence, J Iin - II dIJLx,,11 < IIIn - IllooIJLx,,1I(f!), so J I dJLx,'1 = limn-+ooJ IndJLx,'1 for each x,y E H. It follows immediately that a/ is a sesquilinear form on H. 0 2.5.2. Theorem. Let f! be a compact Hausdorff space, H a Hilbert space, and E a spectral measure relative to (n, H). Then for each I E Boo(f!) the function Uf:H 2 ..... C, (x,y) 1-+ J f dE x,1/' is a bounded sesquilinear form on H, and lIa/11 < 1111100. Proof. That a/is a sesquilinear form follows from the preceding lemma, so we need only show Iia / II < 1111100. Suppose that f! = Sl u. . . USn, where Sl, . . . , Sn are pairwise disjoint Borel sets of f!. Then n n L I{E(Sj)(x), y)1 = L I{E(Sj)(x), E(Sj)(y))1 j=l j=l n n < (L IIE(Sj)(x)11 2 )1/2(L IIE(Sj)(y)112)1/2 j=l j=l = IIE(f!)(x)IIIIE(f!)(y)11 = Ilxlillyll. 
2.5. The Spectral Theorem 69 Hence, IIEx,,1I < Ilxlillyli. Therefore, Iff dE.", I < IIflloollE.",1I < IIflloollxllllyll. so II (1 f II < II f II 00 · o 2.5.3. Theorem. Let n be a compact Hausdorff space, H a Hilbert space, and E a spectral measure relative to (n, H). Then for each f E Boo(n) there is a unique bounded operator u on H such that (u(x), y) = J f dE.", (x, Y E H). Proof. Immediate from the preceding theorem and Theorem 2.3.6. 0 We write J f dE for u and call it the integral of f with respect to E. Note that J Xs dE = E(S) for each Borel set S. 2.5.4. Theorem. With the same asswnptions on n, H, and E as in the preceding theorem, the map tp: Boo(f2) -+ B(H), f 1-+ J f dE, is a unital *-homomorphism. Proof. Linearity is routine and boundedness follows from Theorems 2.5.2 and 2.3.6. To show that cp(f g) = cp(f)cp(g) and cp(l) = (cp(f))-, we need only show these results when f, 9 are simple, because the simple elements of Boo(n) are dense. Hence, we may reduce further and suppose that f = Xs and 9 = Xs' by linearity of cp and the fact that all simple elements of Boo(n) are linear combinations of such characteristic functions. Then cp(fg) = J XsXs' dE = E(S n S') = E(S)E(S') J XS dE J XS' dE = cp(f)cp(g). Also, cp(l) = cp(f) = E( S) = (cp(f)) - . 0 2.5.5. Theorem. Let n be a compact Hausdorff space and H a Hilbert space, and suppose that cp: C(n)  B(H) is a unital *-homomorpmsm. Then there is a unique spectral measure E relative to (n, H) such that tp(J) = J f dE (f E C(n)). Moreover, if u E B(H), then u commutes with cp(f) for all f E C(n) if and only ifu commutes with E(S) for all Borel sets S ofn. 
70 2. C*-Algebras and Hilbert Space Operators Proof. If x, y E H, then the function Tx,y: C(n)  c, f  (cp(f)(x), y), is linear and II T x ,y II < II x 1111 y II. By the Riesz- Kakutani theorem, there is a unique measure J.Lx,y in M(n) such that Tx,y(f) = J f dJ.Lx,y for all f E C(n). Also, lIJ.Lx,yll = IITx,yll. Since the function H 2  C, (x,y)  (cp(f)(x),y), is sesquilinear, the maps from H to M(n) given by x  J.Lx,y and y  J.Lx,y are, respectively, linear and conjugate-linear. Hence, for each f E Bex>(n) the function H 2 -+ C, (x, y) 1-+ J f dJ1.x,y, is a sesquilinear form, by Lemma 2.5.1. Also, Iff dJ1.x,yl < IIflloollJ1.x,yll < IIflloollxllllyll, so this sesquilinear form is bounded and its norm is not greater than Ilfllex>. By Theorem 2.3.6, there is a unique operator, 1/;(f) say, in B(H) such that (tjJ(f)(x), y} = J f dJ1.x,y Moreover, 111/J(f)1I < Ilfllex>. Now suppose that f E C(n). Then (x, Y E H). (tjJ(f)(x), y} = J f dJ1.x,y = TX,y(f) = (tp(f)(x), y} so 1/;(f) = <p(f). It is straightforward to check that the map (x, Y E H), 1/;: Bex>(n)  B(H), f  1/;(f), is linear and we already know it is norm-decreasing. We show now that 1/; is a *-homomorphism. If f E C(n) and 1 = f, then cp(f) is hermitian, so J f dJ.Lx,x = (cp(f)( x), x) is a real number. Thus, the measure J.Lx,x is real, that is, flx,x = J.Lx,x, and therefore if f is an arbitrary function in Bex>(n) such that 
2.5. The Spectral Theorem 71 1 = f, then (1/;(f)( x), x) = J f dJLx ,x is real. Since x is arbitrary, this shows that 1/;(f) is hermitian. Therefore, 1/; preserves the involutions. Let f E Boo(n) and x E H. Assertion: If the equation (1/;(fg)(x),x) = (1/;(f)1/;(g)(x),x) (1) holds for all 9 E C(n), then it also holds for all 9 E Boo(n). Observe that Eq. (1) is equivalent to J gl dJ.Lz,z = J 9 dJ.Lz,,p(/)(z) " To prove the assertion, note that if Eq. (1) holds for all 9 E C(n), then the regular measures fd/-lx,x and /-lx,1/J(J)(x) are equal because Eq. (2) holds for all 9 E C(O). Hence, Eq. (2) holds for all 9 E Boo(O); that is, Eq. (1) holds for all such g, as claimed. Since c.p is a *-homomorphism, Eq. (1) holds for all f, 9 E C(n). Hence, by the assertion, Eq. (1) holds if f E C(n) and 9 E Boo(n). Replacing f, 9 with their conjugates, we get (1/;(19)( x), x) == (1/;(1)1/;(9)( x), x). Taking conjugates of both sides of this equation and using the fact that 1/; preserves the involutions, we get (2) (1/;(g f)( x), x) = (1/;(g )1/;(f)( x ), x), (3) for all 9 E Boo(n) and all f E C(n). Using the assertion again (with the roles of f and 9 interchanged), we get Eq. (3) holds for all f, 9 E Boo (0). Since x was an arbitrary element of H, this implies that 1/;(g f) == 1/;(g )1/;(f), so 1/; is a homomorphism. If S is a Borel set of 0, we put E(S) == 1/;(xs). Obviously, E(S) is a projection on H, and it is easily verified that the map E: S  E( S) from the a-algebra of Borel sets of 0 to B(H) is a spectral measure relative to (0, H)-we have Ex,y = /-lx,y E M(n), since Ex,y(S) == (E(S)(x), y) == (1/;(Xs )(x), y) == J XS d/-lx,y. If f E Boo(n), then ((J 1 dE)(x), y) = J 1 dEz,y = J 1 dJ.Lz,y = (1jJ(f)(x), y), so 1/;(f) = J f dE. In particular, c.p(f) == J f dE for all f E C(n). To see uniqueness of E, suppose that E' is another spectral mea- sure relative to (n, H) such that 'P(f) == J f dE' for all f E C(O). Then J f dE,y = (c.p(f)(x), y) == J f dEx,y. Hence, E,y = Ex,y, and therefore (E'(S)(x),y) = (E(S)(x),y), so E = E'. Now suppose u is an operator on H commuting with all of the elements of the range of c.p. Then if f E C(O), J f d/-lu(x),y == (1/;(f)u( x), y) == 
72 2. C*-Algebras and Hilbert Space Operators (u1jJ(f)(x), y) = (1jJ(f)(x), u*(y)) = J f dllx,u.(y). Hence Eu(x),y = Ex,u.(y), so E(S)u = uE(S) for all Borel sets S. Conversely, suppose now that u commutes with all the projections E(S). Then (E(S)u(x),y) = (uE(S)(x),y) = (E(S)(x),u*(y)), so Eu(x),y = Ex,u.(y). Hence, for every f E C(f2), J J dEu(x),y = J J dEx,u.(y) j that is, (cp(f)u(x), y) = (cp(f)(x), u*(y)), so cp(f)u = ucp(f). 0 The next result (which is a special case of Theorem 2.5.5) is one of the most important in single operator theory, and is called the spectral theorem. 2.5.6. Theorem. Let u be a normal operator on a Hilbert space H. Then there is a unique spectral measure E relative to (O'( U ), H) such that u = J z dE, where z is the inclusion map of 0'( u) in C. Proof. Let cp: C(O'(u))  B(H) be the functional calculus at u. By the preceding theorem, there exists a unique spectral measure E relative to (O'(u), H) such that cp(f) = J f dE for all f E C(O'(u)). In particular, u = cp( z) = J z dE. If E' is another spectral measure such that u = J z dE', then J f dE' = J f dE = cp(f) for all f E C( 0'( u)), since 1 and z generate C( 0'( u)). Therefore, E = E'. 0 The spectral measure E in Theorem 2.5.6 is called the resolution of the identity for u. Since f(u) = J f dE for all f E C(O'(u)), we can unambiguously define feu) = J f dE for all f E Bex>(O'(u)). The unital *- homomorphism Bex>(O'(u))  B(H), f  f(u), is called the Borel functional calculus at u. If v E B( H) commutes with both u and u*, then v commutes with f( u) for all f E B ex> ( 0'( U ) ). For in this case v commutes with all polynomials in u and u *, and since 1 and z generate C( 0'( U )) by the Stone-Weierstrass theorem, v commutes with f(u) for all f E C(O'(u)). By Theorem 2.5.5, therefore, v commutes with E(S) for all Borel sets S of 0'( u). It follows that Ex,v.(y) = Ev(x),y for all x, y E H. Hence, if f E Bex>(O'(u)), ((vJ(u))(x),y) = J JdEx,v.(y) = J J dEv(x),y = (f(u)v(x),y). Therefore, v f( u) = f( u )v. Incidentally, if S is a Borel set of u( u), then X s( u) = E( S). 
2. Exercises 73 2.5.7. Theorem. Let u be a normal operator on a Hilbert space H, and suppose that g: C --+ C is a continuous function. Then (g 0 f)( u) = g(f( u)) for all f E Boo( a( u)). Proof. The result is easily seen by first showing it for 9 a polynomial in z and z, and then observing that an arbitrary continuous function g: C --+ C is a uniform limit of such polynomials on the compact disc  = {,,\ Eel 1,,\1 < IIflloo}, using the Stone-Weierstrass theorem applied to C(). 0 We give an application of this to writing a unitary as an exponential. 2.5.8. Theorem. Let u be a unitary operator in B(H), where H is a Hilbert space. Then there exists a hermitian operator v in B(H) such that u = e iv and IIvll < 27r. Proof. The function f: [0, 27r) --+ T, t  e it , is a continuous bijection with Borel measurable inverse g. Since a( u) C T, we can set v = g( u). The operator v is self-adjoint because 9 is real-valued. Moreover, Ilvll < IIglloo < 27r. By Theorem 2.5.7, (f 0 g)(u) = f(g(u)) = f( v) = e iv . But (f 0 g)("\) = ,,\ for all ,,\ E T, so (f 0 g)( u) = u. Therefore, u=e iv . 0 2. Exercises 1. Let A be a Banach algebra such that for all a E A the implication Aa = 0 or aA = 0 =} a = 0 holds. Let L, R be linear mappings from A to itself such that for all a, b E A, L(ab) = L(a)b, R(ab) = aR(b), and R(a)b = aL(b). Show that Land R are necessarily continuous. 2. Let A be a unital C*-algebra. ( a) If a, b are positive elements of A, show that a( ab) C R + . (b) If a is an invertible element of A, show that a = ulal for a unique unitary u of A. Give an example of an element of B(H) for some Hilbert space H that cannot be written as a product of a unitary times a posi ti ve operator. (c) Show that if a E Inv(A), then Iiall = Ila- 1 11 = 1 if and only if a is a uni tary. 
74 2. C*-Algebras and Hilbert Space Operators 3. Let 0 be a locally compact Hausdorff space, and suppose that the C*-algebra Co(O) is generated by a sequence of projections (Pn)=I. Show that the hermitian element h = E=1 Pn/3n generates Co(O). 4. We shall see in the next chapter that all closed ideals in C*-algebras are necessarily self-adjoint. Give an example of an ideal in the C*-algebra C(D) that is not self-adjoint. 5. Let cp: A  B be an isometric linear map between unital C*-algebras A and B such that cp(a*) = cp(a)* (a E A) and cp(l) = 1. Show that cp(A+) C B+. 6. Let A be a unital C*-algebra. (a) If r( a) < 1 and b = (E=o a*na n )1/2, show that b > 1 and Ilbab- 1 11 < 1. (b) For all a E A, show that r(a) = inf IIbab- 1 II = inf Ilebae-bil. bElnv(A) bEA.a 7. Let A be a unital C*-algebra. ( a) If a, b E A, show that the map f: C  A, A  eiAbae-iAb , is differentible and that f'(O) = i(ba - ab). (b) Let X be a closed vector subspace of A which is unitarily invariant in the sense that uXu* C X for all unitaries u of A. Show that ba-ab E X if a E X and b E A. ( c) Deduce that the closed linear span X of the projections in A has the property that a E X and b E A implies that ba - ab EX. 8. Let a be a normal element of a C*-algebra A, and b an element commut- ing with a. Show that b* also commutes with a (Fuglede's theorem). (Hint: Define f( A) = e iAa * be- 1Aa * in A and deduce from Exercise 2.7 that this map is differentiable and f'(O) = i( a* b - ba*). Since e lXa and b commute, f(A) = e2ic(A)be-2ic(A), where c(A) = Re(Aa*). Hence, IIf(A)11 = IIbll, so by Liouville's theorem, f( A) is constant.) In the following exercises H is a Hilbert space: 9. If I is an ideal of B( H), show that it is self-adjoint. 
2. Addenda 75 10. Let u E B(H). ( a) Show that u is a left topological zero divisor in B ( H) if and only if it is not bounded below (cf Exercise 1.11). (b) Define 0- ape u) = {,,\ E C I u - ,,\ is not bounded below}. This set is called the approximate point spectrum of u because ,,\ E 0- ape u) if and only if there is a sequence (x n) of unit vectors of H such that lim n --+ oo lI(u - "\)(xn)1I = o. Show that O'ap(u) is a closed subset of 0-( u) containing 80-( u). (c) Show that u is bounded below if and only if it is left-invertible in B(H). ( d) Show that 0-( u) = 0- ape u) if u is normal. 11. Let u E B(H) be a normal operator with spectral resolution of the identity E. (a) Show that u admits an invariant closed vector subspace other than 0 and H if dim(H) > 1. (b) If,,\ is an isolated point of 0-( u), show that E("\) = ker( u - ,,\) and that ,,\ is an eigenvalue of u. 12. An operator u on H is subnormal if there is a Hilbert space K con- taining H as a closed vector subspace and there exists a normal operator v on K such that H is invariant for v, and u is the restriction of v. We call v a normal extension of u. (a) Show that the unilateral shift is a non-normal subnormal operator. (b) Show that if u is subnormal, then u*u > uu*. (c) A normal extension v E B(K) of a subnormal operator u E B(H) is a minimal normal extension if the only closed vector subspace of K reducing v and containing H is K itself. Show that u admits a minimal normal extension. In the case that v is a minimal normal extension, show that I{ is the closed linear span of all v* n ( x) (n EN, x E H). (d) Show that if v E B(I{) and v' E B(K') are minimal normal extensions of u, then there exists a unitary operator w: I{  I{' such that Vi = wvw* (so there is only one minimal normal extension). 2. Addenda In the following, H is an infinite-dimensional separable Hilbert space. If u is a self-adjoint operator on H, then there exists a self-adjoint diagonalisable operator v and a self-adjoint compact operator w on H, such that u = v + w (Weyl-von Neumann). Similarly, if u is a normal operator on H, there exists a diagonalisable operator v and a compact operator w on H, such that u = v + w (I. D. Berg). An operator u E B(H) is essentially normal if u*u - uu* is a compact operator. If u is the sum of a normal and a compact operator, then ob- viously it is essentially normal. The unilateral shift is essentially normal, 
76 2. C*-Algebras and Hilbert Space Operators but it is not the sum of a normal and a compact operator, since it has non- zero Fredholm index. It turns out that the index is the only obstruction to an essentially normal operator being the sum of a normal and a compact operator. More precisely, for u an essentially normal operator on H, the following conditions are equivalent: (a) u is the sum of a normal operator and a compact operator. (b) u is the sum of a diagonalisable operator and a compact operator. ( c) For all ,\ E C \ a e ( U ), the operator u - ,\ has zero Fredholm index. An operator u is e33entially unitary if u*u -1 and uu* -1 are compact operators. If v is the unilateral shift and u is an essentially unitary operator on H of Fredholm index n, then there exists a compact operator w such that (a) u - w is unitary if n = 0; (b) u - w = v- n if n is negative; (c) u - w = v*n if n is positive. Let u, v be essentially normal operators on H. The following conditions are equivalent: (a) There exists a compact operator w on H such that v - w is unitarily equivalent to u. (b) The esssential spectra of u and v are the same set, K say, and for each ,\ E C \ K the operators u -,\ and v - ,\ have the same Fredholm index. These surprising and elegant results on essentially normal operators are due to L. Brown, R. G. Douglas, and P. Fillmore: We shall see in the next chapter that B(H)/K(H) is a C*-algebra. If 7r: B(H) --+ B(H)/ ]«H) is the quotient homomorphism, then for u E B(H) the image 7r( u) is normal if and only if u is essentially normal, and 7r( u) is unitary if and only if u is essentially unitary. Although the BDF results are expressed purely in terms of single oper- ator theory, the proofs involve C*-algebras and homological algebraic tech- niques. The introduction of the latter into the subject of operator algebras has given a revolutionary impetus to its development. Reference: [BDF]. Let u be a subnormal operator on H. If v is the minimal normal extension of u, then a( v) C a( u) (P. Halmos). Hence, r( u) = II u II. A much deeper result is that u necessarily has an invariant closed vector subspace other than 0 and H (Scott Brown). For the theory of subnormal operators see [Cnw 1]. 
CHAPTER 3 Ideals and Positive Functionals In this chapter we show that every C*-algebra can be realised as a C*-subalgebra of B(H) for some Hilbert space H. This is the Gelfand- Naimark theorem, and it is one of the fundamental results of the theory of C*-algebras. A key step in its proof is the GNS construction which sets up a correspondence between the positive linear functionals and some of the representations of the algebra. This correspondence will be exploited in many situations in the sequel. There are also deep connections between the positive linear functionals and the closed ideals and closed left ideals of the algebra. We also look at hereditary C*-subalgebras. These are a sort of gener- alisation of ideals and are of great importance in the theory. In the final section of this chapter we apply some of the results we have developed so far to an interesting and highly non-trivial class of operators, the Toepli tz operators. 3.1. Ideals in C*-Algebras In this section we prove basic results on ideals and homomorphisms. First, we show the existence of approximate units in C*-algebras. Of course, if a C*-algebra is non-unital, one can simply adjoin a unit, as we have frequently done. This is not always appropriate, however-consider the problem of showing that closed ideals are self-adjoint (this is shown by using approximate units). An approximate unit for a C*-algebra A is an increasing net (U..x)..xEA of positive elements in the closed unit ball of A such that a = lim..x au.,\ for all a E A. Equivalently, a = lim..x U..xa for all a E A. 77 
78 3. Ideals and Positive Functionals 3.1.1. Ezample. Let H be a Hilbert space with an orthonormal basis (en)  1. The C*-algebra K(H) is of course non-unital, since dim(H) = 00. If Pn is the projection onto Cel + . . · + Ce n , then the increasing sequence (Pn) is an approximate unit for K(H). To see this we need only show that U = limnoo Pn u if u E F( H), since F( H) is dense in K( H). Now if u E F(H), there exist Xl,. . . , X m , YI,. . . , Ym in H such that u = E=I Xk 0 Yk. Hence, pnU = E  1 Pn(Xk) 0 Yk. Since limnooPn(x) = X for all x E H, therefore for each k, lim IIPn(xk) 0 Yk - Xk 0 Ykll = lim IIPn(Xk) - xkllllYkll = O. n-.oo n-.oo Hence, limnooPnu = u. Let A be an arbitrary C*-algebra and denote by A the set of all positive elements a in A such that II a II < 1. This set is a poset under the partial order of Asa. In fact, A is also upwards-directed; that is, if a, b E A, then there exists c E A such that a, b < c. We show this: If a E A +, then 1 + a is of course invertible in A, and a(l + a)-l = 1 - (1 + a)-I. We claim a, b E A+ and a < b =} a(l + a)-l < b(l + b)-I. (1) Indeed, if 0 < a < b, then 1 + a < 1 + b implies (1 + a)-I > (1 + b)-I, by Theorem 2.2.5, and therefore 1 - (1 + a)-I < 1 - (1 + b)-I; that is, a(l + a)-I < b(l + b)-I, proving the claim. Observe that if a E A+, then a(l + a)-I belongs to A (use the Gelfand representation applied to the C*-subalgebra generated by 1 and a). Suppose then that a, b are an arbitrary pair of elements of A. Put a' = a(l - a)-I, b' = b(l - b)-l and c = (a' + b')(l + a' + b')-I. Then c E A, and since a' < a' + b', we have a = a'(l + a')-l < c, by (1). Similarly, b < c, and therefore A is upwards-directed, as asserted. 3.1.1. Theorem. Every C*-algebra A admits an approximate unit. In- deed, if A is the upwards-directed set of all a E A+ such that lIall < 1 and u A = A for all A E A, then (u A) AEA is an approximate unit for A (called the canonical approximate unit). Proof. From the remarks preceding this theorem, (UA)AEA is an increasing net of positive elements in the closed unit ball of A. Therefore, we need only show that a = limA uAa for each a E A. Since A linearly spans A, we can reduce to the case where a E A. Suppose then that a E A and that c > o. Let c.p: C*(a)  Co(r!) be the Gelfand representation. If f = <p(a), then K = {w E r! Ilf(w)1 > c} is compact, and therefore by Urysohn's lemma there is a continuous function g: r!  [0,1] of compact support such that g(w) = 1 for all w E K. Choose 6 > 0 such that 6 < 1 and 1 - 6 < c. Then Ilf - 6gfll < c. If Ao = 
3.1. Ideals in C*-Algebras 79 c.p-l( 8g), then Ao E A and /la - u'\oa/l < £. Now suppose that A E A and A > Ao. Then 1 - u,\ < 1 - u'\o, so a(l - u,\)a < a(l - u..\o)a. Hence, Ila-u,\aI1 2 = /1(1- u..\)1/2(1_u,\)1/2a/l 2 < /I(1-u,\)1/2aIl 2 = lIa(l-u..\)all < Ila(l - u'\o)all < 11(1 - u'\o)all < £. This shows that a = lim,\ u..\a. 0 3.1.1. Remark. If a C*-algebra A is separable, then it admits an ap- proximate unit which is a sequence. For in this case there exist finite sets F 1 C F 2 C ... C Fn C ... such that F = U _ 1 Fn is dense in A. Let (U'\)'\EA be any approximate unit for A. If £ > 0, and Fn = {al,...,a m } say, then there exist AI'...' Am E A such that lIaj - aju'\l1 < £ if A > Aj. Choose A E A such that A > AI'...' Am. Then lIa - au..\1/ < £ for all a E Fn and all A > A. Hence, if n is a positive integer and £ = l/n, then there exists An = A E A such that Ila-aAnll < l/n for all a E Fn. Also, we may obviously choose the An such that An < An+l for all n. Consequently, lim n -. oo lIa - aU..\n II = 0, for all a E F, and since F is dense in A, this also holds for all a E A. Therefore, (u'\n )=1 is an approximate unit for A. 3.1.2. Theorem. If L is a closed left ideal in a C*-algebra A, then there is an increasing net (U'\)..\EA of positive elements in the closed unit ball of L such that a = lim..\ au,\ for all a E L. Proof. Set B = LnL*. Since B is a C*-algebra, it admits an approximate unit, (U'\)'\EA say, by Theorem 3.1.1. If a E L, then a*a E B, so 0 = lim,\a*a(l- u,\). Hence, lim,\ lIa - au..\11 2 = lim,\ 11(1 - u,\)a*a(l - u,\)11 < lim,\ Ila*a(l - u,\)11 = 0, and therefore lim..\ Iia - auAII = o. 0 In the preceding proof we worked in the unitisation A of A. We shall frequently do this tacitly. 3.1.3. Theorem. If I is a closed ideal in a C*-algebra A, then I is self- adjoint and therefore a C*-subalgebra of A. If (U'\)'\EA is an approximate uni t for I, then for each a E A Iia + III = lim Iia - uAal1 = lim Iia - au'\ll. ,\ ,\ Proof. By Theorem 3.1.2 there is an increasing net (U'\)'\EA of positive elements in the closed unit ball of I such that a = lim..\ au..\ for all a E I. Hence, a* = lim,\ u,\a*, so a* E I, because all of the elements u,\ belong to I. Therefore, I is self-adjoint. Suppose that (U'\)'\EA is an arbitrary approximate unit of I, that a E A, and that £ > O. There is an element b of I such that lIa + bll < Ila+ III +£/2. Since b = lim,\ u,\b, there exists Ao E A such that lib - u,\bll < £/2 for all A > Ao, and therefore Iia - u,\all < 11(1 - uA)(a + b)11 + lib - u,\bll < Iia + bll + lib - u,\bll < Iia + III + £/2 + £/2. 
80 3. Ideals and Positive Functionals It follows that lIa + III = limA Iia - uAall, and therefore also lIa + III lIa* + III = limA lIa* - uAa*11 = limA Iia - auAII. 0 3.1.2. Remark. Let I be a closed ideal in a C*-algebra A, and J a closed ideal in I. Then J is also an ideal in A. To show this we need only show that ab and ba are in J if a E A and b is a positive element of J (since J is a C*-algebra, J+ linearly spans J). If (U"\)AEA is an approximate unit for I, then b 1 / 2 = limA u A b 1 / 2 because b 1 / 2 E I. Hence, ab = limA aUAbl/2bl/2, so ab E J because b 1 / 2 E J, au A b 1 / 2 E I, and J is an ideal in I. Therefore, a* b E J also, so ba E J, since J is self-adjoint. 3.1.4. Theorem. If I is a closed ideal of a C*-algebra A, then the quotient AI I is a C*-algebra under its usual operations and the quotient norm. Proof. Let (UA)..\EA be a approximate unit for I. If a E A and bEl, then lIa + 111 2 = lim lIa - au A II 2 (by Theorem 3.1.3) A = lim 11(1 - u A )a*a(l - uA)11 A < sp 11(1 - uA)(a*a + b)(l - u A )1I + lirn 11(1 - u A )b(l - u A )1I < lIa*a + bll + lim lib - uAbl1 - A = lIa*a + bll. Therefore, Iia + 111 2 < lIa*a + III. By Lemma 2.1.3, AI I is a C*-algebra. 0 3.1.5. Theorem. If c.p: A  B is an injective *-homomorpmsm between C*-algebras A and B, then c.p is necessarily isometric. Proof. It suffices to show that 1Ic.p(a)112 = Ilall2, that is, 1Ic.p(a*a)11 = Ila*all. Thus, we may suppose that A is abelian (restrict to C(a*a) if necessary), and that B is abelian (replace B by c.p(A)- if required). Moreover, by extending c.p: A  B to cp: A  iJ if necessary, we may further assume that A, B, and c.p are unital. If r is a character on B, then r 0 c.p is one on A. Clearly the map c.p': f2(B)  f2(A), r  r 0 c.p, is continuous. Hence, c.p'(f2(B)) is compact, because f2(A) is compact, and therefore c.p'(f2(B)) is closed in f2(A). If c.p'(f2(B)) =I f2(A), then by Urysohn's lemma there is a non-zero continuous function f: f2(A)  C such that f vanishes on c.p'(f2(B)). By the Gelfand representation, f = a for some element a E A. Hence, for each r E f2( B), r( c.p( a)) = a( r 0 c.p) = o. 
3.1. Ideals in C*-Algebras 81 Therefore, c.p( a) = 0, so a = O. But this implies that f is zero, a contradic- tion. The only way to avoid this is to have c.p'(f2(B)) = f2(A). Hence, for each a E A, II all = lIali oo = sup Ir(a)1 = sup Ir(c.p(a))1 = 1Ic.p(a)lI. TEO(A) TEO(B) Thus, c.p is isometric. o 3.1.6. Theorem. Ifc.p: A -+ B is a *-homomorpmsm between C*-algebras, then c.p(A) is a C*-subalgebra of B. Proof. The map AI ker( c.p) -+ B, a + ker( c.p)  c.p( a), is an injective *-homomorphism between C*-algebras and is therefore iso- metric. Its image is c.p(A), so this space is necessarily complete and therefore closed in B. 0 3.1.7. Theorem. Let B and I be respectively a C*-subalgebra and a closed ideal in a C*-algebra A. Then B + I is a C*-subalgebra of A. Proof. We show only that B + I is complete, because the rest is trivial. Since I is complete we need only prove that the quotient (B + 1)11 is complete. The intersection B n I is a closed ideal in B and the map c.p from B I(B n I) to AI I defined by setting c.p(b + B n I) = b + I (b E B) is a *-homomorphism with range (B + 1)11. By Theorem 3.1.6, (B + 1)11 is complete, because it is a C* -algebra. 0 3.1.3. Remark. The map c.p: B I (B n I) -+ (B + I) I I, b + B n I  b + I, in the preceding proof is in fact clearly a *-isomorphism. We return to the topic of multiplier algebras, because we can now say a little more about them using the results of this section. Suppose that I is a closed ideal in a C*-algebra A. If a E A, define La and Ra in B(I) by setting La(b) = ab and Ra(b) = ba. It is a straightfor- ward exercise to verify that (La, Ra) is a double centraliser on I and that the map c.p: A -+ M(I), a  (La, Ra), is a *-homomorphism. Recall that we identified I as a closed ideal in M(I) by identifying a with (La, Ra) if a E I. Hence, c.p is an extension of the inclusion map I -+ M(I). 
82 3. Ideals and Positive Functionals If 1 1 ,1 2 , . . . , In are sets in A, we define 1 1 1 2 . . . In to be the closed linear span of all products al a2 . . . an, where aj E Ij. If I, J are closed ideals in A, then I n J = I J. The inclusion I J C In J is obvious. To show the reverse inclusion we need only show that if a is a positive element of In J, then a E IJ. Suppose then that a E (I n J)+. Hence, a 1 / 2 E In J. If (UA)AEA is an approximate unit for I, then a = lim A ( uAal/2)al/2, and since u A a 1 / 2 E I for all ,,\ E A, we get a E I J, as required. Let I be a closed ideal I in A. We say I is e33ential in A if aI = 0 :::} a = 0 (equivalently, Ia = 0 :::} a = 0). From the preceding observations it is easy to check that I is essential in A if and only if I n J i= 0 for all non-zero closed ideals J in A. Every C*-algebra I is an essential ideal in its multiplier algebra M(I). 3.1.8. Theorem. Let I be a closed ideal in a C*-algebra A. Then there is a unique *-homomorphism c.p: A  M(I) extending the inclusion I  M(I). Moreover, c.p is injective if I is essential in A. Proof. We have seen above that the inclusion map I  M(I) admits a *-homomorphic extension <p: A  M(I). Suppose that "p: A  M(I) is another such extension. If a E A and bEl, then c.p( a)b = <p( ab) = ab = 1/;( ab) = 1/;( a )b. Hence, (c.p( a) - "p( a))I = 0, so c.p( a) = "p( a), since I is essential in M(I). Thus, c.p = "p. Suppose now that I is essential in A and let a E ker( c.p). Then aI = La(I) = 0, so a = o. Thus, <p is injective. 0 Theorem 3.1.8 tells us that the multiplier algebra M(I) of I is the largest unital C*-algebra containing I as an essential closed ideal. 3.1.2. Ezample. If H is a Hilbert space, then I{(H) is an essential ideal in B(H). For if u is an operator in B(H) such that uK(H) = 0, then for all x E H we have u(x) 0x = u(x 0x) = 0, so u(x) = O. By Theorem 3.1.8, the inclusion map K(H)  M(I«H)) extends uniquely to an injective *-homomorphism <p: B(H)  M(I{(H)). We show that <p is surjective, that is, a *-isomorphism. Suppose that (L, R) E M(K(H)), and fix a unit vector e in H. The linear map u:H  H, x  (L(x 0 e))(e), is bounded, since lIu(x)1I < IIL(x 0 e)11 < IILllllx 0 ell = IILllllxll. If x, y, z E H, then (Lu (x 0 y))( z) = (u( x) 0 y)( z) = (z, y)(L(x 0 e ))( e) = (L(x 0 e))((z,y)e) = (L( x 0 e))( e 0 y)( z). 
3.2. Hereditary C*-Subalgebras 83 Hence, Lu(x 0 y) = L(x 0 e)(e 0 y) = L(x 0 y) for all x, y E H. Therefore, ( 'P ( u) - (L, R) )I< ( H) = 0, so 'P ( u) = (L, R). Thus, we may regard B(H) as the multiplier algebra of I«H). This ex- ample is the motivating one for the use of the multiplier algebra in K-theory. 3.1.3. Ezample. If n is a locally compact Hausdorff space, then it is easy to check that C o (f2) is an essential ideal in the C*-algebra C b (f2). Therefore, by Theorem 3.1.8 there is a unique injective *-homomorphism 'P: Cb(n)  M(C o (f2)) extending the inclusion Co(n)  M(Co(n)). We show that 'P is surjective, that is, a *-isomorphism. To see this, it suffices to show that if 9 E M(Co(n)) is positive, then it is the range of 'P. If (UA)AEA is an approximate unit for Co(n), then for each wEn the net of real numbers (guA(w)) is increasing and bounded above by IIgll, and therefore it converges to a number, h( w) say. The function h:f2  C, w  h(w), is bounded. Moreover, if f E Co(n), then hf = gf, since f = limA fu).. To see that h is continuous, let (WJl)JlEM be a net in f2 converging to a point w. Let 1< be a compact neighbourhood of W in n. To show that h(w) = limJl h(wJl)' we may suppose WJl E 1< for all indices J.L (there exists J-lo such that W Jl E 1< for all indices J-l > J-lo, so, if neccessary, replace the net (w Jl) JlEM by the net (w Jl) JlJlo). Use U rysohn's lemma to choose a function f E Co(n) such that f = 1 on I<. Since fh E Co(n), h(w) = fh(w) = limfh(wJl) = limh(wJl). Jl Jl Therefore, h is continuous, so h E Cb(n). For f an arbitrary function in Co(n) we have 'P(h)f = 'P(hf) = hf = gf, so ('P(h) - g)Co(n) = o. Consequently, 9 = 'P( h). 3.2. Hereditary C*-Subalgebras This section introduces a new class of C*-subalgebras, namely the hereditary ones. These are particularly well-behaved, especially with re- spect to extending positive linear functionals, an important topic to be taken up in the next section. We illustrate the nice behaviour of hereditary C*-subalgebras in connection with the concept of simplicity of an algebra. A C*-subalgebra B of a C*-algebra A is said to be hereditary if for a E A+ and b E B+ the inequality a < b implies a E B. Obviously, 0 and A are hereditary C*-subalgebras of A, and any intersection of hereditary C*-subalgebras is one also. The hereditary C*-subalgebra generated by a subset S of A is the smallest hereditary C*-subalgebra of A containing S. 
84 3. Ideals and Positive Functionals 3.2.1. Ezample. If p is a projection in a C*-algebra A, the C*-subalgebra pAp is hereditary. For, assuming 0 < b < pap, then 0 < (1 - p)b(l - p) < (1 - p)pap(l - p) = 0, so (1 - p)b(l - p) = o. Hence, IIb 1 / 2 (1 _ p)1I 2 = 11(1 - p)b(l - p)1I = 0, so b(l - p) = o. Therefore, b = pbp E pAp. The correspondence between hereditary C*-subalgebras and closed left ideals in the following theorem is very useful. 3.2.1. Theorem. Let A be a C*-algebra. (1) If L is a closed left ideal in A, then LnL* is a hereditary C*-subalgebra of A. The map L  L n L * is a bijection from the set of closed left ideals of A onto the set of hereditary C*-subalgebras of A. (2) If L 1 , L 2 are closed left ideals of A, then L 1 C L 2 if and only if L 1 nL c L 2 n L. (3) If B is a hereditary C*-subalgebra of A, then the set L(B) = {a E A I a*a E B} is the unique closed left ideal of A corresponding to B. Proof. If L is a closed left ideal of A, then clearly B = L n L* is a C*-subalgebra of A. Suppose that a E A+ and b E B+ and a < b. By Theorem 3.1.2 there is an increasing net (U'\)'\EA in the closed unit ball of L+ such that lim,\ bu,\ = b. Now 0 < (1- u,\)a(l - u,\) < (1- u,\)b(l- u,\), so lIa 1 / 2 -a 1 / 2 u'\1I2 = II(l-u,\)a(l-u'\)11 < lI(l-u,\)b(l-u'\)11 < IIb-bu'\lI. Hence, a 1 / 2 = lim,\ a 1 / 2 u,\, so a 1 / 2 E L, since u,\ E L (,,\ E A). Therefore, a E B, so B is hereditary in A. Suppose now that L 1 , L 2 are closed left ideals of A. It is evident that L 1 C L 2 => L 1 n L C L 2 n L. To show the reverse implication, suppose that L 1 n Lr c L 2 n L; and let (u.\),\EA be an approximate unit for L 1 n L, and a ELI. Then lim,\ Iia - au,\11 2 = lim,\ 11(1- u,\)a*a(l- u,\)11 < lim,\ lIa*a(l - u,\)11 = 0, since a*a E L 1 n L. It follows that lim,\ au.\ = a. Therefore, a E L 2 , since u,\ E L 1 n L c L 2 . This proves Condition (2). Now let B be a hereditary C*-subalgebra of A and let L = L(B). If a,b E L, (a+b)*(a+b) < (a+b)*(a+b)+(a-b)*(a-b) = 2a*a+2b*b E B, so a + bEL. If a E A and bEL, then (ab)*(ab) = b*a*ab < Ilall 2 b*b E B, so ab E L. Similarly, L is closed under scalar multiplication. Thus, L is a left ideal, and it is obviously closed, since B is closed. If b E B, then b* b E B, so bEL. Hence, B C L n L *. If 0 < bEL n L *, then b 2 E B, so b E B, and therefore L n L* C B. Hence, L n L* = B. This proves Condi tion (3), and Condition (1) follows directly. 0 3.2.2. Theorem. Let B be a C*-subalgebra of a C*-algebra A. Then B is hereditary in A if and only if bab' E B for all b, b' E B and a E A. 
3.2. Hereditary C*-Subalgebras 85 Proof. If B is hereditary, then by Theorem 3.2.1 B = L n L * for some closed left ideal L of A. Hence, if b, b' E B and a E A, we have b( ab') E L and b'*(a*b*) E L, so bab' E B. Conversely, suppose B has the property that bab' E B for all b, b' E B and a E A. If (U'\)'\EA is an approximate unit for B and a E A+, b E B+, and a < b, then 0 < (1 - u.\)a(l - u,\) < (1 - u.\)b(l - u.\), and therefore lIa 1 / 2 - a 1 / 2 u'\l1 < IIb 1 / 2 - b 1 / 2 u,\ll. Since b 1 / 2 = lim.\ b 1 / 2 u,\, therefore, a 1 / 2 = lim,\ a 1 / 2 u,\, so a = lim..\ u,\au,\ E B. Thus, B is hereditary. 0 The following corollary is obvious. 3.2.3. Corollary. Every closed ideal of a C*-algebra is a hereditary C*-subalgebra. 3.2.4. Corollary. If A is a C*-algebra and a E A+, then (aAa)- is the hereditary C*-subalgebra of A generated bya. Proof. The only thing we show is that a E (aAa) -, because the rest is routine. If (U'\)'\EA is an approximate unit for A, then a 2 = lim,\ au,\a, so a 2 E (aAa) -. Since (aAa) - is a C* -algebra, a = Vdi E (aAa) - also. 0 In the separable case, every hereditary C*-subalgebra is of the form in the preceding corollary: 3.2.5. Theorem. Suppose that B is a separable hereditary C*-subalgebra of a C*-algebra A. Then there is a positive element a E B such that B = (aAa) - . Proof. Since B is a separable C*-algebra, it admits a sequential ap- proximate unit, (un)=l say (cf. Remark 3.1.1). Set a = E=l u n /2n. Then a E B+, so B contains (aAa) -. Since u n /2n < a, and (aAa)- is hereditary by Corollary 3.2.4, therefore Un E (aAa)-. If b E B, then b = lim n -. oo Un bUn, and unbu n E (aAa) -, so b E (aAa) -. This shows that B = (aAa) - . 0 If the separability condition is dropped in Theorem 3.2.5, the result may fail. To see this let H be a Hilbert space, and suppose that U is a positive element of B(H) such that K(H) = (uB(H)u)-. If x E H, then x 0 x = limn-.oouvnu for a sequence (v n ) in B(H), and therefore x is in the closure of the range of u. This shows that H = (u(H))-, and therefore H is separable, since the range of a compact operator is separable (cf. Remark 1.4.1). Thus, if H is a non-separable Hilbert space, then the hereditary C*-subalgebra K( H) of B( H) is not of the form (uB( H)u)- for any u E B(H)+. 
86 3. Ideals and Positive Functionals 3.2.6. Theorem. Suppose that B is a hereditary C*-subalgebra of a unital C*-algebra A, and let a E A+. If for each c > 0 there exists b E B+ such that a < b + c, then a E B. Proof. Let £ > O. By the hypothesis there exists b E B+ such that a < b + €2, so a < (be + € t. Hence, (be + € )-1 a(b e + €) -1 < 1, and there- fore lI(b + £ )-1 a(b + £)- II < 1. Using the fact that 1 - b(b + c )-1 = £(b + £)-1 , we get lIa 1 / 2 _ a1/2bE(b + £)-111 2 = £21Ia1/2(b + £)-111 2 = £2 II (bE + £)-1a(b + £)-1" < £2. Hence, a 1 / 2 = lim a 1 / 2 bEe bE + £) -1 , E-+O and therefore also a 1 / 2 = lim (bE + £)-1 ba1/2, -+O by taking adjoints. Thus, a = lim (bE + £ ) -1 bE ab E (bE + £) -1 . E-+O Now bE(b E + £)-1 E B, and therefore (bE + £)-1bEabE(bE + £)-1 E B, since B is hereditary in A. It follows that a E B. 0 We briefly indicate the connection between the ideal structure of a C*-algebra and its hereditary C*-subalgebras in the following results, but we shall defer to Chapter 5 a fuller consideration of this matter. 3.2.7. Theorem. Let B be a hereditary C*-subalgebra of a C*-algebra A, and let J be a closed ideal of B. Then there exists a closed ideal I of A such that J = B n I. Proof. Let I = AJ A. Then I is a closed ideal of A. Since J is a C*-algebra, J = J3, and since B is hereditary in A, we have B n I = BIB (both of these assertions follow easily from the existence of approximate units). Therefore, BnI = BIB = B(AJ A)B = BAJ 3 AB C BJB, because B AJ and JAB are contained in B by Theorem 3.2.2. Since B J B = J, be- cause J is a closed ideal in B, we have B n I C J, and the reverse inclusion is obvious, so B n I = J. 0 A C*-algebra A is said to be simple if 0 and A are its only closedjdeals. These algebras are (loosely) thought of as the building blocks of the theory of C*-algebras, and it is important to compile a large stock of examples. We shall be presenting some as we proceed, but for the present we content ourselves with one class of examples: 
3.3. Positive Linear Functionals 87 3.2.2. Eample. If H is a Hilbert space, then the C*-algebra K(H) is simple. For if I is a closed non-zero ideal of K(H), it is also an ideal of B(H) (cf. Remark 3.1.2), so I contains the ideal F(H) by Theorem 2.4.7, and therefore I = I{(H). It is not true that C*-subalgebras of simple C*-algebras are necessarily simple. For instance, if p, q are finite-rank non-zero projections on a Hilbert space H such that pq = 0, then A = Cp+Cq is a non-simple C*-subalgebra of the simple C*-algebra I{(H) (the closed ideal Ap = Cp of A is non- trivial ). 3.2.8. Theorem. Every hereditary C*-subalgebra of a simple C*-algebra is simple. Proof. Let B be a hereditary C*-subalgebra of a simple C*-algebra A. If J is a closed ideal of B, then J = B n I for some closed ideal I of A by Theorem 3.2.7. Simplicity of A implies that I = 0 or A, and therefore J = 0 or B. 0 3.3. Positive Linear Functionals For abelian C*-algebras we were able completely to determine the structure of the algebra in terms of the character space, that is, in terms of the one-dimensional representations. For the non-abelian case this is quite inadequate, and we have to look at representations of arbitrary dimension. There is a deep inter-relationship between the representations and the pos- itive linear functionals of a C*-algebra. Representations will be defined and some aspects of this inter-relationship investigated in the next section. In this section we establish the basic properties of positive linear functionals. If <p: A  B is a linear map between C*-algebras, it is said to be positive if <p( A +) C <p( B+). In this case <p( Asa) C B sa, and the restriction map c.p: Asa  B sa is increasing. Every *-homomorphism is positive. 3.3.1. Eample. Let A = C(T) and let m be normalised arc length measure on T. Then the linear functional G(T)  c, f  J f dm, is positive (and not a homomorphism). 3.3.2. Eample. Let A = Mn(C). The linear functional n tr: A  C, (Aij)  L Aii, i=l 
88 3. Ideals and Positive Functionals is positive. It is called the trace. Observe that there are no non-zero *-homomorphisms from Mn(C) to C if n > 1. Let A be a C*-algebra and r a positive linear functional on A. Then the function A 2 -+ C, (a, b)  r(b*a), is a positive sesquilinear form on A. Hence, r(b*a) = r(a*b)- and Ir(b*a)1 < r(a*a)1/2r(b*b)1/2. Moreover, the function a  r(a*a)1/2 is a semi-norm on A. Suppose now only that r is a linear functional on A and that M is an element of R+ such that Ir(a)1 < M for all positive elements of te closed unit ball of A. Then r is bounded with norm Ilrll < 4M. We show this: First suppose that a is a hermitian element of A such that Iiall < 1. Then a+, a- are positive elements of the closed unit ball of A, and therefore Ir(a)1 = Ir(a+)-r(a-)I < 2M. Now suppose that a is an arbitrary element of the closed unit ball of A, so a = b+ic where b, c are its real and imaginary parts, and IIbll, lIell < 1. Then Ir(a)1 = Ir(b) + ir(c)1 < 4M. 3.3.1. Theorem. If r is a positive linear functional on a C*-algebra A, then it is bounded. Proof. If r is not bounded, then by the preceding remarks sUPaES r( a) = +00, where S is the set of all positive elements of A of norm not greater then 1. Hence, there is a sequence (an) in S such that 2 n < r( an) for all n E N. Set a = E':=oa n /2 n , so a E A+. Now 1 < r(a n /2n) and therefore N < E:Ol r(a n /2n) = r(E:ol a n /2 n ) < rea). Hence, rea) is an upper bound for the set N, which is impossible. This shows that r is bounded. 0 3.3.2. Theorem. If r is a positive linear functional on a C*-algebra A, then r(a*) = r(a)- and Ir(a)12 < IIrllr(a*a) for all a E A. Proof. Let (UA)AEA be an approximate unit for A. Then r( a*) = lim r( a*u A ) = lim r( uAa)- = r( a)-. A A Also, Ir(a)12 = limA Ir(u A a)12 < SUPA r(u)r(a*a) < Ilrllr(a*a). 0 3.3.3. Theorem. Let r be a bounded linear functional on a C*-algebra A. The following conditions are equivalent: (1) r is positive. (2) For each approximate unit (UA)AEA of A, Ilrll = limA r(u A ). (3) For some approximate unit (UA)AEA of A, IIrli = limA r( u A ). 
3.3. Positive Linear Functionals 89 Proof. We may suppose that IITII = 1. First we show the implication (1) =} (2) holds. Suppose that T is positive, and let (U.x)..\EA be an ap- proximate unit of A. Then (T( u..\) .x).xEA is an increasing net in R, so it converges to its supremum, which is obviously not greater than 1. Thus, lim.x T(U.x) < 1. Now suppose that a E A and lIall < 1. Then IT(u.xa)12 < T(ui)T(a*a) < T(u.x)T(a*a) < lim.x T(U.x), so IT(a)12 < lim.x T(U.x). Hence, 1 < lim.x T( u.x). Therefore, 1 = lim.x T( u.x), so (1) =} (2). That (2) => (3) is obvious. Now we show that (3) => (1). Suppose that (U.x).xEA is an approximate unit such that 1 = lim.x r( u.x). Let a be a self-adjoint element of A such that lIall < 1 and write T(a) = a+i/3 where a,/3 are real numbers. To show that T( a) E R, we may suppose that /3 < o. If n is a positive integer, then Iia - inu.x112 = II(a + inu.x)(a - inu.x)1I = IIa 2 + n 2 ui - in( au.x - u.xa)1I < 1 + n 2 + nllau.x - u.xall, so IT( a - inu).) 1 2 < 1 + n 2 + nllau.x - u.xall. However, lim.x T(a - inu.x) = T(a) - in, and lim.x au.x - U.xa = 0, so in the limi t as ,,\ --+ 00 we get la + i/3 - inl 2 < 1 + n 2 . The left-hand side of this inequality is a 2 + /32 - 2n /3 + n 2 , so if we cancel and rearrange we get -2n/3 < 1 - /32 - a 2 . Since /3 is not positive and this inequality holds for all positive integers n, (3 must be zero. Therefore, r( a) is real if a is hermitian. Now suppose that a is positive and lIall < 1. Then U.x - a is hermitian and Ilu.x-all < 1, so T(U.x-a) < 1. But then 1-T(a) = lim.x T(u..\-a) < 1, and therefore T(a) > o. Thus, T is positive and we have shown (3) => (1).0 3.3.4. Corollary. 1fT is a bounded linear functional on a unital C*-algebra, then T is positive if and only if T(l) = IITII. Proof. The sequence which is constantly 1 is an approximate unit for the C*-algebra. Apply Theorem 3.3.3. 0 3.3.5. Corollary. If T, T' are positive linear functionals on a C*-algebra, then liT + T'II ____ IITII + IIT'II. Proof. If (U.x).xEA is an approximate unit for the algebra, then liT + T' II = lim.x(T + T')(U.x) = lim.x T(U.x) + lim.x T'(U.x) = IITII + IIT'II. 0 A state on a C*-algebra A is a positive linear functional on A of norm one. We denote by S(A) the set of states of A. 
90 3. Ideals and Positive Functionals 3.3.6. Theorem. H a is a normal element of a non-zero C*-algebra A, then there is a state r of A such that lIall = Ir(a)l. Proof. We may assume that a i= o. Let B be the C*-algebra generated by 1 and a in A. Since B is abelian and a is continuous on the compact space f2( B), there is a character r2 on B such that II a II = II a II 00 = I r2 ( a ) I..: By the Hahn-Banach theorem, there is a bounded linear functional rl on A extending r2 and preserving the norm, so Ilrlll = 1. Since rl (1) = r2(1) = 1, rl is positive by Corollary 3.3.4. If r denotes the restriction of rl to A, then r is a positive linear functional on A such that lIall = Ir( a) I. Hence, IIrllllall > Ir(a)1 = lIall, so IIrll > 1, and the reverse inequality is obvious. Therefore, r is a state of A. 0 3.3.7. Theorem. Suppose that r is a positive linear functional on a C*-algebra A. (1) For each a E A, r(a*a) = 0 if and only if r(ba) = 0 for all b E A. (2) The inequality r(b*a*ab) < Ila*allr(b*b) holds for all a, b E A. Proof. Condition (1) follows from the Cauchy-Schwarz inequality. To show Condition (2), we may suppose, using Condition (1), that r(b*b) > o. The function p: A  C, c  r(b*cb)/r(b*b), is positive and linear, so if (UA)AEA is any approximate unit for A, then IIpll = limp(u A ) = limr(b*uAb)/r(b*b) = r(b*b)/r(b*b) = 1. A A Hence, p(a*a) < Ila*all, and therefore r(b*a*ab) < Ila*allr(b*b). 0 We turn now to the problem of extending positive linear functionals. 3.3.8. Theorem. Let B be a C*-subalgebra of a C*-algebra A, and sup- pose that r is a positive linear functional on B. Then there is a positive linear functional r' on A extending r such that II r' II = II r II. Proof. Suppose first that A = iJ. Define a linear functional r' on A by setting r'(b + A) = r(b) + Allrll (b E B, A E C). Let (UA)AEA be an approximate unit for B. By Theorem 3.3.3, Ilrll = limA r( u A ). Now suppose that b E Band J.L E C. Then Ir'(b + J.L)I = I limA r(bu A ) + J.L limA r( uA)1 = I limA r((b + J.L)(uA))1 < SUPA IIrllll(b + J.L)uAII < Ilrllllb + J.LII, since IluAIl < 1. Hence, IIr'lI < IIrll, and the reverse inequality is obvious. Thus, IIr'll = IIrll = r'(l), so r' is positive by Corollary 3.3.4. This proves the theorem in the case A = B. 
3.3. Positive Linear Functionals 91 Now suppose that I A is an arbitrary _C*-algepra containing B as a C*-subalgebra. Replacing B and A by B and A if necessary, we may suppose that A has a unit 1 which lies in B. By the Hahn-Banach theorem, there is a functional r' E A * extending r and of the same norm. Since r'(l) = r(l) = IIrll = IIr'lI, it follows as before from Corollary 3.3.4 that r' is positive. 0 In the case of hereditary C*-subalgebras, we can strengthen the above result-we can even write down an "expression" for r': 3.3.9. Theorem. Let B be a hereditary C*-subalgebra of a C*-algebra A. If r is a positive linear functional on B, then there is a unique positive linear functional r' on A extending r and preserving the norm. Moreover, if(u).)).EA is an approximate unit for B, then r' (a) = lim r( u).au).) ). (a E A). Proof. Of course we already have existence, so we only prove uniqueness. Let r' be a positive linear functional on A extending r and preserving the norm. We may in turn extend r' in a norm-preserving fashion to a positive functional (also denoted r') on A. Let (U).)).EA be an approximate unit for B. Then lim). r(u).) = Ilrll = Ilr'll = r'(l), so lim). r'(l - u).) = o. Thus, for any element a E A, Ir'(a) - r(u).au.x)1 < Ir'(a - u).a)1 + Ir'(u).a - u).au).)1 < r'((l - u).)2)1/2r'(a*a)1/2 + r'(a*ua)1/2r'((1 - u).)2)1/2 < (r'(l- u).))1/2r'(a*a)1/2 + r'(a*a)1/2(r'(1- u).))1/2. Since lim). r'(l- u).) = 0, these inequalities imply lim). r(u).au).) = r'(a).D Let f! be a compact Hausdorff space and denote by C(f!, R) the real Banach space of all real-valued continuous functions on f!. The operations on C(f!, R) are the pointwise-defined ones and the norm is the sup-norm. The Riesz-Kakutani theorem asserts that if r: C(f!, R) -+ R is a bounded real-linear functional, then there is a unique real measure J-l E M(f!) such that r(f) = J f dJ-l for all f E C(f!, R). Moreover, 11J-l11 = Ilrll, and J-l is positive if and only if r is positive; that is, ref) > 0 for all f E C(f!, R) such that f > o. The Jordan decomposition for a real measure J-l E M(f!) asserts that there are positive measures J-l+, J-l- E M(f!) such that J-l = J-l+ - J-l- and IIJ-lil = IIJ-l+ II + IIJ-l-II. We translate this via the Riesz-Kakutani theorem into a statement about linear functionals: If r: C(f!, R) -+ R is a bounded real- linear functional, then there exist positive bounded real-linear functionals r+,r_:C(f!,R) -+ R such that r = r+ - r_ and IIrll = IIr+11 + IIr-li. We are now going to prove an analogue of this result for C*-algebras. 
92 3. Ideals and Positive Functionals Let A be a C*-algebra. If r is a bounded linear functional on A, then IIrll = sup IRe(r(a))I. lIalll (1) For if a E A and II a II < 1, then there is a number ,\ E T such that ,,\ r ( a) E R, so Ir( a)1 = IRe( r("\a)) I < IIrl!, which implies Eq. (1). If r E A*, we define r* E A* by setting r*(a) = r(a*)- for all a E A. Note that r** = r, IIr* II = IIrll, and the map r  r* is conjugate-linear. We say a functional rEA * is self-adjoint if r = r*. For any bounded linear functional r on A, there are unique self-adjoint bounded linear func- tionals rl and r2 on A such that r = rl + ir2 (take rl = (r + r*)/2 and r2 = (r - r*)/2i). The condition r = r* is equivalent to r(Asa) C R, and therefore if r is self-adjoint, the restriction r': Asa -+ R of r is a bounded real-linear functional. Moreover, Ilrll = Ilr'II; that is, IIrll = sup Ir( a )1. aEA. a lIalll For if a E A, we have Re(r(a)) = r(Re(a)), so IIrli = sup I Re( r( a))1 < sup Ir(b)1 < Ilrll. lIalll bEA.a IIblll We denote by A: a the set of self-adjoint functionals in A *, and by A+ the set of positive functionals in A * . We adopt some temporary notation for the proof of the next theorem: If X is a real-linear Banach space, we denote its dual (over R) by Xb. The space Asa is a real-linear Banach space and it is an easy exercise to verify that .£4: a is a real-linear vector subspace of A * and that the map A: a -+ Aa' r  r', is an isometric real-linear isomorphism. We shall use these observations in the proof of the following result. 3.3.10. Theorem (Jordan Decomposition). Let r be a self-adjoint bounded linear functional on a C*-algebra A. Then there exist positive linearfunctionalsr+,T_ on A such thatr = r+-r_ andllrll = Ilr+II+llr-ll. Proof. Let n denote the set of all r E A+ such that IIrll < 1. Then n is weak* closed in the unit ball of A*, so by the Banach-Alaoglu theorem n is a (Hausdorff) weak* compact space. If a E Asa, def¥1e B(a) E C(n, R) by setting B( a)( r) = r( a). The map B: Asa -+ C(n, R), a  B(a), 
3.4. The Gelfand-Naimark Representation 93 is clearly real-linear, and also order-preserving; that is, if a is a positive ele- ment of A, then (}(a) > 0 on f!. Moreover, (} is isometric by Theorem 3.3.6. If T E A: a , then T' E Aa. By the Hahn-Banach theorem, there exists a real-linear functional p E C(f!, R)b such that p(} = T' and IIpil = IIT'II. By the remarks preceding this theorem, there exist positive functionals p+, p- E C(f!, R)b such that p = p+ - p- and IIpll = IIp+ II + lip_II. Set T = p+ o(} and T = p_ o(}. Clearly, T, T E Aa. We denote the corresponding self-adjoint functionals in A: a by T + and T _. Then T = T + - T _, and since IITII = IIT'II = IIpll = IIp+1I + lip-II > IITII + IITII = IIT+II + liT_II > IITII, we have "T" = "T +" + "T _ ". Clearly, T +, T _ E Ai-. 0 One can show that the functionals T + and T _ in the preceding theorem are unique ([Ped, Theorem 3.2.5]), but we shall have no need of this. 3.4. The Gelfand-Naimark Representation In this section we introduce the important GNS construction and prove that every C*-algebra can be regarded as a C*-subalgebra of B(H) for some Hilbert space H. It is partly due to this concrete realisation of the C*-algebras that their theory is so accessible in comparison with more gen- eral Banach algebras. A representation of a C*-algebra A is a pair (H, <p) where H is a Hilbert space and <p: A -+ B(H) is a *-homomorphism. We say (H, <p) is faithful if <p is injective. If (H,\, <p ,\) '\EA is a family of representations of A, their direct sum is the representation (H,cp) got by setting H = ffi,\H,\, and cp(a)((x,\),\) = (cp,\(a)(x,\)),\ for all a E A and all (x,\),\ E H. It is readily verified that (H, <p) is indeed a representation of A. If for each non-zero element a E A there is an index ,,\ such that cp,\ ( a) :F 0, then (H, <p) is faithful. Recall now that if H is an inner product space (that is, a pre-Hilbert space), then there is a unique inner product on the Banach space completion II of H extending the inner product of H and having as its associated norm the norm of II. We call if endowed with this inner product the Hilbert space completion of H. With each positive linear functional, there is associated a represent- ation. Suppose that T is a positive linear functional on a C*-algebra A. Set ting NT = {a E A I T(a*a) = OJ, it is easy to check (using Theorem 3.3.7) that NT is a closed left ideal of A and that the map (A/N T )2 -+ C, (a + NT, b + NT) ...... T(b*a), 
94 3. Ideals and Positive Functionals is a well-defined inner product on A/N r . We denote by Hr the Hilbert completion of A/N r . If a E A, define an operator cp( a) E B( A/ N r) by setting cp(a)(b + N r ) = ab + N r . The inequality I/cp(a)1I < I/al/ holds since we have I/cp(a)(b + N r )1/2 = r(b*a*ab) < l/aIl 2 r(b*b) = lIall 2 11b + N r ll 2 (the latter inequality is given by Theorem 3.3.7). The operator cp( a) has a unique extension to a bounded operator CPr(a) on Hr. The map CPr: A  B(Hr), a  CPr(a), is a *-homomorphism (this is an easy exercise). The representation (Hr,CPr) of A is the Gelfand-Naimark-Segalrepre- sentation (or G N S repres entation) associated to r. If A is non-zero, we define its universal representation to be the direct sum of all the representations (Hr,CPr), where r ranges over S(A). 3.4.1. Theorem (Gelfand-Naimark). If A is a C*-algebra, then it has a faithful representation. Specifically, its universal representation is faithful. Proof. Let (H, cp) be the universal representation of A and suppose that a is an element of A such that cp( a) = O. By Theorem 3.3.6 there is a state r on A such that IIa*all = r(a*a). Hence, if b = (a*a)1/4, then l/al/ 2 = r(a*a) = r(b 4 ) = I/CPr(b)(b + N r )1I2 = 0 (since CPr(b 4 ) = CPr(a*a) = 0, so CPr(b) = 0). Hence, a = 0, and cP is injective. 0 The Gelfand-Naimark theorem is one of those results that are used all of the time. For the present we give just two applications. The first application is to matrix algebras. If A is an algebra, Mn(A) denotes the algebra of all n x n matrices with entries in A. (The operations are defined just as for scalar matrices.) If A is a *-algebra, so is M n (A), where the involution is given by (aij ):,j = (aji )i,j. If cp: A  B is a *-homomorphism between *-algebras, its inflation is the *- homomorphism (also denoted cp) cp: Mn(A)  Mn(B), (aij)  (cp(aiJ)). If H is a Hilbert space, we write H(n) for the orthogonal sum of n copies of H. If u E Mn(B(H)), we define cp(u) E B(H(n») by setting n n CP(U)(X1'... ,x n ) = (2: U1j(Xj),..., 2: Unj(Xj)), j=l j=l 
3.4. The Gelfand-Naimark Representation 95 for all (Xl'. . . , X n ) E H(n). It is readily verified that the map c.p: Mn(B(H))  B(H(n»), U  c.p( u), is a *-isomorphism. We call c.p the canonical *-isomorphism of Mn(B(H)) onto B(H(n»), and use it to identify these two algebras. If v is an operator in B(H(n») such that v = c.p(u) where u E Mn(B(H)), we call u the operator matrix of v. We define a norm on Mn(B(H)) making it a C*-algebra by setting lIuli = 1Ic.p(u)lI. The following inequalities for u E Mn(B(H)) are easy to verify and are often useful: n lIuijll < Ilull < L lIuklli k,l=l (i,j = l,...,n). 3.4.2. Theorem. If A is a C*-algebra, then there is a unique norm on Mn(A) making it a C*-algebra. Proof. Let the pair (H, c.p) be the universal representation of A, so the *-homomorphism c.p: Mn(A)  Mn(B(H)) is injective. We define a norm on Mn(A) making it a C*-algebra by setting lIall = 1Ic.p(a)II for a E Mn(A) (completeness can be easily checked using the inequalities preceding this theorem). Uniqueness is given by Corollary 2.1.2. 0 3.4.1. Remark. If A is a C*-algebra and a E Mn(A), then n lIaij II < Iiall < L lIadl k,l=l (i,j = 1,...,n). These inequalities follow from the corresponding inequalities in Mn(B(H)). Matrix algebras playa fundamental role in the K-theory of C*-algebras. The idea is to study not just the algebra A but simultaneously all of the matrix algebras M n ( A) over A also. Whereas it seems that the only way known of showing that matrix algebras over general C* -algebras are themselves normable as C* -algebras is to use the Gelfand-Naimark representation, for our second application of this representation alternative proofs exist, but the proof given here has the virtue of being very "natural." 3.4.3. Theorem. Let a be a self-adjoint element of a C*-algebra A. Then a E A + if and only if T( a) > 0 for all positive linear functionals T on A. Proof. The forward implication is plain. Suppose conversely that T( a) > o for all positive linear functionals T on A. Let (H, c.p) be the universal representation of A, and let X E H. Then the linear functional T: A  C, b  (c.p(b)(x), x), 
96 3. Ideals and Positive Functionals is positive, so rea) > 0; that is, (<p(a)(x),x) > O. Since this is true for all x E H, and since <p( a) is self-adjoint, therefore <pc a) is a positive operator on H. Hence, <pea) E <p(A)+, so a E A+, because the map <p: A  <peA) is a *-isomorphism. 0 3.5. Toeplitz Operators In this section we apply some of the theory we have developed so far. Our objective is to develop some aspects of the theory of Toeplitz operators. The literature on these operators is vast, and their theory is deep. We shall for the most part confine ourselves to Toeplitz operators with continuous symbol, since their theory is directly accessible to C*-algebraic methods. Apart from their great intrinsic interest, we are concerned with these op- erators for another reason-the C*-algebra that they generate (called the Toeplitz algebra) will play an indispensible role in the proof of Bott Period- icity in K-theory that we present in Chapter 7. With this application in mind, we shall develop a number of the properties of this algebra. Endow the circle group T with its normalised arc length measure (= Haar measure), denoted by dA, and write LP(T) for LP(T, dA). Thus, if f E L 1 (T), then f f(A) dA = 211f fo21f f(e it ) dt. For each integer n, the function en: T  T, A  An, is of course continuous. We denote by r the linear span of the en (n E Z). The elements of r are called trigonometric polynomia13. The set r is a *-subalgebra of G(T), and as we have observed already (in Example 2.4.2) it follows from the Stone-Weierstrass theorem that r is norm-dense in G(T). Since G(T) is LP- norm dense in LP(T) for 1 < p < +00, therefore r is also LP- norm dense in LP(T). Hence, (cn)nEZ is an orthonormal basis of the Hilbert space L 2 (T). If fELl (T), recall that the nth Fourier coefficient of f is defined to be j( n) = f f()..)>.. n d)", and that the function f:z  C, n  fen), is the Fourier tran3form of f. If j = 0, then f = 0 a.e. For in this case, fg(A)f(A)dA = 0 for all 9 E r, and therefore for all 9 E G(T) by sup-norm density of r in G(T). Hence, the measure f dA is zero, so f = 0 a.e. For p E [1, +00] set HP = {f E-LP(T) I j(n) = 0 (n < O)}. This is an LP-norm closed vector subspace of LP(T), called a Hardy 3pace. 
3.5. Toeplitz Operators 97 We write r + for the linear span of the functions Cn (n EN), and call the elements of r + the analytic trigonometric polynomials. The set r + is L 2 -norm dense in H 2 and (cn)nEN is an orthonormal basis {or H 2 . It is an easy exercise to verify that for <p E LOO(T) the inclusion <pH2 C H 2 holds if and only if <p E Hoo, and to show from this that HOO is a subalgebra of LOO(T). The Hardy spaces have an interpretation in terms of analytic functions on the open unit disc in the plane satisfying certain growth conditions approaching the boundary (see Exercise 3.10). This explains the "analytic- type" behaviour displayed in the following result. 3.5.1. Lemma. If f, f E HI, then there exists a scalar a E C such that f = a a.e. Proof. Suppose first that f = f a.e. Set a = J f(A) dA, and observe that a = f M dA = f f(A ) dA = a. If n ::; 0, then (f - acor(n) = J(f(A) - a)cn(A) dA = f f(A)cn(A) dA - a f €n(A) dA = 0, and hence, also, (f - a€o)"( n) = 0 if n > 0, so f - aco has zero Fourier transform, and therefore f = a a.e. If we now suppose only that f, f E HI, then Re(f) and Im(f) are in HI, so by what we have just shown, these functions are constant a.e., and therefore f is constant a.e. 0 Recall from Example 2.5.1 that if <p E LOO(T), then the multiplication operator Mcp with symbol <p is defined by Mcp(f) = <pf (f E L 2 (T)); that M", E B(L 2 (T)); and that the map LOO(T)  B(L 2 (T)), <p...... M"" is an isometric *-homomorphism. In this context M", is called a Laurent operator. Set v = Ml and observe that v is the bilateral shift on the basis (cn)nEZ, since V(€n) = €n+l for all n E Z. The restriction u of v to H 2 is the unilateral shift on the basis (€n)nEN of H 2 . We are now going to characterise the invariant subspaces of v, and for this we determine the comffiutant of v, that is, the set of operators commuting with v. 3.5.2. Theorem. ffw is a bounded operator on L 2 (T), then w commutes with v if and only if w = M", for some <p E LOO(T). Proof. We show the forward implication only because the reverse is clear. Suppose then wv = vw. If"p E r, then Mt/J is in the linear span of all the powers v n (n E Z), so Mt/J commutes with w. If now "p is an arbitrary element of LOO(T), then there is a sequence ("pn) in r converging in the L 2 -norm to"p. Hence, lim n -. oo IIw("pn) - W("p)1I2 = 0, and, by going to 
98 3. Ideals and Positive Functionals subsequences if necessary, we may suppose that (1/Jn) converges to 'l/J a.e. and (w(1/Jn)) converges to w(1/J) a.e. If c.p = w(co), then w(1/Jn) = wMt/ln(co) = Mt/ln w( co) = 1/Jnc.p a.e. Hence, w( 1/J) = 1/Jc.p a.e. Let En = {A E T 11c.p(A)1 > IIwll + l/n}, so En is a measurable set, and since IIwll 2 11XEn II > IIw(XEn )II = j Itp(>')12XEn (>') d>' > j(/lw ll + 1/n)2XEn(>') d>' = (lIwll + 1/n)21IXEn II, En is of measure zero. Hence, the set of points A E T such that 1c.p(A)1 > IIwll, which is just the union U=lEn' is a set of measure zero. It follows that 1c.p(A)1 < IIwll a.e., and therefore c.p E LOO(T). Because w is equal to Mcp on LOO(T), and therefore on L 2 (T) by L 2 -norm density of LOO(T) in L 2 (T), the theorem is proved. 0 If E is a Borel set of T, then M XE is a projection on L 2 (T). We call its range K a Wiener vector subspace of L 2 (T). Note that v(I{) = !{. If c.p is a unitary of LOO(T), then c.pH 2 is a closed vector subspace of L 2 (T), called a Beurling vector subspace. Note that c.pH 2 is invariant for v also, but v(c.pH2) i= c.pH 2 (otherwise, c.pH 2 = vc.pH 2 = c.pvH 2 ; therefore, H2 = u(H 2 ), so the unilateral shift is surjective and therefore invertible, which is false). 3.5.3. Theorem. The closed vector subspaces of L 2 (T) invariant for the bilateral shift v = Ml are precisely the Wiener and Beurling spaces. If K is an invariant closed vector subspace for v, then (1) v(K) = K if and only if I{ is a Wiener space, and (2) v(K) i= K if and only if K is a Beurling space. Proof. Let K be a closed vector subspace of L 2 (T) invariant for v. Sup- pose first that v(I{) = I{, and let p be the projection of L 2 (T) onto K. Since K reduces v therefore pv = vp. Hence, by Theorem 3.5.2 there is an element c.p E LOO(T) such that p = Mcp. Because p is a projection, so is c.p, and therefore c.p = XE a.e. where E is a measurable set. Hence, p = M XE ' so K is a Wiener space. Now suppose instead that v(K) i= K. Then there is a unit vector c.p E K such that c.p is orthogonal to v(I{). Since vn(c.p) E v(I{) for all n > 0, it follows that 0 = (vn(c.p), c.p) = J cn(A)Ic.p(A)1 2 dA. Therefore, for any non-zero integer n, we have J €n(A)Ic.p(A)1 2 dA = 0, and so 1c.p12 = a a.e. for some scalar a. Since 1Ic.p112 = 1, therefore a = 1. Thus, c.p is a unitary 
3.5. Toeplitz Operators 99 in LOO(T), and clearly, cpH 2 C I{. Also, (€nCP )nEZ is an orthonormal basis for L 2 (T), and €nc.p E K.l. for n < 0 (because (€nCP, 'l/;) = (cp, €-n'l/;) = 0 for all 'l/; E K, since €-n'l/; E v(K)). It follows from these observations that (€nc.p)nEN is an orthonormal basis for K, and therefore K = cpH2. Thus, K is a Beurling space, and the theorem is proved. 0 We give an interesting application of Theorem 3.5.3 to derive an im- portant result in function theory. 3.5.4. Theorem (F. and M. Riesz). If f is a function in H 2 that does not vanish a.e., then the set of points of T where f vanishes is a set of measure zero. Proof. Let E = f-l{O}, and let K be the L 2 -norm closed vector subspace of H 2 consisting of all elements 9 E H 2 such that 9X E = 0 a.e. Obviously, K is invariant for u, and therefore for v. Observe that n  ou n ( K) c n=oun(H2) = o. Hence, if v(I{) = I{, then K = 0, and therefore since f E K, f = 0 a.e. This contradicts the hypothesis. Hence, v( K) =f I{, so by Theorem 3.5.3, [{ = cpH 2 for some unitary element cp E LOO(T). Consequently, CPXE = 0 a.e., so XE = 0 a.e. Therefore, E is of measure zero. 0 3.5.5. Theorem. The only closed vector subspaces of H 2 reducing for the unilateral shift u are the trivially invariant spaces 0 and H 2 . Proof. Suppose I{ is a non-trivial closed vector subspace of H 2 that re- duces u. Since n=lun(I{) C n=lun(H2) = 0 and K =f 0, therefore u(I{) f:. K, so K is a Beurling space by Theorem 3.5.3. Similarly, H 2 8I{ is a Beurling space. Hence, there are unitaries cP, 'l/; E LOO(T) such that K = cpH 2 and H 2 8 I{ = 'l/;H 2 . For all n > 0, we have €nCP E K and €n'l/; E H 8 [{, so (CnCP, 'l/;) = (cp, €n'l/;) = O. Hence, cp1f has zero Fourier transform, so cp;j; = 0 a.e., a contradiction (since c.p, 'l/; are unitaries). Thus, the only reducing closed vector subspaces for u are the spaces 0 and H 2 . 0 The irreducibility of u is important for the analysis of the Toeplitz algebra which we undertake below. Denote by p the projection of L 2 (T) onto H 2 . If c.p E LOO(T), the operator T",: H 2  H 2 , cp  p( cp f), (that is, the compression of M", to H2) has norm IIT",II < 11c.plloo. We call T", the Toeplitz operator with symbol cpo The map LOO(T)  B(H 2 ), cp  T"" is linear and preserves adjoints; that is, T; = T cp. The latter is true because if f, 9 E H 2 , then (T;(f), g) = (f, T",(g)) = (f, p( c.pg)} = (p(f), cpg) = (<pf,g) = (<pf,p(g)) = (p(<pf),g) = (Tcp(f),g). 
100 3. Ideals and Positive Functionals Therefore, if rp = c.p, then T", is self-adjoint. If c.p E Loo(T), then the matrix (Aij) of M", with respect to the basis (tn )nEZ is constant along diagonals; that is, Aij = Ai+1,j+1 for all i, j. This follows from the fact that M", commutes with v = Ml. Conversely, if w is a bounded operator on L 2 (T) whose matrix with respect to (en) is constant along the diagonals, then it is easily verified that w commutes with v, and therefore w is a Laurent operator by Theorem 3.5.2. From these remarks it is clear that the matrix of a Toeplitz operator with respect to the basis (en)nEN is also constant along its diagonals. One can show conversely (but we shall not) that a bounded operator on H 2 whose matrix with respect to (en) is constant along the diagonals is a Toeplitz operator. The F. and M. Riesz theorem has a bearing on the spectral theory of Toeplitz operators: If c.p E HOO and c.p is not a scalar a.e., then T", has no eigenvalues. For suppose that f E H 2 and A E C and (T", - A )(f) = o. Then (c.p - A)f = 0 a.e. Since c.p - A E H 2 and is not the zero element, the set of points where it vanishes is a null set by Theorem 3.5.4. Therefore, f = 0 a.e. A complication that arises in the theory of Toeplitz operators and distinguishes it from the theory of Laurent operators is that although M",M..p = M",..p for arbitrary c.p, 'ljJ E Loo(T), the corresponding statement for Toeplitz operators is not in general true. For instance, if c.p = e1 and 'ljJ = t-1, then Tcp = u, the unilateral shift, and Tt/J = u*. As u*(eo) = 0, uu* =11, but T"'t/J = Teo = 1, so T",Tt/J =I T",t/J. We can get T",Tt/J = T"'t/J in certain important special cases, as for instance in the following result. 3.5.6. Theorem. Let c.p E LOO and'ljJ E HOO. Then T ",t/J = T ",T t/J and T1j;", = T1j;T",. Proof. Since 'ljJ E Hoo, therefore 'ljJH 2 C H 2 . If f E H 2 , then T",Tt/J(f) = p(c.pp('ljJf)) = p(c.p'ljJf) = T",t/J(f), so T",Tt/J = T",t/J. To get the second equality in the statement of the theorem, observe that Tcj)Tt/J = Tcpt/J, so by taking adjoints, T;T = Tt/J; that is, T1j;T", == T1j;",. 0 Now we examine some aspects of the elementary spectral theory of Toepli tz operators. 3.5.7. Theorem (Hartman-Wintner). Let c.p E Loo(T) and let a(c.p) denote the spectrum of c.p in Loo(T). Then a( c.p) C a( T", ) and r(T",) == IIT",II == 1I'P1100. 
3.5. Toeplitz Operators 101 Proof. Since T", -,,\ == T",-A if ,,\ E C, in order to show that a( c.p) C a(Tcp), it suffices to show that if T", is invertible, then c.p is invertible in LOO(T). Assume then T", is invertible and denote by M the positive number IIT;II1. For all f E H2, IIT;I(f)1I < Mllfll, so replacing f by Tr.p(f) we get IIfll < MIIT",(f)lI. If nEZ, then IIM",(cnf)1I == lIc.pcnfll == lIc.pfll > IIT",(f)1I > IlfiliM == Ilenfll/ M . However, the functions enf are dense in L 2 (T) relative to the L2-norm, since r is L 2 -norm dense in L 2 (T). Hence, for all 9 E L 2 (T) we have IIM",(g)1I > IIglll M, and therefore (M;Mr.p(g), g) > (g, g) 1M 2 , so M;M", > M- 2 > O. It follows that M;M", is invertible, so M", is invertible (by normality of M",). Since the map LOO(T)  B(L 2 (T)), c.p  M"" is an isometric *-homomorphism, c.p is invertible in LOO(T). Now suppose that c.p is an arbitrary element of LOO(T). Since a( c.p) C a(T",), we get liT", II < 11c.p1! 00 == r( c.p) < r(T",) < liT", II, so we have liT", II == r(T",) == 11c.p1100. 0 3.5.8. Theorem. If c.p E LOO(T), then T", is compact if and only if c.p == o. Proof. Let u denote the unilateral shift. Then *n ( ) _ { Cm-n, U em- 0, ifm > n if m < n. Therefore, if f E H 2 , then 00 00 Ilu*n(f)112 == II L (f,cm)cm-nI1 2 == L l(f,em)12, m=n m=n so the sequence (u*n(f)) converges to zero as n  00. If v E B(H 2 ) is of finite rank, then by Theorem 2.4.6 there exists fl, . . . , f nand gl, . . . , gn in H 2 such that v = 2::7=1 fj 0 gj. Therefore, for each positive integer m we have u*mv = 2::7=1 u*m(fj) 0 gj, so lim m -. oo u*mv == o. Hence, for all v E K(H 2 ), lim m -. oo u*mv = 0, because the finite-rank operators are norm- dense in I«H2). Observe that u*T",u == Ttl T",Tl == Ttl"'l == T"" so if T", is compact, then since IIT",II = Ilu*mT",u m II < lIu*mT",1I and lim m -. oo u*mT", = 0, we have T '" = 0, and therefore c.p == o. 0 3.5.9. Lemma. If c.p E C(T) and 'l/J E LOO(T), then T",Tt/J - T"'t/J and T1jJT", - Tt/J", are compact operators. Proof. We show Tt/JT", - Tt/J", E I«H 2 ), and this implies T",Tt/J - Tt/J", == (T;jJTtjJ - T-;;;-)* E I{(H2). By density of the set r of trigonometric poly- nomials in G(T), we may suppose that c.p is a trigonometric polynomial, 
102 3. Ideals and Positive Functionals and by linearity of the map c.p  Tcp, we may even suppose that c.p = Cn for some integer n. If n > 0, then by Theorem 3.5.6 TtPTn = TtPn. Therefore, we need only show TtPT_1e - TtP_1e E I«H2) for all k positive. We show this by induction on k. If f E H 2 , then T tP T _ 1 (f) = p( 1/; p( € -1 f) ) = p( 1/; ( € -1 f - (f , co) € -1 ) ) = TtP_1 (f) - (f, €o)p( 1/;c-1). Hence, T tP T -1 - T tP -1 is an operator of rank not greater than one. Suppose now we have shown that TtPTe_1e - TtP-1e E I«H2) for all 1/; E LOO(T) and some k. Then T tP T _ Ie _ 1 - T tP _ Ie _ 1 = (T tP T _ Ie - T tP _ Ie ) T _ 1 + T tP _ Ie T _ 1 - T( tP - Ie )  _ 1 is compact. This proves the result. o Let A denote the C*-algebra generated by all Toeplitz operators Tcp with continuous symbol c.p, and call A the Toeplitz algebra. We are going to use A to analyse these operators. To do this we need to identify the commutator ideal of A. If A is a C* -algebra, then its commutator ideal [ is the closed ideal generated by the commutators [a, b] = ab - ba (a, b E A). It is easily verified that the commutator ideal is the smallest closed ideal I in A such that AI I is abelian. 3.5.10. Theorem. The commutator ideal of the Toeplitz algebra A is K(H 2 ). Proof. If K is a closed vector subspace of H 2 invariant for A, then 1< is reducing for the unilateral shift u, so by Theorem 3.5.5 K = 0 or K = H2. Thus, A is an irreducible subalgebra of B(H 2 ). Now p = 1 - uu* is a rank-one operator, so pEA n K(H 2 ), and therefore [{(H2) C A by Theorem 2.4.9. The quotient algebra AI [{(H2) is abelian, since it is generated by the elements Tcp + K(H 2 ) (c.p E C(T)), which are commuting and normal by Lemma 3.5.9. Hence, K(H 2 ) contains the commutator ideal I of A. Since I contains p = [u*, u], it is non-zero. Therefore, I = [«H 2 ) because K(H 2 ) is simple (cf. Example 3.2.2). 0 3.5.11. Theorem. The map 1/;: C(T) -+ AI I«H2), c.p  Tcp + J{(H2), is a *-isomorphism. 
3.5. Toeplitz Operators 103 Proof. That 'l/J is linear and preserves adjoints is clear, and by Lemma 3.5.9 it is multiplicative, so it is a *-homomorphism. Since the Toeplitz operators Tcp (cp E G(T)) generate A, the elements Tcp + I{(H2) (cp E G(T)) generate AI K( H2), and therefore 'l/J is surjective. Injectivity of 'l/J is immediate from Theorem 3.5.8. 0 3.5.12. Corollary. If c.p E G(T), then Tcp is a Fredholm operator if and only if cp vanishes nowhere. Proof. By the Atkinson characterisation (Theorem 1.4.16), Tcp is Fred- holm if and only if Tcp + K(H 2 ) is invertible in the quotient B(H 2 )1 K(H2). Hence, Tcp is Fredholm if and only if Tcp + I{(H2) is invertible in AI K(H2). By Theorem 3.5.11, therefore, Tcp is Fredholm if and only if c.p is invertible in G(1r). 0 3.5.13. Corollary. If c.p E C(1r), then ae(Tcp) = c.p(1r). Hence, a Toeplitz operator with continuous symbol has connected essential spectrum. Proof. From the Atkinson characterisation of Fredholm operators and from Theorem 3.5.11, ae(Tcp) = a(Tcp + I{(H2)) = a(c.p) = c.p(1r). 0 A case of particular interest is the unilateral shift u = Tl. It follows from the corollary that a e ( u) = 1r. If nEZ, then the Fredholm index of TEn is -no To see this, one may suppose that n > 0, and then observe that in this case ind(Tn) = ind( un) = n ind( u) = n( -1) (u is the unilateral shift). We shall be generalising this remark shortly and shall need the following elementary result. 3.5.14. Lemma. If c.p is an invertible function in G(T), then there exists a unique integer n E Z such that c.p = €net/J for some 'l/J E G(T). Proof. First let us remark that if 111 - c.p II < 1, then c.p == e t/J for some 1/J E C(1r). In fact we can take 'l/J == in 0 c.p, where in: C \ (-00, 0]  C is the principal branch of the logarithmic function (the hypothesis 111 - c.pll < 1 implies that the range of c.p lies in the domain of in). Hence, if c.p, c.p' are invertible elements of G(1r) such that 1Ic.p - c.p'11 < 11c.p- 1 11- 1 , then cp' = c.pe tP for some 'l/J E G(T). Suppose then c.p is an invertible element of G(1r), and we shall show that c.p = € n e t/J for some n E Z and some 'l/J E C (T). Since r is dense in G(T), we may suppose that c.p E r, by the observations in the first paragraph of this proof. Hence, we may write c.p == ElnlN An€n, for some N > 0 and some An E C. Therefore, c.p = £-Nc.p' for some cp' E r +, so we may suppose that c.p E r +. In this case c.p is a polynomial in z == €1, and therefore a product of a constant and factors of the form z - A, where A f/. 1r. Thus, we may further reduce and suppose that c.p = z - A with IAI =11. If I A I < 1, then II c.p - z II = I A I < 1 = II Z -1 11-1 , so c.p is of the form z e t/J for some 
104 3. Ideals and Positive Functionals 1/J E G(T). Likewise, if IAI > 1, then 11(1 - A-I z) - 111 < 1, so 1 - A-I Z is of the form e t/J for some 1/J E G (T), and therefore c.p == - A e t/J == e t/J' for some 1/J' E G(T). Thus, we have shown that if 'P is invertible in G(T), then c.p == cn e 1/1 for some n E Z and 1/J E G(T). To show uniqueness of n, we need only show that if Cn is of the form etP for some 1/J E G(T), then n == o. Suppose then that Cn == et/J, where n E Z and 'ljJ E G(T). The map a: [0, 1] -+ Z, t  ind(Tet ), is continuous and has discrete range and connected domain, so it is necessar- ily constant. Hence, -n == ind(Te) == 0(1) == 0(0) == ind(T 1 ) == ind(l) == o. This completes the proof. 0 The integer n in Lemma 3.5.14 is called the winding number of c.p (with respect to the origin). We denote it by wn( c.p). 3.5.15. Theorem. Let c.p be an invertible element in G(T). Then the Fredhobn index ofT", is minus the winding number of 'P, that is, ind(T",) == - wn( 'P). Moreover, T", is invertible if and only if it is Fredhobn of index zero, if and only if c.p == e 1/1 for some 1/J E G (T). Proof. If 1/J is a trigonometric polynomial, say 1/J == 2:lnlN Ancn for some N > 0 and An E C, write 1/J' == 2::=0 Ancn and 1/J" == 2::=1 A-nc-n. Then 1/J == 1/J' + 1/J" and 1/J', 1/J" E Hoo. Since Hoo is a closed subalgebra of I -II LOO(T), it follows that et/J , et/J E HOO. Hence, Te-I Te1 == Te-I e' (by Theorem 3.5.6), so Te-I Te1 == T 1 == 1. Likewise, Te1 Te-I == T 1 == 1. Thus, Te1 is invertible. By a similar argument Te1I is invertible, with inverse Te-II. If we suppose now that c.p is an arbitrary element of G(T), then using the density of r in G(T) we may choose a trigonometric poly- nomial 1/J as above such that 111 - e",-1/1 II < 1. Then Te'P == Te1I e'P- e' == Te1I Te'P-Te1 (by Theorem 3.5.6). Since 111 - Te'P- II == 111 - e",-t/JII < 1, the operator Te'P -  is invertible, and we have already seen that Te1I and Te1 are invertible. Hence, Te'P is a product of invertible operators and is therefore invertible. Now suppose that 'P is an arbitrary invertible element of G(T) with winding number n. We show ind(T",) = -n, and to do this we may suppose n > 0 (replace 'P by <p if necessary). Now c.p == e1/1cn for some 1/J E G(T), and T", == T e 1/1 Ten by Theorem 3.5.6. Hence, ind(T",) == ind(T e 1/1 ) + ind(Te n ) == -n, since Te 1/1 is of index zero because it is invertible. The theorem follows. 0 
3.5. Toeplitz Operators 105 3.5.16. Theorem. The spectrum of a Toeplitz operator with continuous symbol is connected. Proof. If c.p E C(T), then by Theorem 3.5.15 u(Tcp) == c.p(T) U {,,\ Eel Tcp - ,,\ is Fredholm of non-zero index}. Therefore, u(T",) is a compact set consisting of the connected compact set c.p(T) and some of its holes, and therefore by elementary plane topology a(Tcp) is connected. 0 The preceding theorem is a simple special case of a deep theorem of Widom which asserts that all Toeplitz operators have connected spectra [Dou 1, Corollary 7.46]. If (H>..)>"EA is a family of Hilbert spaces, u>.. E B(H>..) for all ,,\ E A, and M == sUP.x Ilu.x1l < 00, we define u E B( ffi>..H.x) by u((x>..)>..) == (u>..(x>..))>.. ((x>..)>.. E (fJ>..H>..). It is easily checked that Ilull == M. We call u the direct sum of the family (U>")>"EA, and denote it by (fJ>"EAU>... It is straightforward to verify that the map ffi>..B(H.x)  B(ffi.xH.x), (u.x).x  (fJ.x u >.., is an isometric *-homomorphism of C*-algebras. If u == Tl' then it is easily seen that u generates the C*-algebra A using the fact that C} generates C(T). As we mentioned in the introduction to this section, the algebra A plays a role in K-theory. What makes it useful is that it is the "universal" C*-algebra generated by a non-unitary isometry. This is made precise in Theorem 3.5.18. To establish that theorem we shall need the following result, which is an important structure theorem for isometries. 3.5.17. Theorem (Wold-von Neumann). If v is an isometry on a Hilbert space H, then v is a unitary, or a direct sum of copies of the unilateral shift, or a direct sum of a unitary and copies of the unilateral shift. Proof. We may suppose that v is neither a unitary nor a sum of copies of the unilateral shift. Set I{ == n=ovn(H). Then v(K) == K, and therefore !{ reduces v. Let w be the compression of v to K and w'the compression of v to I{...L. Since w is an isometry, and the equation v( I{) == K implies that w is surjective, therefore w is a unitary. Now set L == (vH)1... For all n > 0, vn(L) C v(H) == L1.., so if m,n E Nand m =I n, then vn(L) is orthogonal to vm(L). We claim 
106 3. Ideals and Positive Functionals that the internal orthogonal sum EI1=ovn(L) is equal to K.l.. To see this, first observe that vn(L) is orthogonal to vn(L.l.) = v n + 1 (H), so vn(L) C K.l., since K C v n + 1 (H). Consequently, EI1=ovn(L) C K.l.. To show the opposite inclusion, it suffices to show that if x E H is orthogonal to vn(L) for all n E N, then x E K. Suppose then x E n=o(vn(L)).l.. We show by induction that x E vn(H) for all n. This is trivially true for n = o. If x E v n ( H), then x = vn(y) for some y E H, and since vn(y) ..1 v n ( L), then y E L.l. = v(H), and therefore x E v n + 1 (H). This proves our claim. If E is an orthonormal basis for L, then U=ovn(E) is an orthonormal basis for K.l.. For each e E E, let Le be the Hilbert subspace of I{.l. having (vn(e))  0 as orthonormal basis. Then K.l. is the internal orthogonal sum tBeEELe, each Le is invariant for v, the compression V e of v to Le is the unilateral shift, and Wi = EI1eEEV e . 0 An interesting consequence of the Wold-von Neumann decomposition is that every non-unitary isometry has spectrum the closed unit disc. This is the case since the unilateral shift has such a spectrum (cf Example 2.3.2), and is a direct summand of every non-unitary isometry. 3.5.1. Remark. If v is a unitary in a unital C*-algebra B, and z: T  C is the inclusion function, then there is a unique unital *-homomorphism c.p: G(T)  B such that c.p(z) = v. To construct c.p, first observe that a( v) C T and that the "restriction" map G(T)  G(a(v)), f  fu(vb is a unital *-homomorphism. One gets c.p by composing this map with the functional calculus G( a( v))  B at v. Since z generates G(T), c.p is unique. 3.5.18. Theorem (Coburn). Suppose that v is an isometry in a unital C*-algebra B, and let u = Tl E A. Then there exists a unique unital *-homomorphism c.p: A  B such that c.p(u) = v. Moreover, if vv* =11, then c.p is isometric. Proof. Sinc u generates A, therefore c.p is unique. We use the universal representation of B to reduce to the case where B is a C*-subalgebra of B(H) for some Hilbert space H, and idH E B. By Theorem 3.5.17, we can write H = EI1).. EA H).. and v = tB)..EAV).., where H).. are Hilbert spaces and each v).. E B( H)..) is a unitary or a unitlateral shift. If v).. is a unitary, then combining Theorem 3.5.11 and Remark 3.5.1, there is a unital *-homomorphism c.p)..: A  B(H)..) such that 'P'\(u) = v)... If v).. is a unilateral shift, then there exists a unitary w)..: H 2  H).. such that v).. = w,\uw (cf. Example 2.3.2). Hence, the map c.p)..: A  B(H)..), a  wAaw1, 
3. Exercises 107 is an isometric unital *- homomorphism such that cP..\ ( u) = v..\. Let (H, c.p) be the direct sum of the family of representations (H..\, cP ..\ ) ..\ of A. Then c.p:A  B(H) is a unital *-homomorphism such that (u) = E9..\ V..\ = v. Moreover, since c.p( u) E Band u generates A, therefore c.p( A) is contained in B. Now suppose that vv* :F 1. Then some v..\o is a unilateral shift. Hence, the representation (H..\o, CP"\o) is faithful, so (H, c.p) is faithful. Therefore, c.p is isometric. 0 3. Exercises In Exercises 1 to 7, A denotes an arbitrary C*-algebra. 1. Let a, b be normal elements of a C*-algebra A, and c an element of A such that ac = cb. Show that a*c = cb*, using Fuglede's theorem (Exer- cise 2.8) and the fact that the element d= ( ) is normal in M 2 (A) and commutes with dl=( ). This more general result is called the Putnam-Fuglede theorem. 2. Let T be a positive linear functional on A. (a) If I is a closed ideal in A, show that I C ker( T) if and only if I C ker( CPr ). (b) We say T is faithful if T( a) = 0 => a = 0 for all a E A +. Show that if T is faithful, then the GNS representation (Hr, CPr) is faithful. (c) Suppose that a is an automorphism of A such that T(a(a)) = T(a) for all a E A. Define a unitary on Hr by setting u(a + N r ) = a(a) + N r (a E A). Show that CPr(a(a)) = ucp(a)u* (a E A). 3. If cp: A  B is a positive linear map between C*-algebras, show that cP is necessarily bounded. 4. Suppose that A is unital. Let a be an automorphism of A such that a 2 = id A . Define B to be the set of all matrices c = C(b) ata)) , 
108 3. Ideals and Positive Functionals where a, b E A. Show that B is a C*-subalgebra of M 2 (A). Define a map c.p: A  B by setting ( a) = ( aa) ) · Show that c.p is an injective *-homomorphism. We can thus identify A as a C*-subalgebra of B. If we set u = ( ), then u is a self-adjoint unitary and B = A + Au. If C is any unital C*-algebra with a self-adjoint unitary element v, and "p: A  C is a *-homomorphism such that "p( a( a )) = v"p( a )v * (a E A), show that there is a unique *-homomorphism "p': B  C extending "p (that is, "p' 0 c.p = "p) such that "p' ( u) = v. (This establishes that B is a (very easy) example of a crossed product, namely B = A x a Z2, the crossed product of A by the two-element group Z2 under the action a. The theory of crossed products is a vast area of the modern theory of C*-algebras. One of its primary uses is to generate new examples of simple C*-algebras. For an account of this theory, see [Ped].) 5. An element a of A + is strictly positive if the hereditary C*-subalgebra of A generated by a is A itself, that is, if (aAa)- = A. (a) Show that if A is unital, then a E A + is strictly positive if and only if a is invertible. (b) If H is a Hilbert space, show that a positive compact operator on H is strictly positive in ]{(H) if and only if it has dense range. (c) Show that if a is strictly positive in A, then r( a) > 0 for all non-zero posi ti ve linear functionals r on A. 6. We say that A is a-unital if it admits a sequence (Un)=l which is an approximate unit for A. It follows from Remark 3.1.1 that every separable C* -algebra is a-unital. (a) Let a be a strictly positive element of A, and set Un = a(a + l/n)-l for each positive integer n. Show that (un) is an approximate unit for A. (Hint: Define gn: a(a)  R by gn(t) = t 2 /(t + l/n). Show that the sequence (gn) is pointwise-increasing and pointwise-convergent to the inclusion z:a(a)  R, and use Dini's theorem to deduce that (gn) converges uniformly to z. Hence, a = lim n -+ oo au n. ) (b) If (Un) _ l is an approximate unit for A, show that a = E=l u n /2n is a strictly positive element of A. Thus, A is a-unital if and only if it admits a strictly positive element. 
3. Exercises 109 7. Let n be a locally compact Hausdorff space. Show that C o (f2) admits an approximate unit (Pn)=I' where all the Pn are projections, if and only if f2 is the union of a sequence of compact open sets. Deduce that if a C*-algebra A admits a strictly positive element a such that a(a) \ {OJ is discrete, then A admits an approximate unit (Pn)=l consisting of projections. (Show that G*(a) is *-isomorphic to Co(a(a) \ {O}).) 8. Let z: T -+ C be the inclusion map. Let (J E [0, 1]. Show that there is a unique aptomorphism a of G(T) such that a(z) = e i27rS z. Define the faithful posit  e linear functional T: C(T) -+ C by setting T(f) = J f dm where m is n malised arc length on T. Show that T(a(f)) = T(f) for all f E C(T). Ded ce from Exercise 3.2 that there is a unitary v on the Hilbert space H T such that CPT( a(f)) = v'PT(f)v* for all f E C(T). Let u be the unitarY'PT(z). Show that vu = e i27rS uv. If (} is irrational, the C*-algebra As generated by u and v is called an irrational rotation algebra, and As can be shown to be simple. See [Rie] for more details concerning As. These al- gebras form a very important class of examples in C*-algebra theory. They are motivating examples in Connes' development of "non-commutative dif- ferential geometry," a subject of great future promise [Con 2]. 9. Let m be normalised Haar measure on T. If A E C, IAI < 1, define T,\: HI -+ C by setting T),(f) = J / w dmw (f E HI). Show that T,\ E (H I )*. By expanding (1 - Aw)-I in a power series, show 00 '" that T,\(f) = En=o f( n)A n. Deduce that the function f:intD -+ C, A  T'\(f), is analytic, where int D = {A E C I IAI < I}. If f, 9 E H 2 , show that fg E HI and T,\(fg) = T'\(f)T,\(g). (Hint: There exist sequences (CPn) and ("pn) in r + converging to f and g, respectively, in the L 2 -norm. Show that the sequence ('Pn1/Jn) converges to fg in the LI-norm, and deduce the result by first showing it for functions in r +. ) 10. If f: int D -+ C is an analytic function and 0 < r < 1, define fr E C(T) by setting fr(A) = f(rA). Set IIfll2 = sUPO<r<I IIfr112' and let H 2 (D) denote the set of all analytic functi ons f: int D -+ C such that II f 112 < 00. If f E H2(D), show that IIfll2 = v E=o I A nI 2 , where f(A) = E  0 An An is the Taylor series expansion of f. Show that H 2 (D) is a Hilbert space with inner product {f, g} = E=o Ani1n, where An = f(n)(O)/n! and J-Ln = 
110 3. Ideals and Positive Functionals g(n)(O)fn! (the operations are pointwise-defined), and show also that the map H 2  H 2 (D), f  j, is a unitary operator. (Thus, the elements of H 2 can be interpreted as analytic functions on int D satisfying a growth condition approaching the boundary. A similar interpretation can be given for the other HP -spaces.) 11. Show that if c.p is a function in LOO(T) not almost everywhere zero, then either Tcp or T; is injective (Coburn). (Hint: If f E ker(Tcp) and 9 E ker(T;), show that c.pfg and cplg E HI. Deduce that c.pfg = 0 a.e. and apply Theorem 3.5.4 to show that f or 9 = 0 a.e.) Deduce that Tcp is invertible if and only if it is a Fredholm operator of index zero. 3. Addenda An ordered group is a pair (G, < ) consisting of an abelian (discrete) group G and a partial order < on G such that for all x, y, z E G the implication x < y => x + z < y + z holds, and either x < y or y < x. If G is an arbitrary abelian group, then there exists an order < on G such that (G, < ) is an ordered group if and only if G is torsion-free, if and only if the Pontryagin dual group G is connected. Let (G, < ) be an ordered group, and set G+ = {x E G I 0 < x}. If f E L 1 (G), denote by j: G  C its Fourier transform, j(x) = fa f(-Y h(x) dm"( (x E G). Here m is the unique Haar measure on G such that m(G) = 1, and LP(G) = LP( G, m). The generalised Hardy space HP = HP( G, < ) is defined to be HP = {f E LP(G) I lex) = 0 (x E G, x < O)}. This is an LP-closed vector subspace of LP( G) for all p E [1,00]. Denote by q the projection of the Hilbert space L2( G) onto H 2 . If c.p E Loo(G), define Tcp E B(H 2 ) by setting Tcp(f) = q(c.pf), and call Tcp a (generalised) Toeplitz operator on H 2 . Much of the classical theory of Toeplitz operators carries over to this situation. Denote by T( G) the C*-subalgebra of B(H 2 ) generated by all Tcp where c.p E C( G), and let I{T( G) be the commutator ideal of T( G). Then T( G) acts irreducibly on H 2 . Let V x = TEll:' where Cx: G  T, ,  ,(x). If W: G+  B is a map to a unital C*-algebra B such that all W x are isometries and W x + y = W x W y (x, Y E G+), then there is a unique *-homomorphism c.p:T(G)  B such that c.p(V x ) = W x (x E G+). 
3. Addenda 111 If G is an ordered subgroup of R, that is, G is a subgroup of R with the order induced from that of R, then KT(G) is a simple C*-algebra (it is *-isomorphic to some J{(H) for a Hilbert space H if and only if G is order isomorphic to 0 or Z). Conversely, if I{T( G) is simple, then G is order isomorphic to an ordered subgroup of R. References: [Dou 2], [Mur]. 
CHAPTER 4 Yon Neumann Algebras A useful way of thinking of the theory of C*-algebras is as "non- commutative topology." This is justified by the correspondence between abelian C*-algebras and locally compact Hausdorff spaces given by the Gelfand representation. The algebras studied in this chapter, von Neumann algebras, are a class of C*-algebras whose study can be thought of as "non- commutative measure theory." The reason for the analogy in this case is that the abelian von Neumann algebras are (up to isomorphism) of the form Loo(n, J-L), where (n, J-L) is a measure space. The theory of von Neumann algebras is a vast and very well-developed area of the theory of operator algebras. We shall be able only to cover some of the basics. The main results of this chapter are the von Neumann double commutant theorem and the Kaplansky density theorem. 4.1. The Double Commutant Theorem There are a number of topologies on B(H) (H a Hilbert space), apart from the norm topology, that playa crucial role, and each has valuable prop- erties that the others lack. The two most important are the strong (opera- tor) topology and the weak (operator) topology. This section is concerned with the former (we shall introduce the weak topology in the next section). One of the reasons for the usefulness of the strong topology is the "order completeness" property asserted in Vigier's theorem (Theorem 4.1.1) which is analogous to the order completeness property of the real numbers R. Henceforth, we shall be using some results concerning locally convex spaces. The relevant definitions and the required results are given in the appendix. Let H be a Hilbert space, and x E H. Then the function Px: B(H)  R+, u  Ilu(x)lI, 112 
4.1. The Double Commutant Theorem 113 is a semi-norm on B(H). The locally convex topology on B(H) generated by the separating family (Px)xEH is called the 3trong topology on B(H). Thus, a net (UA)AEA converges strongly to an operator U on H if and only if u( x) = limA U A (x) for all x E H. It follows that the strong topology is weaker than the norm topology on B(H). With respect to the strong topology, B(H) is a topological vector space, so the operations of addition and scalar multiplication are strongly contin- uous. This is not the case in general for the multiplication and involution operations. 4.1.1. Ezample. Let H be an infinite-dimensional Hilbert space with an orthonormal basis (en)=l. Set Un = el Q9 en. If x E H, then un(x) = (x, en)el, so lim n -. oo Ilun(x)11 = lim n -. oo I{x, en)1 = o. Thus, the sequence (un) is strongly convergent to zero in B(H). Now u = enQgel, so Ilu(x)1I = I(x, el)l. Therefore, lim n -. oo lIu(x)1I = 1 for x = el, so the sequence (u) does not converge strongly to zero. This shows that the operation U .-...+ U * on B(H) is not strongly continuous, and therefore the strong and the norm topologies on B(H) do not coincide. Observe also that Ilunll = lIelllllenll = 1, so the sequence (1Iunll) is not convergent to zero, and therefore the norm 11.11: B(H) -+ R+ is not strongly continuous. The operation of multiplication B(H) x B(H) -+ B(H), (u,v).-...+ uv, is not strongly continuous either (cf. Exercise 4.3). The preceding example shows that the strong topology behaves badly in some respects, but it also has some very good qualities, as we shall prove in the next theorem. Let H be an arbitrary Hilbert space and suppose that (UA)AEA is an increasing net in B(H)sa that converges strongly to U (so U also belongs to B(H)sa). Then U = SUPAUA and (u(x),x) = SUPA{UA(X),x) (x E H). The corresponding statement for decreasing nets in B(H)sa is also true. Both of these observations follow from the fact that if a net (UA)AEA converges strongly to an operator u, then (u(x), y) = limA(uA(x), y) (x, y E H). 4.1.1. Theorem (Vigier). Let (UA)AEA be a net of hermitian operators on a Hilbert space H. Then (UA)AEA is strongly convergent if it is increasing and bounded above, or if it is decreasing and bounded below. Proof. We prove only the case where (UA)AEA is increasing, since the decreasing case can be got from this by multiplying by minus one. Suppose then that (u A ) is increasing and bounded above. By trun- cating the net (that is, by choosing a point Ao E A and considering the truncated net (UA)AAO) we may suppose that (u A ) is also bounded below, by v say. We may further suppose that all U A are positive (by considering the net (u A - v) if necessary). Hence, there is a positive number M such 
114 4. Yon Neumann Algebras that II u,\ II < M for indices ,,\ . It follows that the increasing net (( U A ( X), X) ) is bounded above (by MllxII2), so this net is convergent. Using the polari- sation identity 3 (u'\(x), y) = t L i k (uA(x + iky), x + iky}, k=O we see that ((uA(x),y}) is a convergent net for all x,y E H. Letting a(x,y) denote its limit, it is easy to check that the function a: H 2  C, (x, y)  a(x, y), is a sesquilinear form on H. Moreover, la(x, y)1 = limA l(uA(x), y)1 < M II x 1111 y II, so a is bounded. Hence, there is an operator U on H such that (u(x), y) = a(x, y) for all x, y. Clearly, lIull < M, U is hermitian, and U A < u for all ,,\ E A. Also, lIu(x) - u A (x)1I 2 = II(u - u A )1/2(u - u A )1/2(x)1I2 < Ilu - uAIIII(u - U A )1/2(x)1I2 < 2M((u - uA)(x),x), and limA((u - uA)(x), x) = 0, so u(x) = limA uA(x). Thus, (u.x) converges strongly to u. 0 4.1.1. Remark. If (PA) is a net of projections on a Hilbert space strongly convergent to an operator u, then u is a projection. For u is self-adjoint and (u(x), y) = limA(PA(x), y) = limA(PA(x),p.x(y)) = (u(x), u(y)) = (u 2 (x), y), so U = U 2 . 4.1.2. Theorem. Suppose that (PA)AEA is a net ofprojections on a Hilbert space H. (1) If (PA) is increasing, then it is strongly convergent to the projection of H onto the closed vector subspace (U.xPA(H))-. (2) If (PA) is decreasing, then it is strongly convergent to the projection of H onto n.xp.x(H). Proof. An easy exercise. o Just as for normed vector spaces, we say a family (XA)AEA of elements of a locally convex space is summabZe to a point x if the net (EAEF XA)F (where F runs over all non-empty finite subsets of A) is convergent to x, and in this case we write x = EAEA x A . 
4.1. The Double Commutant Theorem 115 4.1.3. Theorem. Let (PA)AEA be a family of projections on a Hilbert space H that are pairwise orthogonal (that is, PAPA' == 0 if A, A' are distinct indices in A). Then (PA) is summable in the strong topology on B(H) to a projection, P say, such that !!p( x)1I == (L I!PA( X )11 2 )1/2 AEA (x E H). If P == 1, then the map H  E9 PA(H), x  (PA(X)), A is a uni tary. Proof. If F is a finite non-empty subset of A, then PF == LAEF PA is a projection. Therefore, (PF)F is an increasing net of projections, hence strongly convergent to a projection P; that is, the family (PA) is strongly summable to p. Moreover, IIp( X )11 2 = lijP IlpF( x) 11 2 = lijP L IIp.x(x) 11 2 = L IIp.x( x) 11 2 . AEF AEA The observation concerning the case where P == 1 is clear. o If C is a subset of an algebra A, we define its commutant C' to be the set of all elements of A that commute with all the elements of C. Observe that C' is a subalgebra of A. The double commutant C" of C is (C')'. Similarly, C'" == (C")'. Always C C C" and C' == C"'. If A is a normed algebra, then C' is closed. If A is a *-algebra and C is self-adjoint, then C' is a *-subalgebra of A. All of these facts are elementary with easy proofs. 4.1.4. Lemma. Let H be a Hilbert space and A a *-subalgebra of B(H) containing idH. Then A is strongly dense in A". Proof. Let u E A", x E H, and I{ == cl{v(x) ! v E A}. Then I{ is a closed vector subspace of H which is invariant, and therefore reducing, for all v E A, since A is self-adjoint. Thus, if P is the projection of H onto K, then pEA', so pu == up. Hence, u( x) E I{, and therefore there is a sequence (v n ) in A such that u(x) == limnoo vn(x). For each positive integer n the map <p: B(H)  B(H(n»), v  (c5 ij v), is a unital *-homomorphism, so <peA) is a *-subalgebra of B(H(n») con- taining idH(n). Moreover, <p(u) E (<p(A))", for if W E (<p(A))' and v E A then <p( v)w == w<p( v) => VWij == WijV. Hence, Wij E A', so UWij == WijU. 
116 4. Yon Neumann Algebras Therefore, <.p( u)w == w<.p( u). Suppose now that x == (Xl'...' Xn) E H(n). Then by the first paragraph of this proof there is a sequence (vm)m in A such that <.p( u)( x) == limmoo <.p( v m )( x). Hence, u( x j) == limmoo v m ( x j) for j == 1, . . . , n. We show that this implies that u is in the strong closure of A. If W is a strong neighbourhood of u, we must show that W n A is non-empty, and to do this we may suppose that W - u is a basic neighbourhood of o. Therefore, there are elements Xl, . . . , X n E H and a positive number £ such that W - u == {v E B ( H) I II v( x j ) II < £ (j == 1, . . . , n ) } . Hence, there is a sequence (vm)m in A such that u ( x j) == lim v m ( X j ) moo (j == 1, . . . , n ). Consequently, for some N the operator VN E W, so W n A =I 0. 0 Let H be a Hilbert space. If A is a strongly closed *-subalgebra of B(H), we call A a von Neumann algebra on H. Since the strong topology is weaker than the norm topology, a strongly closed set is also norm-closed. Hence, a von Neumann algebra is a C*-algebra. Obviously, B(H) is a von Neumann algebra on H, as is C1 where 1 is the identity map on H. If (H>..)..EA is a family of Hilbert spaces and A).. is a von Neumann algebra on H).. for each index '\, then it is an easy exercise to show that the direct sum EB)..A).. is a von Neumann algebra on EB)..H)... If A is a *-algebra on a Hilbert space H, then its commutant A' is a von Neumann algebra on H (it is straightforward to verify that A' is strongly closed). If A is a von Neumann algebra on Hand p is a projection in A, then pAp is a von Neumann algebra on H. Also, Mn(A) is a von Neumann algebra on H(n). If H is infinite-dimensional, then I«H) is not a von Neumann algebra on H. To see this, let E be an orthonormal basis for H, and for each finite non-empty subset F of E let PF == EeEF eQge. Then PF is a finite-rank pro- jection and the net (PF)F (where F runs over all finite non-empty subsets of E) converges strongly to 1 on H. If I{(H) were a von Neumann algebra, this would imply that it contains 1, and so dim(H) < 00, contradicting our assumption on H. A fundamental result concerning von Neumann algebras is the follow- ing, known as the double commutant theorem. 4.1.5. Theorem (von Neumann). Let A be a *-algebra on a Hilbert space H and suppose that id H E A. Then A is a von Neumann algebra on H if and only if A == A". 
4.1. The Double Commutant Theorem 117 Proof. Immediate from Lemma 4.1.4. o The intersection of a family of von Neumann algebras on a Hilbert space H is also a von Neumann algebra. Thus, for any set C C B(H) there is a smallest "on Neumann algebra A containing C. We call A the von Neumann algebra generated by C. If C is self-adjoint and contains id H , then A. = C". If in addition C consists of commuting elements, then A is abelian (for in this case C C C' => A = C" C C' => A C A'). This implies that there are non-trivial examples of abelian von Neumann algebras. We give an explicit example: 4.1.2. Eample. Let f2 be a compact Hausdorff space, and suppose that I" is a finite positive regular Borel measure on f2. We saw in Example 2.5.1 that the map L OO (f2, 1")  B(L 2 (f2, 1")), c.p  Mcp, is an isometric *-homomorphism. Its range A is a C*-subalgebra of B(H). Denote by B the C*-subalgebra of A of all multiplication operators on L 2 (f2, 1") with continuous symbol. The commutant of B is A (to see this, mimic the proof of Theorem 3.5.2 using the L2- norm density of C(f2) in L2(0,,1-")). Hence, A is a von Neumann algebra on the Hilbert space L2(0" J-L). Since A C A' (because A is abelian) and A' C B' = A, there- fore A = A'. Consequently, A = B", so B is strongly dense in A by Lemma 4.1.4. Let I( be a closed vector subspace of a Hilbert space H and let p be the projection of H onto I(. If u E B(H), let Up = UK be the compression of u to K. It is easy to verify that the map pB(H)p  B(I(), u  UK, is a *- isomorphism. If A is a *-algebraon H andp is in A', set Ap = {up I U E A}. 4.1.6. Lemma. Let A be a *-algebra on a Hilbert space H, and p a projec- tion in A'. Then pAp and Ap are *-algebras on Hand p( H), respectively, and the map pAp  Ap, U  up, is a *-isomorphism. Moreover, if also pEA", then (A')p = (A p )'. Proof. We show only that pEA" =} (A')p = (A p )', because the rest is a routine exercise. Suppose that U E (A')p and v E Ap. Then there exist u and v in A' and A, respectively, such that u = up and v = v p . Hence, for any x E p(H) we have uv(x) = pupv(x) = pvpu(x) = vu(x). Therefore, U E (.4. p )', so (A')p C (A p )'. 
118 4. Yon Neumann Algebras Conversely, suppose now that u E (A p )', and write u = up for some u E pB(H)p. If v E A, then VpU = uV p , so (pvp)pu p = up(pvp)p. Hence, pvpu = upvp, so vu = uv. Consequently, U E A', and therefore u E (A')p. This shows that the inclusion (Ap)' C (A')p holds. 0 The reader should be aware that some authors define a von Neumann algebra on a Hilbert space H to be a *-algebra A on H such that A = A". This automatically ensures that id H E A. However, proofs appear to run more smoothly if von Neumann algebras are defined as we have done. Moreover, we can frequently reduce to the case where A = A", by the trick explained in Remark 4.1.2. Using our definition von Neumann algebras are still unital, but the unit may not be the identity map of the underlying Hilbert space: 4.1.7. Theorem. If A is a non-zero von Neumann algebra, then it is unital. Proof. Suppose that A acts on the Hilbert space H, and let (UA)AEA be an approximate unit for A. By Theorem 4.1.1, (u A ) converges strongly to a self-adjoint operator, p say, and obviously pEA, since A is strongly closed. If x E Hand u E A, then pu(x) = limA uAu(x) = u(x), so pu = u. Hence, p is a unit for A. 0 If p is a projection in a von Neumann algebra A, then pAp is a strongly closed hereditary C*-subalgebra of A, and if also pEA', then Ap is a strongly closed ideal of A. We now prove the converse of these statements. 4.1.8. Theorem. Let A be a von Neumann algebra. (1) If B is a strongly closed hereditary C*-subalgebra of A, then there is a unique projection p E B such that B = pAp. (2) If I is a strongly closed ideal in A, then there is a unique projection q in I such that I = Aq. Moreover, q E A'. Proof. The existence of p and q follows from the observation that B and I are von Neumann algebras and therefore unital by Theorem 4.1.7. Uniqueness is clear in each case. 0 4.1.2. Remark. Let A be a von Neumann algebra on a Hilbert space H and let p be the unit of A. Of course, p is a projection in A'. The map A  Ap, u  up, is a *-isomorphism (by Lemma 4.1.6), and Ap is a von Neumann algebra on p( H) containing idp(H), so Ap = (A p )". This device will be frequently used to reduce to the case where the yon Neumann algebra is its own double commutant. 
4.1. The Double COffimutant Theorem 119 If u is an operator on a Hilbert space H, then its range projection [u] is ...L the projection of H on (u(H))-. We have [u] = [(uu*)1/2], since u(H) = ker( u*) = ker( uu*)1/2 (by the polar decomposition of u*) = (uu* )1/2(H)...L. 4.1.9. Theorem. If A is a von Neumann algebra, then it contains the range projections of all of its elements. Proof. Let A act on H, and let u E A. Since (uu*)1/2 E A, to show that the range projection of u is in A, we may suppose that u > O. Obviously, we may also assume that u < 1. Let Un = u 1 / 2n for n E N. Then (un) is an increasing sequence of positive elements in the closed unit ball of A, so by Theorem 4.1.1 (un) is strongly convergent to a positive operator, p say. If x E H, then II(p2 - u)( x) II < lI(p2 - unp )(x )11 + lie unp - u)( x)" < II(p - un)p(x) II + II(p - un)(x )11. Therefore, (u,) converges to p2 strongly. But u = Un-l for all n > 0, so p = p2. The sequence (un) is in C*(u), so p(H) C (u(H))-. The continuous functions O'(u)  R, t  t 1 + 1 / 2n , form an increasing sequence and converge pointwise to the identity function t  t, so by Dini's theorem, they converge uniformly. Therefore, by the f . I 1 I 1 . 1+1/2n h . I . H unctlona ca cu us, u = 1m n --+ oo u ; t at IS, u = 1m n -+ oo uu n . ence, u = up = pU, so (u(H))- C pCB). Therefore, [u] = pEA. 0 4.1.10. Theorem. Let A be a von Neumann algebra on a Hilbert space H and v an element of A with polar decomposition v = ulvl. Then u E A. Proof. Let w be a unitary in the unital C*-algebra A'. Then w = w*uw is a partial isometry on H such that v = w Iv I and ker( w) = ker( v). It follows, therefore, from the uniqueness of the polar decomposition that w = u, so u and w commute. But A' is the linear span of its unitaries, so u must commute with all elements of A', and therefore u E A" = (A + C1)". By Lemma 4.1.4 there is a net (U.x).xEA in A and a net (a.x).xEA in C such that the net (u.x + 0.x1) converges strongly to u on H. If p = (lvl], then by . ...L Theorem 4.1.9 pEA. SInce (1 - p)(H) = Ivl(H) = ker(lvl) = ker(v) = ker(u), we have u(l - p) = 0; that is, u = up. Therefore, u is the strong limit of the net (u.xp + a.xp) which lies in A, so u E A. 0 4.1.11. Theorem. Suppose that A is a von Neumann algebra on a Hilbert space H. (1) A is the closed linear span of its projections. 
120 4. Yon Neumann Algebras (2) Ifid H E A and u is a normal element of A, then E(S) E A for every Borel set S of a( u), where E is the spectral resolution of the identity for u. (3) If id H E A and v E B( H), then v E A if and only if v commutes with all the projections of A'. Proof. We may suppose in all cases that idH E A. We prove Condition (2) first. Let u be a normal element of A with spectral resolution of the identity for u denoted by E. If v E A', then vu = uv and vu* = u*v, so vf(u) = f(u)v for every f E Bex>(a(u)). In particular, vE(S) = E(S)v for every Borel set S of a(u). Therefore, E(S) E A" = A. Condi tion (1) follows directly from Condition (2), using the fact that the closed linear span of the characteristic functions X s (S a Borel set of a( u)) is Bex>( a( u)) for each normal element u of A. Condition (3) follows immediately from Condition (1), since A' is a von Neumann algebra, and therefore it is the closed linear span of its pro- jections. 0 We give an immediate and important application of this result in the next theorem. First we make some observations. 4.1.3. Remark. If H is a Hilbert space, then B(H)' = C1. For it is obvious that C' = B(H), and since C is a von Neumann algebra containing id H , Theorem 4.1.5 implies that C = C", so C = B(H)'. 4.1.4. Remark. If A is a C*-algebra acting on a Hilbert space Hand S C H, denote by AS the linear span of the set {u( x) I u E A, xES}, and denote by [AS] the closure of AS. We say A acts non-degenerately on H if [AH] = H. Equivalently, for each non-zero element x E H there exists u E A such that u( x) =I o. If A acts non-degenerately on Hand (U..x)..xEA is an approximate unit for A, then (U..x)..xEA converges to 1 = idH strongly on H. (We have to show that lim..x U..x ( x) = x, for all x E H. This is clear if x = u( y) for some u E A and y E H. By taking linear combinations, one gets lim..x u..x( x) = x for all x in AH. Using density of AH in H, it follows that lim..x u..x(x) = x for arbitrary x.) If A acts irreducibly on H and A =I 0, then it acts non-degenerately, since [AH] is a non-zero closed vector subspace of H invariant for A, and therefore equals H. 4.1.12. Theorem. Let A be a non-zero C*-algebra acting on a Hilbert space H. The following conditions are equivalent: (1) A acts irreducibly on H. (2) A' = C1. (3) A is strongly dense in B(H). 
4.1. The Double Commutant Theorem 121 Proof. If P is a projection in B(H), then pEA' if and only if the closed vector subspace p(H) of H is invariant for A. Since A' is a von Neumann algebra, it is the closed linear span of its projections by Theorem 4.1.11, so if A acts irreducibly, then A' has no projections except the trivial ones, and therefore A' = C1. Therefore, (1) => (2). The reverse implication (2) => (1) is clear. If we suppose now that A' = C1, then (A + C1)' = C1, and therefore (A + C1)" = B(H). By Lemma 4.1.4 the C*-algebra A + C1 is strongly dense in B(H). However, since A acts irreducibly on H, it acts non- degenerately, and therefore if (UA)AEA is an approximate unit for A, it is strongly convergent to 1. Therefore, 1 E A- and A- = (A+C1)- = B(H), where the symbol - denotes strong closure. Hence, (2) => (3). Finally, if A is strongly dense in B(H), then A' = B(H)' = C, so (3) => (2). 0 If (cn)=o is the usual orthonormal basis for the Hardy space H 2 and U is the unilateral shift on this basis, then the C*-algebra A generated by U acts irreducibly on H 2 , by Theorem 3.5.5. Hence, A is strongly dense in B(H2) by the preceding theorem, and therefore the von Neumann algebra generated by U is B(H 2 ). Yon Neumann algebras have a plentiful supply of projections, as we saw in Theorem 4.1.11, and this is again illustrated by the following "binary expansion." 4.1.13. Theorem. If A is a hereditary C*-subalgebra of a von Neumann algebra and a is a positive element of A such that lIali < 1, then there is a sequence of projections (Pn)=1 in A such that a = E=I pn/ 2n . Proof. First suppose that A itself is a von Neumann algebra. We may suppose that A acts on a Hilbert space H such that id H E A. We con- struct by induction a sequence of projections (Pn) in A such that 0 < a - Ej=IPj/2 j < 1/2n, for n > 1, and this will prove the result. Let X be the Borel function from 0"( a) to C which is defined by t _ { I, if t > 1/2 X( ) - 0, ift < 1/2. If z is the inclusion map of O"(a) in C, then 0 < z-!X < !, so 0 < a- !PI < !, where PI = X( a). We are, of course, using the Borel functional calculus at a here. Note that PI is a projection and lies in A, since A is a von Neumann algebra containing idH (use Theorem 4.1.11). Thus, we have started the inductive construction. Suppose then that PI, . . . , Pn have been constructed with the required properties. Then b = a- Ej=I pj/2 j is positive and u(b) C [0,1/2 n ]. Define 
122 4. Yon Neumann Algebras the Borel function x: a(b)  C by { I if 1 / 2n+l < t < 1 / 2n X(t) =' . - - 0, otherwIse. Then if z is the inclusion function of a(b) in C, we have 0 < z - X/2n+l < 1/2n+l, so again by the Borel functional calculus the element Pn+l = X(b) is a proiection in A, and 0 < b - Pn+1/ 2n +1 < 1/2n+l. Therefore, 0 < a - E j : pj/2j < 1/2n+l. This completes the induction. Now let us suppose only that A is a hereditary C*-subalgebra of a von Neumann algebra, B say. As before, if a is in the closed unit ball of A+, then a = E:=l pn/2n for a sequence (Pn) of projections in B. Since a > Pn/2n and A is hereditary in B, therefore Pn E A. This proves the theorem. 0 It follows from Theorem 4.1.13 that the closed unit ball of A+ is the closed convex hull of the projections of A. This of course is not true for all C*-algebras-consider C[O, 1], for instance. 4.1.14. Corollary. If A is a hereditary C*-subalgebra of a von Neumann algebra, then it is the closed linear span of its projections. Proof. The algebra A is the linear span of A +, and A + is contained in the closed linear span of the projections by Theorem 4.1.13. 0 If p, q are projections in a C*-algebra A, we say they are (Murray- von Neumann) equivalent, and we write P "V q, if there exists u E A such that P = u*u and q = uu*. It is a straightforward exercise to show that this is indeed an equivalence relation on the projections of A. The relation "V is of fundamental importance in the classification theory of von Neumann algebras (see the Addenda section of this chapter), and in K-theory for C*-algebras (Chapter 7). However, for the moment we need only one small result concerning "V: 4.1.5. Remark. If P, q are infinite-rank projections on a separable Hilbert space H, then P "V q. To see this, choose orthonormal bases (en)=l and (fn)  1 for p(H) and q(H), respectively. Let v:p(H)  q(H) be the uni- tary such that v(e n ) = In for all n. Define u E B(H) by setting u = v on p(H) and u = 0 on (1 - p)(H). It is easily verified that p = u*u and q = uu*. 4.1.15. Theorem. If H is a separable infinite-dimensional Hilbert space, then K(H) is the unique non-trivial closed ideal of B(H). 
4.1. The Double Commutant Theorem 123 Proof. Let I be a non-zero closed ideal of B(H). By Theorems 2.4.5 and 2.4.7, K(H) C I. Now if I Cf:: K(H), then by Corollary 4.1.14 K(H) does not contain all of the projections of I. Hence, I has an infinite-rank projection, p say. If q is any other infinite-rank projection on H, then as we saw in Remark 4.1.5 there is an element u E B(H) such that p = u*u and q = uu*. Therefore, q = q2 = upu*, so q also belongs to I. Hence, I contains all the projections of B( H), whether their rank is finite or infinite, and therefore I = B(H). Thus, we have shown that the only closed ideals of B(H) are 0, K(H), and B(H). 0 4.1.6. Remark. Let H be a separable infinite-dimensional Hilbert space. If (Xn)1 is a dense sequence in H, it is easy to check that K(H) is the closed linear span of the operators x n Q9 x m (n, m > 1) using the fact that K(H) is the closed linear span of the rank-one operators (Theorems 2.4.5 and 2.4.6). Hence, I{(H) is separable. However, B(H) is non-separable (and therefore B(H)/I{(H) is non-separable). To see this, choose an ortho- normal basis (en )=1 for H. For each set S of positive integers, let PS be the projection in B(H) defined by settingps(e n ) = en ifn E S, andps(e n ) = 0 otherwise. Clearly, lips - PS' II = 1 if S =I S'. Hence, the family of operators (ps)s cannot be in the closure of the range of any sequence in B(H). 4.1.16. Theorem. If H is an infinite-dimensional separable Hilbert space, then the Calkin algebra B ( H) / I{ ( H) is a simple C* -algebra. Proof. Let C denote the Calkin algebra and 7r: B(H)  C the quotient map. To see that C is simple, let I be a closed ideal in C. Then?r -1 (I) is a closed ideal in B(H), and therefore by Theorem 4.1.15, 7r- 1 (I) = 0, K(H), or B(H). Hence, I = 0 or C. 0 4.1.7. Remark. Let u,v be operators on a Hilbert space H such that uu* < vv*. Then there is an operator w E B(H) such that u = vw. To see this, observe that Ilu*(x)1I2 = (uu*(x), x) < (vv*(x), x) = Ilv*(x)112. Hence, we get a well-defined norm-decreasing linear map Wo: v*(H)  H by setting wov*(x) = u*(x). Clearly, we can extend Wo to a bounded linear WI on H. Setting w = w, we get the required result, u = vw. We give an application of Theorem 4.1.13 to single operator theory. 4.1.17. Theorem. Let H be a Hilbert space and u a bounded operator on H. Then u is compact if and only if its range contains no infinite- dimensional closed vector subspace. Proof. Suppose that u is compact. Let I{ be a closed vector subspace of u(H) and p the projection of H on K. The linear map v: ker(pu)l.  I{, x  pu(x), 
124 4. Yon Neumann Algebras is compact, and bijective, therefore invertible by the open mapping theorem. Hence, idK = vv- 1 is compact, so K is finite-dimensional. Now suppose conversely that u(H) contains no infinite-dimensional closed vector subspaces. Observe that the range of u is the same as the range of I u * I by the polar decom posi tion of u *, and that u is com pact if and only if lu*1 is compact. Thus, to show that u is compact we may suppose that it is positive, and by rescaling if necessary we may also suppose that u < 1. Hence, 0 < u 2 < 1. It follows from Theorem 4.1.13 that there is a sequence of projections (Pn)  1 on H such that u 2 = E - 1 Pn/2n. Now pn/2n < u 2 , so by Remark 4.1.7 there exists W n E B(H) such that Pn = UW n . Hence, Pn(H) is a closed vector subspace of u(H) and is therefore finite-dimensional by assumption. Consequently, all of the projections Pn are compact operators, and therefore u is compact also. 0 4.2. The Weak and Ultraweak Topologies Preparatory to our introduction of the weak and ultraweak topologies, we show now that L 1 (H) is the dual of K(H), and B(H) is the dual of L 1 (H). Let H be a Hilbert space, and suppose that u E L 1 (H). It follows from Theorem 2.4.16 that the function tr( u.): I{( H) -+ C, v  tr( uv), is linear and bounded, and II tr( u. ) II < II u lit. We therefore have a map L 1 (H) -+ I{(H)*, u  tr( u.), which is clearly linear and norm-decreasing. We call this map the canonical map from L 1 (H) to I{(H)*. 4.2.1. Theorem. If H is a Hilbert space, then the canonical map from L 1 (H) to I«H)* is an isometric linear isomoIphism. Proof. Let 8 denote this map. If 8( u) = 0, then tr( u*u) = 0, so u = o. Thus, (J is injective. Now suppose that r E I{(H)*. Then the function a: H 2 -+ C, (x, y)  r(x Q9 y), is a sesquilinear form on Hand lIuli < IIrll. Hence, there is a unique operator u on H such that (u(x),y) = a(x,y) = r(xQ9Y) (x,y E H). Also, Ilull = Iiali. Let E be an orthonormal basis for H and let u = vlul be the 
4.2. The Weak and Ultraweak Topologies 125 polar decomposition of u. If F is a finite subset of E, set PF = EeEF e Q9 e (so PF is a projection). Then L(lul(e),e) = L(u(e),v(e)} eEF xEF = L a(e, v(e)) xEF = Lr(eQ9v(e)) xEF = r((L e Q9 e)v*) xEF = r(PFv*) < Ilrli. Hence, lIulh < Il r ll, so u E L 1 (H). If x, y E H, then tr(u(x Q9 y)) - (u(x), y} = r(x Q9 y), so tr(u.) equals r on F(H) and therefore on K(H); that is, 8( u) = r. Therefore, 8 is an isometric linear isomorphism. 0 4.2.2. Corollary. L 1 (H) is a Banach *-algebra under the trace-class norm. Proof. It is a dual space, and is therefore complete. o Suppose again that H is a Hilbert space and suppose that v E B(H). Then the function tr(.v): L 1 (H) -+ C, u  tr( uv), is linear and bounded, and II tr(.v)1I < Ilvll, by Theorem 2.4.16. We call the norm-decreasing linear map B(H) -+ L 1 (H)*, v  tr(.v), the canonical map from B(H) to Ll(H)*. 4.2.3. Theorem. If H is a Hilbert space, then the canonical map from B(H) to L 1 (H)* is an isometric linear isomorphism. Proof. Let 8 denote this map. If 8( v) = 0, then 8( v)( x Q9 y) = 0, that is, tr(x Q9 v*(y)) = 0, so (x, v*(y)) = 0, for all x, y E H. Hence, v = 0, and therefore 8 is injective. Now let r E L I (H)*. The map a:H 2 -+ C, (x,y)  r(x Q9y), is a sesquilinear form on H, and lIali < IIrll. Hence, there is a unique operator v E B(H) such that (v(x), y} = r(x Q9 y) for all x, y E H. Also, 
126 4. Yon Neumann Algebras IIvll = lIall < Ilrll. Now 8(v)(x @ y) = tr(x @ v*(y)) = (v(x), y) = r(x @ y), so 8(v) equals r on F(H), and therefore on L 1 (H) (by Theorem 2.4.17, F(H) is dense in L 1 (H) with respect to the trace-class norm). Therefore, 8 is an isometric linear isomorphism. 0 If H is a Hilbert space, the Hausdorff locally convex topology on B(H) generated by the separating family of semi-norms B(H)  R+, u  I(u(x), y)l, (x,y E H) is called the weak (operator) topology on B(H). If (UA)AEA is a net in B( H), then (u A ) converges weakly to an operator U if and only if (u(x), y) = limA (u A ( x), y) (x, Y E H). The ultraweak or a-weak topology on B(H) is the Hausdorff locally convex topology on B(H) generated by the semi-norms B(H)  R+, U  Itr(uv)l, (v E L 1 (H)). The weak topology is weaker than the ultraweak topology. For if (UA)AEA is a net converging ultraweakly to an operator u, then for each x, y E H the net ((uA(x) - u(x), Y))A = (tr((u A - u)(XQ9Y)))A converges to 0, so (u A ) converges to u weakly. Clearly, the weak topology is also weaker than the strong topology. The operations of addition and scalar multiplication are of course weakly and ultraweakly continuous, and it is easy to see that the invol- ution operation u  u* is also continuous for these topologies. We showed in Example 4.1.1 that the involution is not strongly continuous in general, so the weak and strong topologies do not coincide in general. Continuity of multiplication in the weak topology does not hold in general: Let H be a Hilbert space with an orthononnal basis (en)=l and let U be the unilateral shift on this basis. Then the sequences (u *n) and (un) both converge weakly to zero, but the product sequence (u*nu n ) is the constant 1. We have seen that for an arbitrary Hilbert space H the Banach space B(H) is the dual of L 1 (H). It is clear from this that the ultraweak topology is just the weak* topology on B(H). Hence, the closed unit ball of B(H) is ultraweakly compact, by the Banach-Alaoglu theorem. 4.2.4. Theorem. If H is a Hilbert space, then the relative weak and ultraweak topologies on the closed unit ball of B(H) coincide, and hence the ball is weakly compact. Proof. The identity map from the ball with the ultraweak topology to the ball with the weak topology is a continuous bijection from a compact space to a Hausdorff space and is therefore a homeomorphism. 0 
4.2. The Weak and Ultraweak Topologies 127 4.2.1. Remark. The closed unit ball of B(H) is not strongly compact in general. For if we suppose the contrary, then the identity map of the ball with the relative strong topology to the ball with the relative weak topology is a continuous bijection from a compact space to a Hausdorff one, and therefore a homeomorphism. This means the relative strong and weak topologies on the ball coincide. The involution operation U  u* is weakly continuous, but it follows from Example 4.1.1 that in general this operation is not strongly continuous when restricted to the ball. We therefore have a contradiction. If C is a set of operators on a Hilbert space H, then the weak closure of C is contained in C". For if u is in this weak closure, there is a net (U..x)..xEA in C converging weakly to u, and therefore if v E C' we have (uv(x), y) == lim..x(u..xv(x), y) = lim..x(vu..x(x), y) = lim..x(u..x(x), v*(y)) == (u(x), v*(y)) (vu(x), y), so u and v commute, which implies that u E C". 4.2.5. Tlleorem. Suppose that A is a *-algebra on a Hilbert space H containing id H . (1) The weak closure of A is A". (2) A is a von Neumann algebra if and only if it is weakly closed. Proof. This is immediate from the preceding remark and Theorem 4.1.5. In fact, this result is basically just the completion of that theorem. 0 4.2.6. Theorem. Let H be a Hilbert space and r a linear functional on B(H). The following conditions are equivalent: ( 1) r is weakly contin uous. (2) r is strongly continuous. (3) There are vectors Xl, . . . , X n and Yl, . . . , Yn in H such that n r(u) == L(u(Xj),Yj} )=1 (u E B(H)). Proof. The implications (3) => (1) => (2) are clear. To show that (2) => (3), suppose that r is strongly continuous. Then by Theorem A.1 there is a positive number M and there exist Xl,...,X n E H such that Ir(u)1 < M maXljn Ilu(xj)11 for all u E B(H). We may obviously suppose that M = 1. Then n Ir(u)\ < (L Ilu(xj)112)1/2 j=l (u E B(H)). Let Ko be the vector subspace of H(n) consisting of all (U(Xl)'...' u(xn)) (u E B(H)), and let !{ be its norm closure. The function a: I( 0  C, (u ( Xl)' . . . , u ( X n ))  r ( u ), 
128 4. Yon Neumann Algebras is well-defined, linear, and bounded, with lIall < 1, so it extends to a linear norm-decreasing functional on 1< which we also denote by a. By the Riesz representation theorem for linear functionals on Hilbert spaces, there is a unique element Y = (Yl,...' Yn) E 1< such that a(z) = (z, y) (z E K). Hence, n T(U) = a(u(xl)'...' u(xn)) = L(u(Xj), Yj) j=l (U E B(H)). This shows that (2) => (3). o 4.2.7. Theorem. Let H be a Hilbert space and C a convex subset of B(H). Then C is strongly closed if and only if it is weakly closed. Proof. Since the weak topology is weaker than the strong topology, a weakly closed set is strongly closed. Suppose, therefore, that C is strongly closed, and let u be a point in its weak closure. Then there is a net of operators (UA)AEA in C converging to U weakly, and hence, for every weakly continuous linear functional T on B(H), we have r(u) = limA T(U A ). By Theorem 4.2.6, the weakly continuous linear functionals on B(H) are the same as the strongly continuous, so by Corollary A.8, U is in the strong closure of C; that is, u E C. Hence, C is weakly closed. 0 4.2.8. Corollary. If A is a *-algebra on H, then A is a von Neumann algebra if and only if it is weakly closed. Proof. Immediate, since A is convex. o We show now that a von Neumann algebra is the dual space of a Banach space. This is not true for arbitrary C* -algebras. In fact, by a theorem of Sakai every C*-algebra that is a dual space is isomorphic to a von Neumann algebra ([Sak, Theorem 1.16.7]). Suppose that A is a von Neumann algebra on a Hilbert space H. We set A1. = {v E L1(H) I tr(uv) = 0 (u E A)}. This is a vector subspace of L 1 (H), closed with respect to the trace-class norm. Set A* = L 1 (H)/A1.. Then A* is a Banach space when endowed with the quotient norm (corresponding to the trace-class norm). If U E A, we have a well-defined bounded linear functional B( u): A*  C, v + A 1.  tr( uv). The map (J:A(A*)*, uB(u), is clearly norm-decreasing and linear. We call it the canonical map from A to (A*)*. 
4.3. The Kaplansky Density Theorem 129 4.2.9. Theorem. Let A be a von Neumann algebra on a Hilbert space H. Then the canonical map from A to (A*)* is an isometric linear isomorphism. Proof. Let 0: B(H) -+ Ll(H)* and 0': A -+ (A*)* be the canonical maps. By Theorem 4.2.3, B is an isometric linear isomorphism. If B' ( u) = 0, then B( u) = 0, so u = o. Thus, 0' is injective. If r E (A*)* and 1r is the quotient map from Ll(H) to Ll(H)/Al.., then r1r E Ll(H)*, so r1r = B(u) for some u E B(H). To show that u E A, we need only show that tr(uw) = 0 for all w E Al.., using the fact that A is strongly closed, the characterisation of strongly continuous linear functionals on B(H) given in Theorem 4.2.6, and Corollary A.9. But tr(uw) = O(u)(w) = r1r(w) = reO) = o. Therefore, u E A. If v E L1(H), then O'(u)(1r(v)) = tr(uv) = O(u)(v) = r(1r(v)), so 0' (u) = r. Therefore, 0' is a bijection. Observe also that if e > 0, then there exists v E L 1 (H) such that IIvlll < 1 and IB(u)(v)1 > IIB(u)ll- e, so IIrll > Ir(1r(v))1 > lIull-e. Since e was arbitrary, this shows that IIrll > lIuli. It follows that 0' is isometric. 0 It is easy to check that the weak* topology on A is just the relative a-weak topology on A. 4.2.10. Theorem. Let A be a von Neumann algebra on a Hilbert space H, and let r: A -+ C be a linear functional. Then r is a-weakly continuous if and only if there exists u E L 1 (H) such that r( v) = tr( uv) for all v E A. Proof. This follows from the identification A = (A*)*, the remark pre- ceding this theorem, and Theorem A.2. 0 4.2.2. Remark. Let A be a von Neumann algebra on a Hilbert space H containing id H . If u is a normal element of A, then for each f E Boo(a(u)) the element f ( u) is in A. This is so because f is in the closed linear span of the characteristic functions in Boo ( a( u)), and if S is a Borel set of a( u), then the spectral projection X s( u) = E( S) belongs to A by Theorem 4.1.11. From this and the proof of Theorem 2.5.8, it follows that if u E A is a unitary, then u = e iv for some hermitian element v E A. 4.3. The Kaplansky Density Theorem We prepare the way for the density theorem with some useful results on strong convergence. 4.3.1. Theorem. If H is a Hilbert space, the involution u  u* is strongly continuous when restricted to the set of normal operators of B(H). 
130 4. Yon Neumann Algebras Proof. Let x E H and suppose that u, v are normal operators in B(H). Then lI(v* - u*)(x)1I 2 = (v*(x) - u*(x), v*(x) - u*(x)) = IIv(x)1I2 -lIu(x)1I2 + (uu*(x),x) - (vu*(x), x) + (uu*(x), x) - (uv*(x), x) = IIv(x)1I2 -lIu(x)1I2 + ((u - v)u*(x), x) + (x, (u - v)u*(x)) < IIv(x)1I2 -lIu(x)112 + 211(u - v)u*(x)lIlIxll. If (v.;\) ..\EA is a net of normal operators strongly convergent to a nonnal operator u, then the net (1Iv..\( x )11 2 ) is convergent to Ilu( x )11 2 and the net ((u - v..\)u*(x)) is convergent to 0, so (v!(x) - u*(x)) is convergent to o. Therefore, (v!) is strongly convergent to u*. 0 4.3.1. Remark. If S is a bounded subset of B(H), where H is a Hilbert space, then the map S x B(H)  B(H), (v, u)  VU, is strongly continuous. The proof of this is the inequality IIvu(x) - VIU1(X)1I < Ilvllllu(x) - ul(x)1I + II(v - Vl)Ul(X)II. We say that a continuous function f from R to C is strongly continuous if for every Hilbert space H and each net (U..\)..\EA of hermitian operators on H converging strongly to a hermitian operator u, we have (f( u..\)) converges strongly to f ( u ). 4.3.2. Theorem. If f: R  C is a continuous bounded function, then f is strongly con tin uous. Proof. Let A denote the set of strongly continuous functions. This is clearly a vector space (for the pointwise-defined operations), and it follows from Remark 4.3.1 that if f, 9 belong to A and one of them is bounded, then fg E A. We show first that Co(R) C A: Let Ao = A n Co(R). It is easy to verify that Ao is a closed subalgebra of Co(R), and by Theorem 4.3.1 it is self-adjoint. If z: R  C is the inclusion function, then f = 1/(1 + z2) and 9 = zf belong to Co(R) and Ilflloo,llgl/oo < 1. We show that f, 9 E Ao. Let H be a Hilbert space and suppose that v, u are hermitian operators on H. Then g(u) - g(v) = u(l + u 2 )-1 - v(l + V 2 )-1 = (1 + u 2 )-1 [u(l + v 2 ) - (1 + u 2 )v](1 + V 2 )-1 = (1 + u 2 )-1 [u - v + u( v - u )v](l + V 2 )-1. 
4.3. The Kaplansky Density Theorem 131 Therefore, if x E H, Ilg(u)(x) - g(v)(x)11 < 11(1 + u 2 )-I(u - v)(l + v 2 )-I(x)11 + 11(1 + u 2 )-I(u(v - u)v)(l + v 2 )-I(x)1I < II(u - v)(l + v 2 )-I(x)1I + lI(v - u)v(l + v 2 )-I(x)ll, since 11(1 + u 2 )-111 and 11(1 + u2)-lull < 1. Hence, 9 is strongly continuous, and therefore 9 E Ao. Since z E A, we have zg E A, so f = 1- zg E A, and therefore f E Ao. The set {f,g} separates the points of R, and f(t) > 0 for all t, so by the Stone-Weierstrass theorem the C*-subalgebra generated by f and 9 is Co(R). Hence, Ao = Co(R). Now suppose that h E Cb(R). Then hf, hg E Co(R), so hf, hg E A, and therefore zhg E A also. Consequently, h == hf + zhg E A. 0 If C is a convex set of operators on a Hilbert space H, then its strong and its weak closures coincide, by Theorem 4.2.7. If A is a *-subalgebra of B(H), then its weak closure is a von Neumann algebra. These observations are used in the proof of the following theorem, which is known as the density theorem. 4.3.3. Tlleorem (Kaplansky). Let H be a Hilbert space and A a C*-subalgebra of B( H) with strong closure B. (1) Asa is strongly dense in B sa. (2) The closed unit ball of Asa is strongly dense in the closed unit ball of Bsa. (3) The closed unit ball of A is strongly dense in the closed unit ball of B. (4) If A contains idH, then the unitaries of A are strongly dense in the uni taries of B. Proof. If u E Bsa, then there is a net (UA)AEA in A strongly convergent to u, so (u ) converges weakly to u *, and therefore (Re( u A)) is weakly convergent to u. Hence, u is in the weak closure of Asa, and therefore in the strong closure of this set, since its weak and strong closures coincide (because it is convex). This proves Condition (1). Suppose now that u is in the closed unit ball of Bsa. Then by Condi- tion (1) there is a net (UA).xEA in Asa strongly convergent to u. The function f: R  C defined by setting f(t) = t for t E [-1,1] and f(t) = lIt else- where belongs to Co(R), and therefore by Theorem 4.3.2, it is strongly continuous. Hence, (f( u A )) is strongly convergent to f( u). But clearly, f( u) = u, since a( u) C [-1, 1]. Moreover, f( u A ) is in the closed unit ball of Asa for all indices '\, since 1 = f and Ilflloo < 1. This proves Condition (2). The algebra M 2 (A) is a C*-subalgebra of M 2 (B(H)) = B(H(2»), and strongly dense in the von Neumann algebra M 2 (B). If u is in the closed unit ball of B, then v = (* ) is a hermitian operator on H(2) lying in the 
132 4. Yon Neumann Algebras strong closure of M 2 (A), and since IIvll < 1, it follows from Condition (2) that there is a net (V'\)'\EA in the closed unit ball of M2(A)..a that strongly converges to v. Hence, ((V'\)12) is strongly convergent on H to u, and (V,\)12 is in the closed unit ball of A for all indices A. Thus, Condition (3) is proved. Suppose now that A contains idH and let U(A) and U(B) denote the sets of unitaries in A and B, respectively. If u E U( B), then by Re- mark 4.2.2 there is a hermitian element v of B such that u = e iv . By Condi tion (1) there is a net (v,\) '\EA in Asa strongly convergent to v. How- ever, the function f: R  C, t  e it , is strongly continuous by Theorem 4.3.2, so (f( v,\)) converges strongly to f( v). Since f( v,\) = e iv ). E U(A) and f( v) = u, Condition (4) is proved. 0 4.3.4. Theorem. Let HI and H 2 be Hilbert spaces, A a von Neumann algebra on HI, and <p: A  B(H2) a weakly continuous *-homomorpmsm. Then <peA) is a von Neumann algebra on H 2 . Proof. Observe first that <peA) is a C*-algebra on H 2 . By applying Re- mark 4.1.2, we may suppose that A contains idH l . Let v E <p(A) and suppose that IIvll < 1, so there is a number a such that IIvll < a < 1. Write v = <p(u) for some element u E A and let u = wlul be the polar decomposition of u. By Theorem 4.1.10 w E A. Let E be the spectral resolution of the identity for lu I and G the set of all points of the spectrum of lul not less than a. Then E(G) E A, aE(G) < luIE(G), and lul(l - E(G)) < a(l - E(G)). Hence, 0 < a<p(E(G)) < <p(lul)<p(E(G)), so all<p(E(G))11 < 1I<p(lul)lIlI<p(E(G))11 < 1I<p(lul)ll = 1I<p(w*wlul)lI < 1I<p(u)11 = IIvll < a. Therefore, 11<p(E(G))1I < 1, so <p(E(G)) = 0, since <p(E(G)) is a projection. Consequently, v = <p(u(l - E(G))). Also, lIu(l- E(G))II < IIlul(l- E(G))II < a111- E(G)II < a < 1. Set R= {Ul E A IlIuIIi < I}. We have shown that <p(R) = {v E <peA) Illvll < I}. The closed unit ball S of A is weakly compact by Theorem 4.2.4, so <peS) is weakly compact. Observe that S is the weak closure of R. We claim that <peS) is the closed unit ball of <peA): Let v be an element of the closed unit ball of <peA) and take a sequence Cn E (0,1) converging to 1. Now v = <p(u) for some u E A, so CnV E <p(A) and IIcnvll < 1. Hence, CnV = <p(u n ) for some Un E R. Thus, CnV is a sequence in <p(S) converging in norm to v, so v E <p(S). This shows that the closed unit ball of <peA) is contained in <peS), and the reverse inclusion is obvious. 
4.4. Abelian Yon Neumann Algebras 133 Now let u be a non-zero element of the weak closure of <peA) in B(H 2 ). By the Kaplansky density theorem, uiliull is in the weak closure of the closed unit ball of <peA), and therefore u/liull E <p(8), so u E <p(A). This shows that the weak closure of <p(A) is equal to <p(A), so <p(A) IS a von Neumann algebra. 0 4.4. Abelian Von Neumann Algebras In this section we represent abelian von Neumann algebras acting on separable Hilbert spaces in terms of "LOO-algebras." We begin by making some observations concerning separability. 4.4.1. Remark. Let A be a C*-algebra. If A is separable, then of course it is countably generated as a C*-algebra. The converse is also true (the proof is an easy exercise). If A is abelian and separable, then its character space n = n(A) is second countable. For if (an)=1 is a dense sequence in A, let r be the smallest topology on n making all an continuous. The set B of all elements a E A such that a is continuous with respect to r is a C*-subalgebra of A containing all the an, so B = A. Since the weak* topology on n is the smallest one making continuous all of the functions a (a E A), therefore it is equal to r. If we now choose a countable base E for the topology of C, it is easily checked that the finite intersections of the sets a;;1 (U) (n > 1, U E E) form a countable base for the topology of n. 4.4.2. Remark. If H is a separable Hilbert space, then the closed unit ball S of B(H) is separable in the strong topology. We show this: By Remark 4.1.6, K(H) is separable for the norm topology. If u E K(H)', then u(x 0 x) = (x 0 x)u, that is, u(x) 0 x = x 0 u*(x), for all x E H. Since (x 0 y)(H) = Cx if x, y =I- 0, therefore u(x) = r(x)x for some scalar rex) E C. If x, yare linearly independent in H, then the equations u(x + y) = r(x + y)(x + y) and u(x + y) = r(x)x + r(y)y imply that r( x + y) = r( x) = r(y). From this it is immediately seen that u E C1. Therefore, K(H)' = C1, so K(H) is strongly dense in B(H) by Theorem 4.1.12. It follows from Theorem 4.3.3 that the closed unit ball of K(H) is strongly dense in S, and since the ball of K(H) is separable for the norm topology, S is therefore separable for the strong toplogy as claimed. The ball S has another useful property in the case that H is separable: It is metrisable for the relative strong topology. For if (x n) is a dense 
134 4. Yon Neumann Algebras sequence in the ball of H, then the equation d(u, v) = f= lJ(u - ;1(x n )11 n=l defines a metric on S inducing the strong topology. The proof is a straight- forward exercise. Let A be a C*-algebra acting on an arbitrary Hilbert space H and let x E H. If [Ax] = H we call x a cyclic vector for A. We say that y E H is a 3eparating vector for A if for all u E A, u(y) = 0 => u = o. If x is cyclic for A, then it is separating for A'. For suppose v E A' and v( x) = o. If u E A, then vu(x) = uv(x) = O. Hence, v(H) = v[Ax] = 0, and therefore v = o. If A acts non-degenerately on H and if x is a separating vector for A', then it is cyclic for A. To see this, let p denote the projection of H on [Ax]. Then pEA', since [Ax] is invariant for A. Note that x E [Ax], for if (U.x).xEA is an approximate unit for A, then x = lim.x u.x(x), and u.x(x) E [Ax]. Hence, (1 - p)(x) = o. By the separating property, 1 - p = 0, so [Ax] = H; that is, x is cyclic for A, as claimed. An abelian von Neumann algebra A on a Hilbert space H is maximal if it is not contained in any other abelian von Neumann algebra on H. It is easily verified that A is maximal if and only if A' = A. A simple application of Zorn's lemma shows that every abelian von Neumann algebra is contained in a maximal abelian von Neumann algebra. 4.4.1. Eample. Let I" be a finite positive regular Borel measure on a compact Hausdorff space O. The vector 1 E L 2 (0, J-L) is cyclic for the C*-algebra A on L 2 (0,1") consisting of all multiplication operators Mcp with continuous symbol <p (because C(O) is L 2 -norm dense in L 2 (0, I" )). Recall that A' is the algebra of all multiplication operators on L 2 (0, 1") (by Example 4.1.2). The vector 1 is separating (and cyclic) for A'. The von Neumann algebra A' is maximal abelian, since A" = A'. 4.4.1. Len1ma. If A is an abelian von Neumann algebra acting non- degenerately on a separable Hilbert space H, then A has a separating vector. Proof. Let E be a maximal set in H of unit vectors such that the spaces [Ax] (x E E) are pairwise orthogonal (E exists by Zorn's lemma). If y E H is a unit vector orthogonal to all [Ax] (x E E), then [Ay] is orthogonal to all [Ax] also, contradicting the maximality of E. Hence, H is the (inner) orthogonal sum of the spaces [Ax] (x E E). Since H is separable, the set E is necessarily countable, so we may write E = {xn In > I}, where (x n ) is a sequence of unit vectors in H. Set x = 2::=1 xn/2n. If u E A and 
4.4. Abelian Yon Neumann Algebras 135 u(x) = 0, then u(xn) = 0 for all n, because the sequence (u(xn)) consists of pairwise orthogonal elements. Hence, if v E A, then UV(xn) = vu(xn) = 0, so u[Axn] = 0, for all n. It follows that U = 0, so x is a separating vector for A. 0 4.4.2. Theorem. If A is a maximal abelian von Neumann algebra on a separable Hilbert space, then A has a cyclic vector. Proof. By Lemma 4.4.1 A = A' has a separating vector, x say. Hence, x is cyclic for A. 0 4.4.3. Theorem. Let A be an abelian von Neumann algebra acting on a separable Hilbert space H, which contains id H and has a cyclic vector. Then there exists a second countable compact Hausdorff space n, a positive measure J.L E M(O), and a unitary u: H  L 2 (n, 1"), such that uAu* is the von Neumann algebra of all multiplication operators Mcp on L 2 (n, J.L). Proof. Let x be a cyclic vector for A. The closed unit ball of B(H) is metrisable and separable for the strong topology by Remark 4.4.2, so the same is true for the ball of A. It follows that there is a separable C* -subalgebra B of A which is strongly dense in A. We may assume 1 = id H E B. Let cp: B  C(n) be the Gelfand representation and note that the compact Hausdorff space n is second countable by Remark 4.4.1. We define a positive linear functional r on C(n) by setting ref) = (cp-l(f)(x), x}. By the Riesz-Kakutani theorem, there exists a positive measure J.L E M(n) such that r(f) = J f dl" for all f E C(n). The map 7r: B  B(L 2 (n, 1")), v  Mcp(v), is an injective *-homomorphism. If v E B, then J 1(vW dIJ = T(I(vW) = (-l(v*v)(x), x} = IIv(x)112. Hence, the map u:v(x)  cp(v) from Bx = {vex) I v E B} to the L 2 -norm dense subset C(n) of L 2 (n, 1") is well-defined and isometric, and it is clearly linear. Since [Ax] = Hand B is strongly dense in A, therefore [Bx] = H. We may therefore extend u to a unitary (also denoted by u) from H onto L 2 (n,I"). Ifv,w E B, then 7r(v)uw(x) = cp(vw) = uvw(x). Hence, 7r(v)u = uv for all v E B. Therefore, the *-isomorphism Ad u: B( H)  B( L 2 (n, 1")), V t---+ uvu* , is equal to 7r on B. Denote by C the von Neumann algebra of all multiplica- tion operators on L 2 (n, 1"). Since B is strongly dense in A, and the algebra of multiplication operators with continuous symbol is strongly dense in C (by Example 4.1.2), therefore uAu* = C, because Ad u is a homeomorphism for the strong topologies. 0 
136 4. Von Neumann Algebras 4.4.4. Theorem. Let A be an abelian von Neumann algebra on a separ- able Hilbert space H. Then there exists a second countable compact Haus- dorff space f2, and a positive measure I" E M(n) such that A is *-isomorphic to the C*-algebra LOC>(f2, J.L). Proof. We may assume that idH E A. By Lemma 4.4.1 there exists a separating vector x for A. If p is the projection of H onto [Ax], then pEA'. The map cp: A  Ap, u  up, is a *-homomorphism onto Ap, and since x is separating for A, this map is injective and therefore a *-isomorphism. Clearly, cp is weakly continuous, so by Theorem 4.3.4, cp(A) = Ap is a von Neumann algebra on p(H). Obviously, x is cyclic for Ap. Note also that id p ( H) E Ap. Thus, to prove the theorem we have shown we may reduce to the case where A contains idH and has some cyclic vector x. The result now follows from Theorem 4.4.3. 0 Suppose that v is a normal operator on a separable Hilbert space H. The von Neumann algebra generated by v is abelian, so there is a max- imal abelian von Neumann algebra A containing v. By Theorems 4.4.2 and 4.4.3, there is a second countable compact Hausdorff space n and a positive measure I" E M(f2), and a unitary u: H  L 2 (f2, 1"), such that uAu* is the von Neumann algebra of all multiplication operators on L 2 (f2, 1"). In particular, v is unitarily equivalent to a multiplication operator. 4. Exercises 1. Let H be a separable Hilbert space with an orthonormal basis (en )=1 . Prove that the relative weak topology on the closed unit ball S of B(H) is metrisable by showing that the equation d( ) =  I{(u - v)(e n ), em)1 u,v  2n+m n,m=l defines a metric on S inducing the weak topology. 2. Let H be a Hilbert space. (a) Show that a weakly convergent sequence of operators on H is neces- sarily norm-bounded. (b) Show that if ( un) and (v n ) are sequences of operators on H converging strongly to the operators u and v, respectively, then (unv n ) converges strongly to uv. (c) Show that if (un) is a sequence of operators on H converging strongly to u, and if v E I{( H), then (un v) converges in norm to uv. Show that (vu n ) may not converge to vu in norm. 
4. Exercises 137 3. Let H be a Hilbert space with an orthonormal basis (en ) - 1. (a) Denote by A the set of all pairs (n, U) where n is a positive integer, and U is a neighbourhood of 0 in the strong topology of B(H). For (n, U) and (n',U') in A, write (n,U) < (n',U') ifn < n' and U' C U. Show that A is a poset under the relation < , and that it is upwards-directed. (b) Let u denote the unilateral shift on ( en), and note that (u n*) is strongly convergent to zero. If ,,\ = (n..x, U..x) E A, then limn-+oo(n..xun*) = 0 in the strong topology, so for some n we have n..x u n* E U..x. Set U..x = n..xun* and V..x = LU n . Show that lim..x u.;\ = 0 in the strong topology n and lim..x V..x = 0 in the norm topology. (Since U..xV.;\ = 1, this shows that the operation of multiplication B(H) x B(H) -+ B(H), (u,v)  uv, is not jointly continuous in either the weak or the strong topologies.) (c) Show that neither the weak nor the strong topologies on B(H) are metrisable, using Exercise 4.2 and the nets (u..x) and (v..x) from part (b) of this exercise. 4. Let A be a von Neumann algebra on a Hilbert space H, and suppose that T is a bounded linear functional on A. We say that T is normal if, whenever an increasing net (u..x) ..xEA in Asa converges strongly to an operator u E Asa, we have lim..x T(U.x) = T(U). Show that every a-weakly continuous functional T E A* is normal (use Theorem 4.2.10 and show that if (v..x)..x is a bounded net strongly convergent to v, and if u E L 1 (H), then lim..x /lv..x u - vu Ih = 0.) 5. The existence and characterisation of extreme points is very important in many contexts (for example, we shall be concerned with this in the next chapter in connection with pure states). See the Appendix for the definition of extreme points. Let H be a non-zero Hilbert space. (a) Show that the extreme points of the closed unit ball of H are precisely the unit vectors. (b) Deduce that the isometries and co-isometries of B( H) are extreme points of the closed unit ball of B(H). (It can be shown that these are all of the extreme points. This follows from [Tak, Theorem 1.10.2].) 6. Let A be a C*-algebra. (a) Show that if A is unital, then its unit is an extreme point of its closed unit ball. (b) If p is a projection of A, show that it is an extreme point of the closed unit ball of A+ (use the unital algebra pAp and part (a)). The con- verse of this result is also true, but more difficult. It follows from [Tak, Lemma 1.10.1]. 
138 4. Yon Neumann Algebras (c) Show that if H is an infinite-dimensional Hilbert space, then the closed unit ball of B(H)+ is not the convex hull of the projections of B(H). 7. Let A be a C*-algebra. Show that if p, q are equivalent projections in A, and r is a projection orthogonal to both (that is, rp = rq = 0), then the projections r + p and r + q are equivalent. If H is a separable Hilbert space and p is a projection not of finite rank, set rank(p) = 00. If p has finite rank, set rank(p) = dimp(H). Show that p rv q in B(H) if and only if rank(p) = rank(q). Thus, the equivalence class of a projection in a C*-algebra can be thought of as its "generalised rank." We say a projection p in a C*-algebra A is finite if for any projection q such that q rv p and q < p we necessarily have q = p. Otherwise, the projection is said to be infinite. Show that if p, q are projections such that q < p and p is finite, then q is finite. A projection p in a von Neumann algebra A is abeZian if the algebra pAp is abelian. Show that abelian projections are finite. A von Neumann algebra is said to be finite or infinite according as its unit is a finite or infinite projection. If H is a Hilbert space, show that the von Neumann algebra B(H) is finite or infinite according as H is finite- or infini te- dimensional. 4. Addenda Let T be a bounded linear functional on a von Neumann algebra A. The following are equivalent conditions: (i) T is normal. (ii) The restriction of T to the closed unit ball of A is weakly continuous. (iii) T is a-weakly continuous. Reference: [Ped, Theorem 3.6.4]. A projection in a von Neumann algebra A is central if it commutes with every element of A. We say that A is Type I if every non-zero central projection in A majorises a non-zero abelian projection in A. Thus, abelian von Neumann algebras are trivially Type I. Just as easy, B(H) is Type I for every Hilbert space H. We say that A is Type II if it has no non-zero abelian projections and every non-zero central projection majorises a non-zero finite projection. We say that A is Type III if it contains no non-zero finite projections. We say that A is properly infinite if it has no non-zero finite central projection. If A is Type II and properly infinite, it is said to be Type 110c>, and if it is Type II and finite, it is said to be Type III. 
4. Addenda 139 Each von Neumann algebra can be decomposed into a direct sum of von Neumann algebras of Types I, III, 11 00 , and III (not all types may be present ). If A is a von Neumann algebra on the Hilbert space H, it is said to be a factor if A n A' = C1, where 1 = id H . Of course, B(H) is a factor, but one has to work harder to get other examples (see Section 6.2). A factor is one and only one of the Types I, 111, 11 00 , III. A factor of Type I is isomorphic to B(H) for some Hilbert space H. Every von Neumann algebra can be "decomposed" into factors, so these are the building blocks of the theory. References: [Dix 1], [Tak]. 
CHAPTER 5 Representations of C*-Algebras This chapter is concerned with the positive linear functionals, the representations, and the (left and two-sided) ideals of a C* -algebra, and with their inter-relationship. Pure states are introduced and shown to be the extreme points of a certain convex set, and their existence is deduced from the Krein-Milman theorem (see Appendix). From this the existence of irreducible representations is proved by establishing a correspondence between them and the pure states. We introduce two analogues of the spectrum of an abelian Banach algebra, the space of primitive ideals and (loosely speaking) the space of irreducible representations. These spaces are related to the structure theory of the underlying algebra. We are also interested in this chapter in the relationship of the repre- sentations and ideals of an algebra to the corresponding objects of a sub- algebra and a quotient algebra (for the case of a subalgebra this works out nicest if it is hereditary). The chapter concludes with a brief introduction to the important classes of lirninal and postliminal C*-algebras. 5.1. Irreducible Representations and Pure States If (H, 'P) is a representation of a C* -algebra A, we say x E H is a cyclic vector for (H, 'P) if x is cyclic for the C*-algebra 'P(A). If (H, 'P) admits a cyclic vector, then we say that it is a cyclic representation. We return now to the GNS construction associated to a state to show that the representations involved are cyclic. 140 
5.1. Irreducible Representations and Pure States 141 5.1.1. Theorem. Let A be a C*-algebra and T E SeA). Then there is a unique vector X T E H T such that T(a) = (a + NT,x T ) (a E A). Moreover, X T is a unit cyclic vector for (HT,'PT) and 'PT(a)XT = a + NT (a E A). Proof. The function po: A / NT --+ C, a + NT t---+ T ( a ), is well-defined, linear, and norm-decreasing, so we can extend it to a norm- decreasing linear functional p on H T . By the Riesz representation theorem, there is a unique vector x T in H T such that p(y) = (y, X T) (y E H T). Thus, X T is the unique element of H T such that T(a) = (a + NT, X T ) (a E A). Let a,b E A. Then (b+NT,'PT(a)x T ) = (a*b+NT,x T ) = T(a*b) = (b + NT, a + NT)' and since this holds for all b, we have 'PT( a )XT = a + NT. Hence, 'PT(A)XT is dense in H T , since it is the space A/NT. Therefore, X T is cyclic for (H T , 'PT). Consequently, 'PT(A) acts non-degenerately on H T . If (UA)AEA is an approximate unit for A, then ('PT(U A )) is one for 'PT(A), and therefore it converges strongly to id Hr . Hence, IIx T II 2 = (XT,X T ) = limA('PT(uA)(xT),x T ) = limA T(U A ) = IITII = 1, so X T is a unit vector. 0 We call the vector X T in Theorem 5.1.1 the canonical cyclic vector for (H T , 'PT). If p, T are positive linear functionals on a C*-algebra A, we write p < T if T - P is positive. We say T majorises p, or p is majorised by T, if p < T. 5.1.2. Theorem. Let T be a state and p a positive linear functional on a C*-algebra A, and suppose that p < T. Then there is a unique operator v in 'PT( A)' such that pea) = ('PT(a)vxT,x T ) (a E A). Moreover, 0 < v < 1  Proof. Define a sesquilinear form a on A/ NT by setting a( a + NT, b + NT ) = p(b*a) (this is well-defined as p < T). Observe that lIall < 1 as Ip(b*a)1 < p(b*b)1/2p(a*a)1/2 < T(b*b)1/2T(a*a)1/2 = lib + NTlilia + NTII. We can therefore extend a to a bounded sesquilinear form (also denoted a) on H T which also has norm not greater than 1. Hence, there is an operator v on H T such that (v(x), y) = a(x, y) for all x, y E H T , and IIvll < 1. Therefore, p(b*a) = a(a + NT,b+ NT) = (v(a + NT),b+ NT) = (v'PT(a)xT,'PT(b)x T ). Consequently, (v( a + NT)' a + NT) > 0 for all a E A, so v is positive. 
142 5. Representations of C*-Algebras Ifa,b,cE A, then (<Pr(a)v(b+Nr),c+N r ) = (v(b+Nr),a*c+N r ) = p(c*ab) = (v(ab+Nr),c+N r ) = (v<pr(a)(b+Nr),c+N r ). Hence, <Pr(a)v = vcpr(a) for all a, so v E <f'r(A)'. Also, p(a*b) = (v(b + N r ), a + N r ) = (v<pr(b)xr,<Pr(a)xr) = (v<pr(a*b)xr,xr), so if (U..\)..\EA is an approximate unit for A, then p(u..\b) = (vcpr(u..\b)xr, x r ), and therefore in the limit p(b) = (v<pr(b)xr,xr). To see uniqueness, suppose that W E <Pr(A)' andp(a) = (<pr(a)wxr,Xr) (a E A). Then (W<pr(a*b)x r , x r ) = p(a*b) = (vcpr(a*b)xr, x r ), and therefore (w(b+Nr),a+N r ) = (v(b+Nr),a+N r ) for all a, b. Hence, W = v. o It is an easy exercise to show that we can go in the opposite direction also; that is, if v E CPr(A)' and 0 < v < 1, then the equation p( a) = (<pr(a)VXr, x r ) defines a positive linear functional p on A such that p < T. A representation (H, cp) of a C*-algebra A is non-degenerate if the C*-algebra cp(A) acts non-degenerately on H. It is clear that a direct sum of non-degenerate representations is non- degenerate, and also that cyclic representations are non-degenerate. There- fore, the universal representation of A is non-degenerate. If (H, <p) is a non-degenerate representation for A and (U..\)..\EA an ap- proximate unit of A, then (cp( u..\))..\ is an approximate unit of cp(A), so the net (<p(u..\)) converges strongly to id H . Let (H, <p) be an arbitrary representation of A. If K is a closed vector subspace of H invariant for cp(A), then the map <PK: A -+ B(K), a  (<p(a))K, is a *-homomorphism, so the pair (K,cpK) is a representation of A also. If K = [<p(A)H], then K is invariant for cp(A) and the representation (K,cpK) is non-degenerate. Moreover, IIcp(a)1I = IIcpK(a)1I (a E A). We shall often use this device to reduce to the case of a non-degenerate representation. 5.1.3. Theorem. Let (H,<p) be a non-degenerate representation of a C*-algebra A. Then it is a direct sum of cyclic representations of A. Proof. For each x E H, set Hz = [cp(A)x]. An easy application of Zorn's lemma shows that there is a maximal set A of non-zero elements of H such that the spaces Hz are pairwise orthogonal for x E A. If y E (UzEAHx)..L, then for all x E A we have (y,<p(a*b)(x)) = 0, so (cp(a)(y),<p(b)(x)) = 0, 
5.1. Irreducible Representations and Pure States 143 and therefore the spaces H y, H x are orthogonal. Observe that since (H, 'P ) is non-degenerate, y E Hy. It follows from the maximality of A that y = o. Therefore, H is the orthogonal direct sum of the family of Hilbert spaces (H x )xEA. Obviously, these spaces are invariant for 'P( A), and the restriction representation <Px: A -+ B(Hx), a  <p( a)H:r:' has x as a cyclic vector. Since (H, 'P) is the direct sum of the representations ( H x, <P x) the theorem is proved. 0 Two representations (HI, 'PI) and (H 2 ,<P2) of a C*-algebra A are uni- tarily equivalent if there is a unitary u : HI -+ H 2 such that 'P2 ( a ) = u'PI(a)u* (a E A). It is readily verified that unitary equivalence is indeed an equivalence relation. 5.1.4. Theorem. Suppose that (HI, 'PI) and (H2' 'P2) are representations of a C*-algebra A with cyclic vectors Xl and X2, respectively. Then there is a unitary u: HI -+ H 2 such that X2 = U(Xl) and 'P2(a) = u'PI(a)u* for all a E A if and only if ('PI(a)(xI), Xl) = ('P2(a)(x2), X2) for all a E A. Proof. The forward implication is obvious. Suppose, therefore, that we have ('PI(a)(xI),XI) = ('P2(a)(x2),X2) for all a E A. Define a linear map Uo: 'PI(A)XI -+ H 2 by setting uO('PI(a)(xI)) = 'P2(a)(x2). That this is well-defined and isometric follows from the equations 1I'P2(a)(X2)11 2 = ('P2(a*a)(x2),X2) = ('PI(a*a)(xI),XI) = II'PI(a)(xI)1I2. We extend Uo to an isometric linear map U:HI -+ H 2 , and since U(Hl) = ['P2(A)X2] = H 2 , U is a unitary. If a, b E A, then u'Pl(a)'PI(b)XI = 'P2(ab)(x2) = 'P2(a)u'PI(b)(XI). Therefore, u'PI(a) = 'P2(a)u (a E A). Now 'P2(a)u(xI) = u<pI(a)(xI) = 'P2(a)(x2)' so 'P2(a)(u(xl) - X2) = o. By non-degeneracy of c.p2, therefore, u( Xl) = X2. 0 A representation (H, 'P) of a C* -algebra A is irreducible if the algebra 'P(A) acts irreducibly on H. If two representations are unitarily equivalent, then irreducibility of one implies irreducibility of the other. If H is a one- dimensional Hilbert space, then the zero representation of any C*-algebra on H is irreducible. 5.1.5. Theorem. Let (H, 'P) be a non-zero representation of a C*-algebra A. (1) (H,c.p) is irreducible if and only ifc.p(A)' = C1, where 1 = id H . (2) If (H, 'P) is irreducible, then every non-zero vector of H is cyclic for (H,'P). 
144 5. Representations of C*-Algebras Proof. Condition (1) is immediate from Theorem 4.1.12. Suppose that (H, cp) is irreducible, and that x is a non-zero vector of H. The space [cp(A)x] is invariant for cp(A), and therefore is equal to 0 or H. Because cP is non-zero, there is some element y of H and some element a of A such that cp(a)(y) :F o. Hence, [<p(A)y] = H, so <p is non-degenerate. It follows that cp(A)x is not the zero space, so [cp(A)x] = H; that is, x is a cyclic vector for (H, cp ). 0 We say a state T on a C*-algebra A is pure if it has the property that whenever p is a positive linear functional on A such that p < T, necessarily there is a number t E [0, 1] such that p = tT. The set of pure states on A is denoted by PS(A). 5.1.6. Theorem. Let T be a state on a C*-algebra A. (1) T is pure if and only if (H r, CPr) is irreducible. (2) If A is abelian, then T is pure if and only if it is a character on A. Proof. Suppose that T is a pure state. Let v be an element of CPr(A)' such that 0 < v < 1. Then the function p: A  C, a  (CPr(a)v(xr), x r ), is a positive linear functional on A such that p < T. Hence, there exists t E [0,1] such that p = tT, and therefore (CPr(a)v(xr),xr) = (tcpr(a)(xr),xr) for all a E A. Consequently, (V (a + Nr),b + N r ) = (vcpr(a)(xr),CPr(b)(xr)) = (vcpr(b*a)(xr), x r ) = (tcpr(b*a)(x r ), x r ) = (t(a+Nr),b+N r ) for all a, b E A. Therefore, v = t1, since A/N r is dense in Hr. It follows that CPr(A)' = C1, so (Hr,CPr) is irreducible by Theorem 5.1.5. Now suppose conversely that (Hr, CPr) is irreducible, and let p be a positive linear functional on A such that p < T. By Theorem 5.1.2 there is a unique operator v in CPr(A)' such that 0 < v < 1 and p(a) = (CPr(a)v(xr), x r ) for all a E A. But CPr(A)' = C1, again by Theorem 5.1.5, so v = t1 for some t E [0,1]. Hence, p = tT, so T is pure. This proves the equivalence in Condition (1). Assume now that A is abelian. If T is pure, then CPr(A)' = C1. But CPr(A) C CPr(A)', so CPr(A) consists of scalars, and therefore B(Hr) C CPr(A)'. Hence, B(Hr) = Cl. There- fore, if u, v E B(Hr) they are scalars, and (UV(xr), x r ) = u(V(xr), x r ) = u(xr,xr)(v(xr),xr) = (u(xr),xr)(v(xr),xr). Hence, T = (CPr(.)xr,xr) is multiplicative and therefore a character on A. 
5.1. Irreducible Representations and Pure States 145 Now suppose conversely that r is a character on A, and let p be a positive linear functional on A such that p < r. If r( a) = 0, then r( a* a) = 0, so p(a*a) = O. Since Ip(a)1 < p(a*a)1/2, therefore p(a) = o. Hence, ker( r) C ker(p), and it follows from elementary linear algebra that there is a scalar t such that p = tr. Choose a E A such that r( a) = 1. Then r(a*a) = 1, so 0 < p(a*a) = tr(a*a) = t < r(a*a) = 1, and therefore t E [0,1]. This shows that r is pure, and the equivalence in Condition (2) is proved. 0 It follows from Theorem 5.1.6 that for an arbitrary abelian C*-algebra A, PS(A) = f2(A). The only thing not obvious is that a character r on A must have norm 1. To see this, let (UA)AEA be an approximate unit for A. Then (ui) is also an approximate unit. Hence, IIrll = limA r(ui) (limA r( u A ))2 = IIr1l2, so IIrll = II r 11 2 , and therefore IIrll = 1. 5.1.7. Theorem. Let (H,c.p) be a representation of a C*-algebra A, and let x be a unit cyclic vector for (H, c.p). Then the function r: A  C, a  (c.p( a)(x), x), is a state of A and (H, c.p) is unitarily equivalent to (H T , c.p r ). Moreover, if (H, c.p) is irreducible, then r is pure. Proof. Clearly r is a positive linear functional on A. If (U"\)AEA is an approximate unit for A, then because (H, c.p) is non-degenerate the net (c.p(UA))A is strongly convergent to id H . Hence, IIrll = limA r(u A ) - limA (c.p( uA)(x), x) = (x, x) = 1, so r E S(A). For all a E A, (c.pr(a)(xr),x r ) = r(a) = (c.p(a)(x),x), so (Hr, c.pr) and (H, c.p) are unitarily equivalent by Theorem 5.1.4. If (H, cp) is irreducible, so is (Hr, CPr), so by Theorem 5.1.6 r is pure.D 5.1.1. Ezample. Let H be a non-zero Hilbert space, and A = K(H). We are going to determine the pure states of A. If x E H, then the functional W x : A  C, U  (u(x), x), is positive, and if x is a unit vector, W x is a state. The pure states of A are precisely the states W x where x is a unit vector of H. To prove this, suppose first that x is a unit vector of H, and let i: A  B(H) be the inclusion map. The representation (H, i) is irreduc- ible, since A' = C (cf. Remark 4.4.2). Hence, x is a cyclic vector for A, and it follows from Theorem 5.1.7 that the representations (H Wz: , CPwz:) and (H, i) are unitarily equivalent and W x is pure. 
146 5. Representations of C*-Algebras Now suppose conversely that T is a pure state of A. By Theorem 4.2.1 there is a trace-class operator u on H such that r( v) = tr( uv) for all v E A. For any unit vector x of H, the operator x0x is a projection and therefore positive, so 0 < r(x 0 x) = tr( u(x 0 x)) = tr( u(x) 0 x) = (u(x), x). This shows that the operator u is positive. Since u is a compact normal operator, it is diagonalisable by Theorem 2.4.4; that is, there is an orthonormal basis E for H and there is a family of scalars (Ae )eEE such that u( e) = Aee (e E E). Choose eo E E. If v E A+, T(V) = tr(vu) = L(vu(e),e) = L Ae(v(e),e) > Aeoweo(v). eEE eEE Thus, the pure state T majorises the positive linear functional Aeoweo, so there exists t E [0, 1] such that Aeoweo = tT. Since both Weo and T are of norm one, Aeo = t, so Weo = T; that is, T is of the required form. An interesting consequence of our characterisation of the pure states of A is that every non-zero irreducible representation (K, 1/J) of A is unitarily equivalent to the identity representation (H, i) of A. To see this, let y be a unit vector in K. The function p: A -+ C, U  (1/J( u )(y), y), is a pure state on A, and (K, 1/J) is unitarily equivalent to (Hp, 'Pp) by Theorem 5.1.7. Hence, there exists a unit vector x in H such that p = W x . Thus, (K, 1/J) is unitarily equivalent to (Hw:e' 'Pw:e)' and we have already seen above that (Hw:e' 'Pw:e) is unitarily equivalent to (H, i). 5.1.8. Theorem. If A is a C*-algebra, then the set S of norm-decreasing positive linear functionals on A forms a convex weak* compact set. The extreme points of S are the zero functional and the pure states of A. Proof. It is easy to check that S is weak* closed in the closed unit ball of A*, and therefore weak* compact by the Banach-Alaoglu theorem. Con- vexi ty of S is clear. Let E be the set of extreme points of S. First we show 0 E E: Suppose 0 = tT + (1 - t)p, where 0 < t < 1 and T, pES. If a E A, then 0 > -tT(a*a) = (1 - t)p(a*a) > O. Hence, T = P = 0 on A+, and therefore on A, so 0 is an extreme point of S. Next we show that PS(A) C E: Suppose that p is a pure state of A, and that p = tT + (1 - t)T', where 0 < t < 1 and T,T' E S. Then tT is a positive linear functional on A majorised by p, so there exists t' E [0, 1] such that tT = t' p, because p is pure. Since 1 = IIpll = tllTIl + (1 - t)IIT'II, we have IITII = IIT'II = 1. It follows that t = IltTIl = Ilt'pll = t', so T = p. Hence, (1 - t)T' = (1 - t)p, so T' = p. Therefore, pEE. Finally, we suppose that p is a non-zero element of E and show that it is a pure state: Since p = Ilpll(p/llpll) + (1 - Ilpll)O and 0, p/llpil E S, we 
5.1. Irreducible Representations and Pure States 147 have IIpll = 1, because pEE. If r is a non-zero positive linear functional on A majorised by, and not equal to, p, then for t = Ilrll E (0,1) we have p = t(r/llrll) + (1 - t)(p - r)/lIp - rll, since 1 - t = lip - rll. Hence, p = r Illrll, since p is an extreme point of S. Therefore, r = IIrlip. This proves that p is a pure state of A. 0 5.1.9. Corollary. The set S is the weak* closed convex hull of 0 and the pure states of A. Proof. Apply Theorem A.14. o 5.1.10. Corollary. If A is a unital C*-algebra, then S(A) is the weak closed convex hull of the pure states of A. Proof. The set SeA) is a non-empty convex weak* compact set, so by Theorem A.14 it is the weak* closed convex hull of its extreme points. It is clear that S( A) is a face of S, where S is as in Theorem 5.1.8, so by that theorem the extreme points of SeA) are the pure states of A. 0 5.1.11. Theorem. Let a be a positive element of a non-zero C*-algebra A. Then there is a pure state p of A such that II a II = p( a ). Proof. We may suppose that a =I o. The function a: A* -+ C, r  rea), is weak* continuous and linear, and by Theorem 3.3.6 lIall = sup{r(a) I rES}, where S is the weak* compact convex set of all norm-decreasing positive linear functionals on A. The set F = {r E S I rea) = lIall} is a weak* compact face of S by Lemma A.13, and therefore has an extreme point p by Theorem A.14. Since F is a face in S, the functional p is an extreme point of S also. Now p =I 0, since lIall = p( a), and a =I o. Therefore, p is a pure state of A by Theorem 5.1.8. 0 It follows from Theorem 5.1.11 that a non-zero C*-algebra has pure states. 5.1.12. Theorem. Let A be a C*-algebra, and a E A. Then there is an irreducible representation (H,cp) of A such that lIall = Ilcp(a)lI. P-.roof. By the preceding theorem, there is a pure state p of A such that p(a*a) = Ila*all. By Theorem 5.1.6, the representation (Hp, cpp) is ir- reducible. Since lIal1 2 = p(a*a) = (cpp(a*a)(x p ), x p ) = IIcpp(a)(x p )1I 2 < Ilcp p( a )11 2 < lIa11 2 , therefore lIall = Ilcp p( a )11. 0 The characterisation given in Theorem 5.1.8 allows us to prove another extension theorem for positive functionals. 
148 5. Representations of C*-Algebras 5.1.13. Theorem. Let B be a C*-subalgebra of a C*-algebra A, and let p be a pure state on B. Then there is a pure state pi on A extending p. Moreover, if B is hereditary in A, then pi is unique. Proof. The set F of all states on A extending p is a weak* compact face of the set S of norm-decreasing positive linear functionals on A (by Theorem 3.3.8 F is non-empty). By Theorem A.14 F admits an extreme point, pi say. Hence, pi is an extreme point of S, and non-zero, so by Theorem 5.1.8 pi is a pure state of A. Uniqueness of pi when B is hereditary is given by Theorem 3.3.9. 0 5.1.14. Theorem. Let A be a unital C*-algebra. Suppose that S is a subset of S(A) such that if a hermitian element a E A satisfies the condition r( a) > 0 for all rES, then necessarily a E A +. Then the weak* closed convex hull of S is S(A) and the weak* closure of S contains PS(A). Proof. Let C denote the weak* closed convex hull of S. It follows from Theorem 5.1.8 that PS(A) is the set of extreme points of S(A), so by Theorem A.14 if we show that SeA) = C, then PS(A) is contained in the weak* closure of S. Suppose that C i= SeA) and we shall obtain a contradiction. Since the containment C C SeA) clearly holds, there exists r E SeA) such that r f/. C. By Theorem A.7 there is a weak* continuous linear functional 8: A *  C and there is a real number t such that Re( 8( r)) > t > Re( 8(p)) for all p E C. By Theorem A.2 there is an element a E A such that 8 = a. If b = Re(a), then Re(8(p)) = Re(p(a)) = pCb) for all p E SeA). Since pet - b) > 0 for all pES, the hypothesis implies that t - b > O. Therefore, r(t - b) > 0, so t > r(b). But r(b) = Re(8(r)) > t, a contradiction. 0 If x is a unit vector in a Hilbert space H, we denote by W x the state B(H)  C, u  (u(x), x). 5.1.15. Theorem. Let A be a C*-algebra and suppose that (H)..,CP)..)..EA is a family of representations of A. Suppose also that p is a pure state of A such that n)..EA ker( cp)..) C ker(p). Then p belongs to the weak* closure in A * of the set S = {wxcp).. I A E A and x E H).., IIxll = I}. Proof. Replacing (H).., cp)..) by the canonically associated non-degenerate representation if necessary, we may suppose that each (H).., c.p)..) is non- degenerate. By passing to the quotient of A by the closed ideal n).. ker( c.p)..) if necessary, we may suppose also that n).. ker( cp)..) = o. In this case the 
5.2. The Transitivity Theorem 149 direct sum (H, c.p) of the representations (H A , c.pA) is faithful. Obse.rve that Wxc.p>.. is a state if IIxli = 1. Denote by f the unique state of A extending a state T of A and by c{; A the unique unital *-homomorphism from A to B(H>..) extending CP>... Then for T = wxc.p A E S we have f = wxc{; A. Suppose that A is non-unital and a E A, J.L E C, and a + J.L E nAEA ker( c{; A). Then for all b E A, we have ab + J.Lb = 0 because ab + J.Lb E nAEA ker( c.p A) = o. Thus, if J.L were nonzero, then -a/ J.L would be a unit for A, which contradicts our assumption. Hence, J.L = 0 and therefore a = o. Thus, if A is non-unital, n A ker( c{; A) = o. From these considerations it follows that to prove the theorem we may suppose that A is unital, replacing A by A, cp >.. by c{; A' and p by P if necessary (p is pure by Theorem 5.1.13). Suppose then that A is unital and that a is a self-adjoint element of A such that T( a) > 0 for all T E S. Then for each ,\ E A we have (cp A (a)( x), x) > 0 for all x E H A' and therefore c.p A( a) > o. Hence, c.p( a) > 0, so a > 0, as cp is an injective *-homomorphism. It now follows from Theorem 5.1.14 that the weak* closure of S contains PS(A). 0 5.2. The Transitivity Theorem The theorem of the title of this section enables us to relate some topological concepts to purely algebraic ones. For instance, we use it to show that for C*-algebras topological irreducibility of a representation is equivalent to algebraic irreducibility. We begin with an elementary result. 5.2.1. Lemma. Let H be a Hilbert space and el, . . . , en, Yl,... , Yn ele- ments of H, where el, . . . , en are orthonormal. Then there is an operator u E B(H) such that u(ej) = Yj (j = 1, . . . , n ) and lIuli <  max{IIYllI,..., llYn II}. Moreover, if there is a self-adjoint operator v on H such that v( ej) = Yj for j = 1, . . . , n, then we may choose u to be self-adjoint also. Proof. Set u = E7=1 Yj @ej. Clearly, u(ej) = Yj (j = 1,..., n). If x E H 
150 5. Representations of C*-Algebras and M = maxj IIYjll, then n Ilu(x)11 = II L(x,ej)yjll j=l n < L I(x, ej)IIIYjll j=l n n < (L I (x, ej) 1 2 L IIYj 112)1/2 j=l j=l < IlxIlVTiM, so lIuli < VTiM. Now suppose that there is a self-adjoint operator v on H such that for all j we have v(ej) = Yj. Then n n U = Lv(ej) 0 ej = v(Lej 0 ej) = vp, j=l j=l where p is the projection 2:7=1 ej Q9 ej. Because v is hermitian, so is u' = vp + pv - pvp, and clearly, u'( ej) = Yj for all j. Moreover, Ilu'1I 2 = Ilvp( vp)* + pv(l - p )(pv( 1 - p))* II < IIvpl12 + Ilpv(l _ p )11 2 < IIvpl12 + IIpvll 2 = 211u 11 2 < 2nM 2 , so Ilu'll < V2nM. o The following important result is called the transitivity theorem. 5.2.2. Theorem (Kadison). Let A be a non-zero C*-algebra acting irreducibly on a Hilbert space H, and suppose that Xl, . . . , X n and Yl, . . . , Yn are elements of H and that Xl, . . . , X n are linearly independent. Then there exists an operator u E A such that u(Xj) = Yj for j = 1,..., n. If there is a self-adjoint operator v on H such that v(Xj) = Yj for j = 1,..., n, then we may choose u to be self-adjoint also. If A contains id H and there is a unitary v on H such that v( X j) = Yj for j = 1, . . . , n, then we may choose u to be a unitary also-we may even suppose that u = e iw for some element w E Asa. Proof. Suppose first that there is a self-adjoint operator v on H such that v( x j) = Y j for all j, and we shall show that there exists U E Asa such 
5.2. The Transitivity Theorem 151 that u( x j) = Y j for all j. We may suppose that Xl, . . . , X n are orthonormal (because if we have proved the result in this case, and now suppose that Xl, . . . , X n are merely linearly independent, then we may choose an ortho- normal basis e1,. . . , en for K = CX1 + . . . + Cx n, and use the fact that there exists u E Asa such that u( ej) = v( ej) for all j, which implies u = v on K, to get u(Xj) = v(Xj) = Yj for all j). We may also suppose that maxj IIYj II < (2n )-1/2. Suppose that € > 0 and set Ut; = {u E B(H) I mx Ilu(Xj)11 < c:}, 11n so U e is a neighbourhood of 0 in the strong topology of B(H). Since A is strongly dense in B(H) by Theorem 4.1.12, it follows from the Kaplansky density theorem, Theorem 4.3.3, that the closed unit ball of Asa is strongly dense in the closed unit ball of B(H)sa. Hence, if w E B(H)sa, there is an element Wi E Asa such that Wi - w E U e and II Wi II < IIwll. By Lemma 5.2.1 there is an element Vo E B( H)sa such that vo( X j) = Yj for all j, and Ilvo II < 1. Hence, there is an element Uo E Asa such that Uo - Vo E U 1 /(2N) (where N = J2;1:) and Iluoll < 1. We now construct by induction two sequences of self-adjoint operators ( Uk) and (V k) on H, such that Uk E A and II u k II, II v k II < 2 - k, uk - V k E U 2 -Ie-l N-l, and for j = 1,. . . , n we have Vk(Xj) = (Vk-l - Uk-l)(Xj) (k > 0). Suppose that Uo,..., U r and Vo,..., V r have been constructed as above. Then, by Lemma 5.2.1 again, there is a self-adjoint operator V r +1 on H such that V r + 1 ( X j) = (V r - U r )( X j ) (j = 1, . . . , n ) and Ilv r +111 < N l!n Ilvr(xj) - ur(Xj)11 _1_ 1 1 < N ( ) - - - 2r+1N - 2 r + 1 ' so Ilv r +111 < 2- r - 1 . Hence, there exists an element U r +1 E Asa such that U r +1 - V r +1 E U 2 -r-2N-l and lIu r +111 < 2- r - 1 . This completes the induction. Since 2::0 Ilurll < 2::0 21r < 00, the series 2:r U r is convergent in A. Set U = 2::0 U r , so U E Asa. For j = 1, . . . , n we have r Y j - U ( X j) = lim (Y j - ""' Uk ( X j )) = lim v r+ 1 ( X j ), r-oo  r-oo k=O 
152 5. Representations of C*-Algebras since the sum telescopes. Since lim r --+ oo Vr+l (x j) == 0 for each j, we there- fore have Yj == u( x j). Now we return to the general case; that is, we drop the assumption that there is a hermitian operator v on H such that v(Xj) == Yj for j = 1,. .., n. However, we may retain the assumption that Xl,..., X n are orthonormal. By Lemma 5.2.1 there is a (possibly non-hermitian) operator v on H such that v( x j) = Y j for all j. If v' and v" are the real and imaginary parts of v, then there are hermitian elements u' and u" in A such that u'(Xj) = v'(Xj) and u" (x j) == v" (x j) for all j, by the first part of this proof. Thus, u == u' + iu" E A and u(Xj) = v(Xj) = Yj for all j. Finally we consider the case where A contains idH and there exists a unitary v on H such that v( x j) == Yj for all j. As before, we may suppose that Xl,..., X n are orthonormal. In this case Yl,..., Yn are also ortho- normal. Let 1< be the linear span of the vectors Xl, . . . , X n, Yl, . . . , Yn. Ex- tend Xl, . . . , X n (respectively, Yl, . . . , Yn) to an orthonormal basis Xl, . . . , X m (respectively, Yl,... ,Ym) of I<. Clearly, there is a unitary Vo E B(I{) such that Vo ( x j) == Y j for j == 1,..., m. Since Vo is diagonalisable (it is a normal operator on a finite-dimensional Hilbert space), there is an orthonormal basis el,..., em for 1< such that vo( ej) == Ajej for some AI'...' Am E C. Now IAjl = 1, so Aj = e itj for some tj E R. The operator w' == L:j:l tjej 0 ej on H is hermitian, and w'(ej) == tjej. Hence, by the first part of this proof there exists w E Asa such that w( e j) == t j e j for j == 1,..., m. Set u == e iw . Then u is a unitary in A such that u(ej) == e itj ej = Ajej == vo(ej), so u equals Vo on I{, and therefore u ( x j) = Vo ( x j) == Y j for j == 1,. . . , n. 0 We say a *-algebra A acting on a Hilbert space H is algebraically ir- reducible if 0 and H are the only vector subspaces (closed or not) of H that are invariant for A. Obviously, in this case A is topologically irreduc- ible, that is, irreducible in our previous meaning of this word. We say a representation (H, 'P) of a C* -algebra A is algebraically irreducible if 'P( A) is algebraically irreducible. Surprisingly, algebraic and topological irre- ducibility are the same, an important result in the representation theory of C* -algebras: 5.2.3. Theorem. Let (H, 'P) be a representation of a C*-algebra A. Then (H, 'P) is algebraically irreducible if and only if i t is topologically irreducible. Proof. Suppose that (H, 'P) is non-zero and topologically irreducible, and let !( be a non-zero vector subspace of H invariant for 'P(A). Let X be a non-zero element of !{ and Y an element of H. Then by Theorem 5.2.2 there exists u E A such that 'P( u)( x) == Y, so Y E 1<. Therefore, 1< == H 1- so (H, 'P) is algebraically irreducible. 0 5.2.4. Theorem. If p is a pure state on a C*-algebra A, then A/N p == Hp. 
5.3. Left Ideals of C*-Algebras 153 Proof. Of course the point here is that A/ N p is complete. Since (H p, cp p) is an irreducible representation of A by Theorem 5.1.6, and A/N p is a vector subspace of Hp invariant for <pp(A), we have A/N p = 0 or Hp, by the preceding theorem. Since N p =I A, therefore A/N p = Hp. 0 5.3. Left Ideals of C*-Algebras In this section we show that there is a bijective correspondence between the pure states and the modular maximal left ideals of a C*-algebra. This is used in the next section to analyse the primitive ideals of the algebra. We begin with a result on hereditary C*-algebras which has its own interest and a nice application in Remark 5.3.1. Moreover, using the cor- respondence between hereditary C*-subalgebras and closed left ideals, it translates immediately into a key result of this section concerning left ideals (Theorem 5.3.2). 5.3.1. Theorem. Let Bl and B 2 be hereditary C*-subalgebras of a C*-algebra A. Suppose that Bl C B 2 and that every positive linear func- tional r of A that vanishes on B} also vanishes on B 2 . Then Bl = B 2 . Proof. We may suppose that A is unital. Let a be a positive element of B 2 and suppose that e > o. Then the set F = {r E S(A) I Tea) > e} is weak* closed in the closed unit ball of A * , and therefore weak* compact by the Banach-Alaoglu theorem. If rEF, then r does not vanish everywhere on B 2 , so it does not vanish everywhere on B 1 , either. Choose aT E Bl such that r( aT) =I O. Then there is a weak* open set U T containing r such that p( aT) =I 0 for all p E U T. Clearly, the family (U T )TEF forms a weak* open cover of F, so by weak* compactness of F there are finitely many functionals rl,..., r n E F such that UTl U . . . U UTn contains F. Set b = E=l a;j a Tj . Then b E B 1 , and for any element rEF we have r(b) > 0 (since there exists some element a Tj such that Ir( a Tj ) I > 0, and r( b) > r( a;j a Tj ) > Ir( a Tj ) 1 2 ). Hence, the weak * continuous linear functional b: A *  C, r ....-..+ r ( b), is positive everywhere on F, so (again using weak* compactness of F) there is a positive number M such that r(b) > M for all rEF. Put c = (liall/M)b. Then e is a positive element of Bl and r( e) > lIall > r( a) (r E F). Now suppose that r is an arbitrary state of A. If r(a) < e, then r(e + e - a) > r(e - a) > o. If r(a) > £, then rEF and r(e + e - a) > r( a) + r( e - a) = e > o. This shows that for every positive linear functional r on A, r(e + e - a) > O. Hence, e + e - a > 0 by Theorem 3.4.3. We have therefore shown that for each £ > 0 there is an element e of Bt such that a < e + c. Because B} is hereditary in A, by Theorem 3.2.6 a E Bl. Consequently, Bt C B 1 , so B 2 C B 1 , and therefore B 2 = Bl. 0 
154 5. Representations of C*-Algebras 5.3.1. Remark. Let A be a C*-algebra and let a be a self-adjoint ele- ment of A such that r( a) > 0 for all non-zero positive linear functionals T on A. Then a is positive and (aAa)- = A. Positivity of a is given by Theorem 3.4.3. If the hereditary C*-subalgebra (aAa)- is not equal to A, then by Theorem 5.3.1 there is a non-zero positive linear functional r of A which vanishes on (aAa) -, and therefore r( a) = 0, contradicting our assumption. This shows that (aAa) - = A as asserted (cf. Exercise 3.5). 5.3.2. Theorem. Let L 1 and L 2 be closed left ideals of a C*-algebra A. Suppose that L 1 C L 2 and that every positive linear functional of A that vanishes on L 1 vanishes on L 2 . Then L 1 = L 2 . Proof. By Theorem 3.2.1 B 1 = L 1 n Lt and B 2 = L 2 n L; are hereditary C*-subalgebras of A. If r is a positive linear functional of A vanishing on B 1 , then it is clear from the inequality Ir(a)12 < Ilrllr(a*a) that r vanishes on L 1 . It follows from the hypothesis that r vanishes on L 2 , and therefore on B 2 . Hence, B 1 = B 2 by Theorem 5.3.1, and therefore L 1 = L 2 by Theorem 3.2.1 again. 0 5.3.3. Theorem. Let L be a proper closed left ideal in a C*-algebra A. Then the set R = {N p I p E PS(A) and L C N p } is non-empty and L = nR. Proof. Let S be the set of all norm-decreasing positive linear functionals on A, and let F be the set of all elements r of S such that L C NT. The functional 0 belongs to F, since No = A, so F is non-empty. Also, F is weak* closed, since it is the intersection of the weak* closed sets {r E S I r(a*a) = OJ, where a ranges over L. It follows from the Banach- Alaoglu theorem that F is weak* compact. It is easily checked that F is a face of S. By Theorem A.14 the set E of extreme points of F is non- empty and F is the weak* closed convex hull of E. By Theorem 5.1.8, E C {O} U PS(A). If E = {O}, then F = 0, so for all rES that vanish on L we have r = 0 (since in this case L C NT and therefore rEF). By Theorem 5.3.2 L equals A, contradicting the assumption that L is proper. This argument shows that E =I {O} and therefore E intersects PS(A), so R is non-empty. Now set L 1 = nR, so L 1 is a closed left ideal of A containing L. If r is a non-zero element of E, then NT contains L and therefore contains L 1 by the definition of R. Thus, if a ELI, then r(a*a) = 0 for all r E E, so r( a* a) = 0 for all rEF, as F is the weak* closed convex hull of E. Hence, every functional in F vanishes on L 1 . Since every functional in S vanishing on L is an element of F, we conclude from Theorem 5.3.2 that L = L 1 .----O If r is a positive linear functional on a C* -algebra A, then it is easily checked that the set NT + N; = {a + b* I a, b E NT} is contained in ker( r). 
5.3. Left Ideals of C*-Algebras 155 5.3.4. Theorem. If r is a state on a C*-algebra A, then r is pure if and only if ker( r) = NT + N: . Proof. If r is pure, then the representation (H T, CPT) is irreducible. Sup- pose that a is a hermitian element of ker( r) which is not in NT. Then a + NT and X T are orthogonal elements of H T (since (a + NT,X T ) = r(a) = 0), so if p is the projection of H T onto C(a + NT)' we have p(a + NT) = a + NT and p(x T ) = 0 + NT. By the transitivity theorem, Theorem 5.2.2, there is a hermitian element b E A such that cp( b)( a + NT) = a + NT and cp(b)(x T ) = 0 + NT. Hence, the elements c = ba - a and b belong to NT. Since a is self-adjoint, a = ba - c = a* b - c* E NT + N;. This shows that the hermitian elements of ker( r) belong to NT + N;, so ker( r) C NT + N: (since ker( r) is self-adjoint), and therefore ker( r) = NT + N;. Now suppose conversely that ker(r) = NT + N;. Suppose also that p is a positive linear functional on A majorised by r. Then NT C N p , and therefore ker(r) = NT + N; C N p + N; C ker(p). By elementary linear algebra, there is a scalar t such that p = tr. If p i= 0, then there exists a E A+ such that pea) > 0, and therefore t is a positive number. Moreover, if (U.x).xEA is an approximate unit for A, then t = Ilpll = limA p( u.x) < lim.x r( u.x) = 1, so t E [0,1]. This shows that r is a pure state of A. 0 5.3.5. Theorem. If A is a non-zero C*-algebra, then the correspondence r  NT is a bijection from PS( A) onto the set R of all modular maximal left ideals of A. Proof. Suppose that r, p E PS( A) are such that NT C N p. Then by Theorem 5.3.4 ker(r) = NT + N; C N p + N; = ker(p), so there is a scalar t such that p = tr. Obviously, t is positive and by equating norms we get t = 1, so p = r. This shows that the map p  N p is injective. If r E PS(A), then H T = A/NT by Theorem 5.2.4, so there is an element u E A such that x T = U + NT. Then for all a E A we have a + NT = 'PT(a)(x T ) = au + NT' so a - au E NT' and therefore NT is modular. Now suppose that L is a proper left ideal of A containing NT. Since L is modular (as NT is), L is a proper left ideal of A, by Remark 1.3.1. Hence, by Theorem 5.3.3 there is a pure state p of A such that L C N p . Therefore, NT C N p , so r = p by the first part of this proof. Hence, L = NT, and this shows that NT E R. Finally, suppose that L is an arbitrary element of R. Since by Re- mark 1.3.1 L is a proper closed left ideal of A, there is a pure state r of A such that L C NT' again using Theorem 5.3.3. By maximality of L, therefore, L = NT. 0 5.3.2. Remark. If A is a non-zero abelian C*-algebra, then for r E f2(A) = PS(A) we have NT = ker(r), so Theorem 5.3.5 asserts that the 
156 5. Representations of C*-Algebras correspondence T  ker( T) is a bijection from f2( A) onto the set of all modular maximal ideals of A. 5.4. Primitive Ideals For abelian C* -algebras the ideal structure is investigated in terms of the modular maximal ideals, that is, the kernels of the characters, and in terms of the topology of the spectrum. In the non-abelian case, the role of the modular maximal ideals is taken over by the primitive ideals. There are a number of candidates for the position of analogue of the character space. The most obvious of these is set of primitive ideals, which we endow with a suitable topology. Another analogue is obtained in terms of unitary equivalence classes of non-zero irreducible representations. We begin with a simple result from pure algebra that allows us to define primitive ideals. 5.4.1. Theorem. Let L be a modular left ideal in an algebra A. Then there is a largest ideal I of A contained in L, namely I = {a E A I aA C L}. Proof. It is clear that I = {a E A I aA C L} is an ideal of A. Since L is modular, there is an element u E A such that a - au E L for all a E A. If a E I, then au and a - au E L, so a E L. Therefore, I C L. If J is an ideal of A contained in L, then for all a E J we have aA C J C L, so J C I. 0 If L is a modular maximal left ideal in an algebra A, we call the ideal I in Theorem 5.4.1 the primitive ideal of A associated to L. We denote by Prim(A) the set of primitive ideals of A. 5.4.1. Remark. If T is a pure state on a C*-algebra A, then ker( CPT) is the ideal associated to the modular left ideal NT, as in Theorem 5.4.1, since ker(CPT) = {a E A I CPT(a)(A/N T ) = O} = {a E A I aA C NT}. 5.4.2. Theorem. An ideal I of a C*-algebra A is primitive if and only if there exists a non-zero irreducible representation (H, cp) of A such that I = ker( cP ). Proof. If I is primitive, then there is a modular maximal left ideal of A to which I is associated, and by Theorem 5.3.5 this left ideal is of the form N p for some p E PS(A). Hence, I = ker(cpp) by Remark 5.4.1. Since (Hp,cpp) is a non-zero irreducible representation of A, the forward implication of the theorem is proved. 
5.4. Primitive Ideals 157 Now suppose conversely that I = ker(cp), where (H,cp) is some non- zero irreducible representation of A. By Theorem 5.1.5 (H, c.p) has a unit cyclic vector, x say. The function p: A -+ C, a  (cp(a)(x), x), is a pure state on A and the representations (H, cp) and (H p, cp p) are uni- tarily equivalent by Theorem 5.1.7. Hence, I = ker( c.p p) is the primitive ideal associated to the modular maximal left ideal N p . 0 If S is a subset of a C*-algebra A, we let hull(S) denote the set of primitive ideals of A containing S. If R is a non-empty set of primitive ideals of A, we denote by ker(R) the intersection of the ideals in R. We set ker(0) = A. 5.4.3. Theorem. If I is a proper modular ideal of a C*-algebra A, then hull(I) is non-empty. If I is a proper closed ideal in A, then I = ker(hull(I)); that is, I is the intersection of the primitive ideals that contain it. Proof. If I is a proper modular ideal of A, an application of Zorn's lemma shows that there is a modular maximal left ideal L of A containing I. If J is the associated primitive ideal, then I C J, since I C L. Therefore, J E hull(I), so hull( I) :F 0. Now suppose that I is a proper closed ideal of A. By Theorems 5.3.3 and 5.3.5, the set R of modular maximal left ideals of A that contain I is non-empty and I = nR. If L is a modular maximal left ideal of A with associated primitive ideal J, then I C J if and only if I C L. Hence, hull(I) :F 0, and I C ker(hull(I)) C nR = I, so I = ker(hull(I)). 0 It follows from Theorem 5.4.3 that a modular maximal ideal of a C*-algebra is primitive. For abelian C*-algebras the two concepts coincide. 5.4.4. Theorem. Let A be an abelian C*-algebra and I an ideal of A. Then I is primitive if and only if it is modular maximal. Proof. Suppose that I is primitive. By Remark 5.4.1 and Theorem 5.3.5, there exists p E PS(A) such that I = ker(c.pp). By Theorem 5.1.6 p is a character on A, so N p = ker(p), and therefore I = ker(p). Hence, I is a modular maximal ideal of A (cf. Remark 5.3.2). 0 A C* -algebra A is primitive if its zero ideal is primitive, that is, if A has a faithful non-zero irreducible representation. The primitive C*-algebras, as with the simple C*-algebras, are thought of as the basic building blocks 
158 5. Representations of C*-Algebras of the theory, and it is important to construct examples of these algebras. We present a few here. More will be given at various points as we proceed. If H is a non-zero Hilbert space, then the identity representation of B(H) on H is irreducible by Theorem 4.1.12, since B(H) is strongly dense in itself. Therefore, B(H) is primitive. The Toeplitz algebra A is primitive because it acts irreducibly on the Hardy space H 2 by Theorem 3.5.5. Since every non-zero C*-algebra admits a pure state and therefore a non-zero irreducible representation, it follows that every non-zero simple C*-algebra is primitive, because the kernel of every non-zero irreducible representation is the zero ideal in this case. Not all primitive C*-algebras are simple. An easy counterexample is provided by B(H), where H is an infinite-dimensional Hilbert space. An abelian C*-algebra is almost never primitive. To be precise, a non-zero abelian C*-algebra A is primitive if and only if A = C. The backward implication is trivial. To see the forward implication, suppose that A admits a faithful irreducible representation (H, cp). Since A is abelian, cp(A) C cp(A)', and since (H,cp) is irreducible, cp(A)' = C1 by Theorem 5.1.5. Therefore, cp(A) = C1, so A = C. 5.4.2. Remark. A closed ideal I in a C*-algebra A is prime if whenever J 1 and J 2 are closed ideals of A such that J 1 J 2 C I, we necessarily have J 1 C I or J 2 C I. If I is a prime ideal in A, and S1, S2 are subsets of A such that S1AS2 is contained in I, then S1 or S2 is contained in I. To see this, set J 1 = AS 1 A and J 2 = AS 2 A. Then J 1 , J 2 are closed ideals in A such that J 1 J 2 C I, so J 1 or J 2 is contained in I (because I is prime). But Sj C ASjA (this follows from the existence of an approximate unit for A), so S1 or S2 is contained in I. We say that A is a prime C*-algebra if the zero ideal of A is prime. Equivalently, every pair of non-zero closed ideals of A has non-zero inter- section (the equivalence holds because InJ = IJ for all closed ideals I, J of A). It is immediate from the following theorem that a primitive C*-algebra . . IS prIme. 5.4.5. Theorem. If I is a primitive ideal of a C*-algebra A, then I is prlme. Proof. First, suppose that p E PS(A), and let L 1 ,L 2 be left ideals of A such that L 1 L 2 C N p . We claim that L 1 or L 2 is contained in N p . For suppose that L 2  N p . Then cpp(L 2 )x p is a non-zero vector subspace of Hp invariant for (Hp, cpp), and therefore by (algebraic) irreducibility of this representation, cpp(L 2 )x p = Hp. Hence, x p = cpp(a)(xp) for some element a E L 2 . If b is an arbitrary element of L 1 , then cpp(b)(xp) = cpp(ba)(xp) = 
5.4. Primitive Ideals 159 ba + N p = 0 + N p , since L 1 L 2 C N p . Hence, b E N p . This shows that L 1 C N p . Now suppose that J 1 and J 2 are ideals of A such that J 1 J 2 C I. There is a pure state p of A such that I = ker( 'P p), so J 1 J 2 C N p, and therefore, by what we have just shown, J 1 C N p or J 2 C N p . Hence, J 1 or J 2 is contained in I. 0 Just as we put a topology on the character space of an abelian Banach algebra, we now endow the set of primitive ideals of a C*-algebra with a topology that reflects the ideal structure of the algebra. 5.4.6. Theorem. If A is a C*-algebra, then there is a unique topology on Prim(A) such that for each subset R the set hull(ker(R)) is the closure R. Proof. If R C Prim(A), set R' = hull(ker(R)). Clearly, R C R' and ker(R) = ker(R'), so R' = (R')'. Also, 0' = 0. If R 1 and R 2 are subsets of Prim(A), then (R 1 U R 2 )' = R U R. To show this we may suppose that R 1 and R 2 are non-empty. Set II = nR I and I 2 = nR 2 . Then n(R I UR 2 ) = II nI 2 = 1 1 1 2 , so for any ideal I E Prim(A), we have I E (R 1 U R 2 )' <=> I 1 I 2 C I <=> II C I or 1 2 C I (since 1 is prime by Theorem 5.4.5) {:} I E R U R. Thus, (R 1 U R 2 )' = R U m, as asserted. If (R)EA is an arbitrary family of subsets of Prim(A), then (nEAR)' = nEAR. This is so because for each index I" E A, nEAR C R, so (nEAR)' C R, and therefore (nEAR)' C nEAR, and the reverse inclusion is obvious. Now define T to be the collection of all sets Prim(A) \ R', where R ranges over all subsets of Prim(A). From what we have just shown, it is easily checked that T is a topology on Prim(A), for which R' is the closure of R for each R C Prim(A), and there can be only one such topology. 0 The topology on Prim(A) in Theorem 5.4.6 is called the Jacobson or hull-kernel topology on Prim( A). Recall that a To -space is a topological space n such that for every pair of distinct points of n there is an open set containing one of the points and not the other. 5.4.7. Theorem. Let A be a C*-algebra. (1) Prime A) is a To -space. (2) The correspondence I  hull(I) is a bijection from the set of closed ideals of A onto the set of closed subsets ofPrim(A). (3) If II, I 2 are closed ideals of A, then II C 1 2 if and only if hull ( 1 2 ) C hull ( II). 
160 5. Representations of C*-Algebras Proof. If I is a closed ideal of A, then I = ker(hull(I)) by Theorem 5.4.3, so hull(I) = hull(ker(hull(I))). Hence, hull(I) is closed in Prim(A). Con- ditions (2) and (3) are immediate from these observations. If II and 1 2 are distinct points of Prim(A), then one is not contained in the other; say II C1 1 2 . Hence, 1 2 rt hull ( II), and since hull ( II) is closed, this shows that Prim(A) is a To-space; that is, Condition (1) holds. 0 5.4.8. Theorem. If A is a unital C*-algebra, then Prim(A) is compact. Proof. Let (CA)AEA be a family of closed sets in Prim(A) with the finite intersection property, and for each index A, let I A be the closed ideal of A such that hull(I A ) = CA. For each non-empty finite subset F = {AI, . . . , An} of A, the ideal IF = IAl + . . . + IAn is proper because C Al n . . . n CAn is non-empty. Hence, 1 f/. IF and the ideal J = UFI F (where F ranges over all non-empty finite subsets of A) is proper. It follows from Theorem 5.4.3 that there is a primitive ideal I containing J. Hence, I A C I, so I E C A, for all A E A. Thus, nAEAC A is non-empty. Therefore, the space Prim(A) is compact. 0 The converse of the preceding theorem is false. An easy counter- example is provided by I«H), where H is an infinite-dimensional Hilbert space. In this case the primitive ideal space is a point space (since K(H) is simple), but I«H) is non-tUlital. We introduce now another topological space which is also an analogue of the character space of an abelian algebra: If A is a non-zero C*-algebra, we denote by A the set of unitary equivalence classes of non-zero irreducible representations of A. If (H, cp) is a non-zero irreducible representation of A, we denote its equivalence class in A by [H, cp], and we set ker[H, cp] = ker( cp). The surjection (J: A  Prim(A), [H, cp]  ker[H, cp], is the canonical map from A to Prim(A). We endow A with the weakest topology making () continuous, and call the topological space A the spectrum of A. By elementary topology () is open and closed. If R is a subset of A, let hull'(R) = (J-l(hull(R)). Then the correspondence I  hull'(I) is a bijection from the closed ideals of A onto the closed subsets of A. ote that if A is a unital C*-algebra, then it follows from Theorem 5.4.8 that A is compact. The proof of the following is a short, easy exercise: 5.4.9. Theorem. Let A be a C*-algebra. The following conditions are equivalent: (1) A is a To-space. 
5.4. Primitive Ideals 161 (2) Any two non-zero irreducible representations of A with the same kernel are unitarily equivalent. (3) The canonical map A  Prim(A) is a homeomorphism. We have seen that the closed ideals of a C*-algebra A correspond to the closed subsets of the primitive ideal space and of the spectrum. They also correspond to certain subsets of PS(A), a result that is very useful, as we shall see in the theory of tensor products (Chapter 6). Let A be a C*-algebra and p E PS(A). If u is a unitary in A, then the linear functional p1t: A  C, a  p(uau*), is a pure state of A. We say that a subset S of PS(A) is unitarily invariant if whenever pES and u is a unitary of A we have pU E S. In this case we set S1. = {a E A I pea) = 0 (p E S)}. If [ is a closed ideal in A, we set [1. = {p E PS(A) I p(a) = 0 (a E I)}. It is clear that [1. is a unitarily invariant weak* closed subset of PS(A), and it follows from the next theorem that S1. is a closed ideal in A. 5.4.10. Theorem. Let A be a C *-alge bra. (1) If S is a non-empty unitarily invariant subset ofPS(A), then S1. = npES ker( cp p). Ifin addition S is relatively weak* closed in PS(A), then S = (S1.)1.. (2) If I is a closed ideal of A, then I = ([1.)1.. (3) The map S  S1. from the set of relatively weak* closed unitarily in- variant subsets S ofPS(A) to the set of closed ideals of A is a bijection. Proof. Suppose that S is a non-empty unitarily invariant subset of PS( A). Ifa E npEsker(cpp), thenforeachp E Swehavep(a) = (<pp(a)(xp),xp) = 0, so a E S1.. Thus, npES ker( cpp) C S1.. If these sets are not equal, then there is an element a E S1. such that for some pES we have <p p( a) i= o. Hence, there exists a unit vector x E H p such that (cp p( a)( x), x) i= O. Since p is pure, (H p, cp p) is an irred uci ble representation, and therefore if cp denotes the unique unital *-homomorphism from A to B(Hp) extending CPP' the representation (Hp, cp) of A is also irreducible. It follows from Theorem 5.2.2 that there exists a unitary u in A such that <p(u)x = xp' Now pU(a) = (cpp(uau*)(xp),x p ) = (cpp(a)cp(u*)(xp),CP(u*)(x p )) = (cpp(a)x,x) i= 0, 
162 5. Representations of C*-Algebras so pU ft S, contradicting the unitary invariance of S. We therefore conclude that npES ker( c.p p) = S.1.. Now suppose that in addition S is relatively weak* closed in PS(A). It is clear that S C (S.1.).1.. To prove the reverse inclusion, suppose that T E (S.1.).1.. By applying Theorem 5.1.15 to the family (Hp,c.pp)pES of representations of A, we see that T is a weak* limit of states of the form wxc.pp with pES and x a unit vector of Hp-the symbol W x denotes the state B(H p )  C, u  (u(x),x). It is clear from the first part of this proof that for any such state Wxc.p p, there is a unitary u in A such that wxc.pp = pU, and therefore wxc.pp E S by unitary invariance of S. Hence, T E S, since S is relatively weak* closed in PS( A). Therefore, S = (S.1.).1.. Let I be a proper closed ideal of A. By Theorem 5.3.3 I = npEsN p , where S = {p E PS(A) I I C N p }. The set S is non-empty and clearly weak* closed in PS( A) and unitarily invariant. Moreover, since ker( c.p p) is the largest ideal contained in N p (cf. Remark 5.4.1), we have I C N p if and only if I C ker( c.p p), and therefore I = n pE s ker( c.p p) = S.1. . It follows that I.1. = (S.1.).L = S, so (I.1.).L = S.L = I. This proves the theorem. 0 5.5. Extensions and Restrictions of Representations As is indicated by the title, we are concerned in this section with extending and restricting representations, usually with the aim of getting representations of the same type as the ones with which we started. We also investigate the relationships between the primitive ideal space and the spectrum of a C*-algebra and the corresponding spaces for its hereditary C*-subalgebras and quotient algebras. Let (H, c.p) be a representation of a C* -algebra A, and suppose that B is a C*-subalgebra of A and that I{ is a closed vector subspace of H invariant for c.p( B). Then the map 1/;: B  B(I{), b  c.p(b)K, is a *-homomorphism. We denote the representation (K,1/J) by (H,c.p)B,K. 5.5.1. Theorem. Let B be a C*-subalgebra of a C*-algebra A and sup- pose that (K, 1/;) is a non-degenerate representation of B. Then there is a non-degenerate representation (H, c.p) of A and a closed vector subspace K' of H invariant for c.p( B) such that (K, 1/;) is unitarily equivalent to (H, c.p )B,KI. If(K, 1/;) is cyclic (respectively irreducible), we may take (H, c.p) cyclic (respectively irreducible). 
5.5. Extensions and Restrictions of Representations 163 Proof. We may assume that 'ljJ is non-zero. First suppose that (K,,,p) is cyclic, and let y be a unit cyclic vector. The function TO: B  C, b  ("p(b)(y), y}, is a state of B, and therefore extends to a state T of A, by Theorem 3.3.8. Set (H,cp) = (HT,CPT). If (U.x).xEA is an approximate unit for B, then the net of positive operators (cp( u.x)) is increasing and bounded above, and therefore by Theorem 4.1.1 it converges strongly to a positive operator, p say, on H. Clearly, pcp(b) = cp(b)p = cp(b) for all b E B. Let x = p(x T ) and K' = [<p(B)x]. Then K' is a closed vector subspace of H invariant for cp(B). Set (I<', 'ljJ') = (H, cp )B,KI. Then x is a cyclic vector for the representation (K' , "p') (x E I{', since x = p( X T) = lim.x cp( u.x)( x)). For each bE B, we have ('ljJ'(b)(x),x) = (cp(b)(XT),X T ) = T(b) = To(b) = ('ljJ(b)(y), y). Hence, by Theorem 5.1.4 the representations (K,,,p) and (K', "p') are uni- tarily equivalent. Now suppose that (I{,,,p) is irreducible. Then TO is a pure state by Theorem 5.1.7. Applying Theorem 5.1.13, we may suppose that T is pure also. It then follows that (H, cp) = (H T, CPT) is irreducible. Finally, suppose only that (I{,,,p) is non-degenerate. In this case, by Theorem 5 .1. 3, we can write (I{, "p) as a direct sum of a family (( K .x, "p .x) ) .x of cyclic representations of B. For each index ,,\ there is a cyclic represent- ation (H.x, cp.x) of A, and a closed vector subspace K of H.x invariant for <p.x(B) such that (1{.x,,,p.x) is unitarily equivalent to (H.x,cp.x)B,K. Hence, if (H, cp) is the direct sum of the representations (H.x, cp .x), it is a non- degenerate representation of A, and the orthogonal direct sum K' of the spaces K is a closed vector subspace of H invariant for cp(B). It is easily checked that (1<,,,p) is unitarily equivalent to (H,cp)B,KI. 0 If (H, <p) is a representation for a C*-algebra A, and B a C*-subalgebra of A, we denote the representation (H, cp )B,[<p(B)H] by (H, cp )B, and call it the restriction of the representation (H, cp) to B. 5.5.2. Theorem. Let B be a hereditary C*-subalgebra of a C*-algebra A, and suppose that (H, cp) is an irreducible representation of A. Then (H,cp)B is an irreducible representation of B. Moreover, cp(B)H is closed. Proof. We may suppose that cp # O. Let p be the orthogonal projection of H onto [<p(B)H], and let (U.x).xEA be an approximate unit for B. Then it is easily checked that (cp( u.x)) converges strongly to p on H. Suppose that x and yare elements of p( H) and that x is non-zero. By (algebraic) irreducibility of (H,cp), there is an element a E A such that cp(a)(x) = y. Set b.x = U.x au.x, so b.x E B (because B is hereditary in A). Then, since IIcp(b.x)(x) - yll < Ilcp(u.\au.x)(x) - cp(u.x)(y)1I + IIcp(u.x)(y) - yll < 11<p( a )<p( u.x)x - yll + 1I<p( u.x)y - yll, 
164 5. Representations of C*-Algebras and limA <p(UA)(X) = x, and limA <p(uA)(y) = Y, therefore limA <p(bA)(x) = y. Hence, y E [<p(B)x]. This argument shows that (H,<p)B is (topologically) irreducible. By Theorem 5.2.3 (H, <p) B is therefore algebraically irreducible, and since <p(B)H is an invariant vector space for (H, <p )B, it is equal to [<p(B)H]. 0 5.5.3. Corollary. Let B be a non-zero hereditary C*-subalgebra of a C*-algebra A, and p a pure state of A. Then there exists t E [0,1] and there is a pure state p' of B such that p B = t p', where p B is the restriction of p to B. Proof. We may suppose that PB =I O. Let p be the projection of H onto <pp(B)H, so p E <pp(B)', since <pp(B)H is invariant for <pp(B). If Y = p(x p ), then for all b E B we have p(b) = (<pp(b)(xp),x p ) = (p<pp(b)(xp),x p ) = (<pp(b)(y), y). Hence, the irreducible representation (H, <pp)B is non-zero, and Y is also non-zero. By Theorem 5.1.7 the function p': B -+ C, b  (<pp(b)y, y) IlIy112, is a pure state of B, since y I II y II is a unit cyclic vector for (H, <p p) B. Clearly, PB = tp', where t = Ily112. Since lIyll = IIp(xp)1I < IIxpll = 1, therefore t E [0, 1]. 0 5.5.4. Lemma. Suppose that B is a non-zero hereditary C*-subalgebra of a C*-algebra A, and that I is a primitive ideal of A not containing B. Then I n B is a primitive ideal of B. Moreover, if J is a closed ideal of A such that J n B C I, then J C I. Proof. The characterisation of primitive ideals given in Theorem 5.4.2 implies that I = ker( <p) for some non-zero irreducible representation (H, <p) of A. N ow I n B is the kernel of the representation (H, <p ) B of B, and this representation is non-zero because B  I. Since (H, <p)B is irreducible (Theorem 5.5.2), we can apply Theorem 5.4.2 again to infer that I n B is a primitive ideal in B. Suppose now that J is a closed ideal of A such that J n B C I, and we shall show that J C I. Because B is hereditary in A, we have BAB C B; hence, B J B C J n B C I, and therefore (B J)A( J B) C I. Since I is prime (Theorem 5.4.5), it follows from Remark 5.4.2 that one of the sets BJ or J B is contained in I. Hence, BAJ or JAB is contained in I, so again applying Remark 5.4.2, B or J is contained in I. The containment B C I is ruled out by hypothesis, so J C I. 0 If B is a non-zero hereditary C*-subalgebra of a C*-algebra A, we call the map Prim(A) \ hull(B) -+ Prim(B), I  In B, 
5.5. Extensions and Restrictions of Representations 165 the canonical map from Prim(A) \ hull(B) to Prim(B). Similarly, the map  ,  A \ hull (B) -+ B, [H, <p]  [(H, 4' )B],  ,  is the canonical map from A \ hull (B) to B. (By Theorem 5.4.3, the intersection of the primitive ideals of A is the zero ideal. Hence, hull ( B) =I ,  Prim(A) and hull (B) =I A.) 5.5.5. Theorem. Suppose that B is a non-zero hereditary C*-subalgebra of a C*-algebra A. Then the following diagram commutes where the maps are the canonical ones: A \ hull'(B) 1 Prim(A) \ hull(B)  -+ B 1 -+ Prim(B). Moreover, the horizontal maps are homeomorphisms. Proof. Commutativity of the diagram is clear. Denote by (J and (J' the upper and lower horizontal maps, respectively. First we show that 8 is injective. Suppose that 8[HI' <PI] == 8[H 2 ,4'2]' If (Kj,1/Jj) == (Hj,<pj)B (j == 1,2), then (K I ,1fJI) and (K 2 ,1fJ2) are non-zero unitarily equivalent irreducible representations of B. Let u: KI -+ K 2 be a unitary such that 1/J2(b) == u1/JI(b)u* (b E B), and choose unit vectors XI,X2 of I<I,K 2 , respectively, such that X2 == U(XI). Then <PI(A)XI == HI and <P2(A)X2 == H 2 (because <PI and <P2 are algebraically irreducible). Theorem 5.1.7 implies that the functions Pj: A -+ C, a  (<pj(a)(xj),Xj), (j == 1,2) are pure states of A such that (Hj, <P j) is unitarily equivalent to (H Pi , <P Pi ) (j == 1,2). Moreover, Pj extends the function Pj: B -+ C, b  (1/Jj(b)(Xj),Xj), which is a pure state of B, since (I{ j, 1/J j) is irred uci ble. However, for each b E B we have p(b) == (1/J2(b)u(XI)' U(X1)) == (1/JI(b)(XI),XI) == p(b), so P == p;. It follows from Theorem 3.3.9 that P2 == PI, and therefore (H2, <P2) and (HI, <PI) are unitarily equivalent. Therefore, 8 is injective. To show surjectivity of (J, suppose that [K,1/J] E E. By Theorem 5.5.1 there is an irreducible representation (H, <p) of A and a closed vector sub- space K' of H invariant for <p(B) such that (K, 1/J) and (H, <P )B,K' are uni- tarily equivalent. Since (H, 'P) is clearly non-zero, [H, <p] E A. Choose x in 
166 5. Representations of C*-Algebras K' such that c.p(B)x = K' (this is possible, since (H, c.p )B,K' is irreducible). Then there is an element b o E B such that x = c.p(bo)(x). Now c.p(A)x = H, as (H, c.p) is irreducible, so for any y E Hand b E B there exists an element a E A such that c.p(b)(y) = c.p(b)c.p(a)(x) = c.p(babo)(x) E c.p(B)x. This shows that c.p(B)H = K'. Therefore, (H, c.p )B,K' = (H, c.p )B, so [H, c.p]  hull'(B) and 8[ H, c.p] = [K, 'l/J]. We have therefore shown that (J is a bijection. It follows directly from commutativity of the diagram in the statement of the theorem that 8' is surjective, and injectivity of 8' is immediate from Lemma 5.5.4. We now show (J' is a homeomorphism (from which it is an elementary exercise to show that (J is also a homeomorphism). Suppose that C is a non-empty closed set of Prim(B). Then C = hullB(I) for some closed ideal I of B (the subscript indicates we are looking at the hull relative to B). Now ((J')-l(C) = {J E Prim(A) I I C J n Band B Cl J} = hullA(I) n (Prim(A) \ hullA(B)), so (8')-1(C) is closed in Prim(A) \ hullA(B). Therefore, (J' is continuous. Now we show that 0' is a closed map. Suppose that C is a non-empty closed set in Prim(A) \ hullA(B), so C = hullA(I) \ hullA(B) for some proper closed ideal I of A. If J E hullB(I n B), then J = J' n B for some J' E Prim(A) \ hullA(B), because 8' is surjective. Hence, I n B C J' . It follows from Lemma 5.5.4 that I C J'. Therefore, J = (J' ( J') E (J'( C). Hence, hullB(I n B) C (J'( C), and the reverse inclusion is obvious, so hullB(I n B) = (J'( C). Consequently, (J' is a closed map. 0 5.5.6. Corollary. Suppose that B is a non-zero hereditary C*-subalgebra of a primitive C*-algebra A. Then B also is primitive. Proof. It follows immediately from Theorem 5.5.5 that since 0 E Prim(A)\ hull(B), therefore 0 E Prim(B). 0 If the assumption in Corollary 5.5.6 that B is hereditary is dropped, then the conclusion may fail. For instance, if H is a Hilbert space of dimension greater than 1, then the primitive C*-algebra B(H) contains non-trivial, and therefore non-primitive, abelian C*-subalgebras. Let 'l/J: A  B be a surjective *-homomorphism of C*-algebras, and suppose that I = ker( 'l/J) . If (H, c.p) is a representation of B, then (H, c.p'l/J ) is a representation of A, and clearly (H, c.p'l/J) is irreducible if (H, c.p) is ir- reducible. If B :F 0, we therefore have a well-defined map '" , B  hull (I), [H, c.p]  [H, c.p1/J], 
5.6. Liminal and Post liminal C* - Algebras 167 which we call the canonical map from iJ to hull' (1). Note that ker( <.p) = -l(ker( <.p )). It follows from Theorem 5.4.2 that if J E Prim(B), then -l(J) E hull(1). The map Prim(B)  hull(I), J  -l(J), is the canonical map from Prim(B) to hull(I). It is an easy exercise to show that these two canonical maps are bijective. 5.5.7. Theorem. Suppose that the map : A  B is a non-zero surjective *-homomorphism from the C*-algebra A onto the C*-algebra B, and sup- pose that I = ker( ). Then the following diagram commutes, where all the maps are canonical: '" B 1 Prim(B)   hull'(1) 1 hull ( I). Moreover, the horizontal maps are homeomorphisms. Proof. We shall show only that the second horizontal map, which we denote by 8, is a homeomorphism, because the rest is then routine. To show that 8 is continuous, let C be a closed set in hull(I). For some closed ideal 10 of A containing I, we have C = hullA(1o), so 8- 1 (C) = {J E Prim(B) I Io C -I(J)} = {J E Prim(B) I (Io) C J}, and therefore 8- 1 (C) is closed in Prim(B). Hence, 8 is continuous. To show that 8 is a closed map, suppose that D is a closed set of Prim(B). Then there is a closed ideal J o in B such that D = hullB(J o ). If J E hullA(-l(Jo)), then 1 C J, so J = -l(J') for some primitive ideal J' of B (because 8 is bijective). Clearly, J o C J', so J = 8(J') E 8(D), and therefore hull A (  -1 ( J 0)) C 8( D). Since the reverse inclusion also holds, we have hullA(-l(Jo)) = 8(D), so 8(D) is closed in hull(I), and therefore 8 is a closed map. 0 5.6. Liminal and Post liminal C*-Algebras The algebras of the title of this section form the best-behaved class of C*-algebras. Their theory is deep and very well-developed, but we shall have space here only to touch upon the elements of the subject. A C*-algebra A is said to be liminal if for every non-zero irreduc- ible representation (H,<.p) of A we have <.p(A) = 1{(H) (equivalently, in- voking Theorem 2.4.9, <.p(A) C 1{(H)). Liminal algebras are also called CCR algebras. CCR is an acronym for completely continuous representa- tions (an older terminology for a compact operator is a completely contin- uous operator). 
168 5. Representations of C*-Algebras 5.6.1. Eample. If A is an abelian C*-algebra, then it is liminal. For if (H, cp) is a non-zero irredutible representation of A, then cp(A)' = C1, and cp(A) C cp(A)', since A is abelian. Hence, cp(A) = C1, so H is one-dimensional, since cp( A) has no non-trivial invariant vector subspaces. Therefore, cp(A) = I{(H). 5.6.2. Eample. Every finite-dimensional C*-algebra A is liminal. For if (H, cp) is a non-zero irreducible representation of A, then H = cp(A)x for some non-zero vector x E H, so H is finite-dimensional and therefore cp(A) C I«H) = B(H). 5.6.3. Eample. If H is a Hilbert space, I{(H) is a liminal C*-algebra. This follows immediately from Example 5.1.1, where it is shown that every non-zero irreducible representation of I{(H) is unitarily equivalent to the identity representation of I{(H) on H. Not every C*-algebra is liminal. The algebra B(H) for H infinite- dimensional provides an easy counterexample (consider the identity repre- sentation of B(H) on H). 5.6.1. Theorem. If A is a liminal C *-alge bra, then its C*-subalgebras and its quotient C*-algebras are li1ninal also. Proof. Suppose that B is a C* -subalgebra of A. If (I{, 'ljJ) is a non-zero irreducible representation of B, then by Theorem 5.5.1 there is an irreduc- ible representation (H,cp) of A and a closed vector subspace 1(' of H such that (1<, 'ljJ) is unitarily equivalent to (I{', 'ljJ') = (H, cp )B,K'. Clearly, cp =I 0 because 'ljJ =I O. By hypothesis, cp(A) = I{(H), and therefore 'ljJ'(b) is a compact operator for all b E B (restrictions of compact operators are com- pact). Therefore, 'ljJ( B) consists of compact operators on I{. Consequently, B is liminal. Now let C be a quotient C*-algebra of A, and let 7r: A  C be the quotient *-homomorphism. Let (H, cp) be a non-zero irreducible represent- ation of C. Then (H, cp7r) is a non-zero irreducible representation of A, and therefore cp7r(A) = I{(H) from the hypothesis; that is, cp(C) = I«(H). Therefore, C is liminal. 0 If I is a closed ideal in a liminal C*-algebra A, then it follows from Theorem 5.6.1 that I and AI I are both liminal. The converse is false. To see this, let us first observe that a unital liminal C*-algebra A has only finite-dimensional irreducible representations. For if (H, cp) is a non-zero irreducible representation of A, then it is non-degenerate, and therefore cp(l) = id H . Hence, idH is compact, and therefore dim(H) < 00. Now if H is any infinite-dimensional Hilbert space, then the C*-algebra A = I«H) + C1 is unital and has an infinite-dimensional non-zero irreducible 
5.6. Liminal and Postliminal C*-Algebras 169 representation, namely, the identity representation on H. Hence, A is not liminal. But [{(H) is a liminal ideal of A such that AI K(H) == C is liminal. A C* -algebra A is said to be postliminal if for every non-zero irreduc- ible representation (H,<p) of A we have K(H) C <p(A) (equivalently, by Theorem 2.4.9, 1{(H) n <peA) =I 0). The postliminal C*-algebras are also called GCR algebras and Type I C*-algebras. Unfortunately, the Type I terminology conflicts with the terminology for von Neumann algebras be- cause Type I von Neumann algebras are not Type I C*-algebras in general. Every liminal C* -algebra is obviously postliminal. 5.6.2. Theorem. Let I be a closed ideal in a C*-algebra A. Then A is postliminal if and only if I and AI I are postliminal. Proof. First suppose that A is postliminal. If ([{,,,p) is a non-zero ir- reducible representation of I, then by Theorem 5.5.1 there is an irreducible representation (H, 'P) of A and a closed vector su bspace K' of H invariant for 'P(I) such that ([{,,,p) is unitarily equivalent to (I<',1/J') == (H,'P)I,KI. Clearly, 'P =I 0, since"p =I O. Since A is postliminal, I«H) C 'P(A). Observe that if u is a compact operator on I{I, then it is the restriction of a com- pact operator v on H (for example, take v == u on I<' and v == 0 on K'J...). Suppose now that x is a non-zero element of I{' and let u be the projection of K ' onto Cx. Since (I{I, "p') is algebraically irreducible and "p'(I)x is a non-zero vector subspace of l{' invariant for "p' (I), we have I{' == "p' (I)x. Hence, x == 1/;' (b)x for some bEl. Moreover, since v is compact, we have v == 'P( a) for some a E A, because A is postliminal. Consequently, u == "p'(b)u == "p'(b)'P(a)K' == 'P(ba)K' == "p'(ba) (as ba E I). Therefore, "p' (I) contains the rank-one projections of K ' , and therefore the compact operators on [<'. Since ([{,,,p) and ([{',,,p') are unitarily equivalent, "p( I) contains the compact operators on [{. This shows that I is postliminal. The proof that AI I is also postliminal when A is postliminal is completely straightforward and left as an exercise. Now we suppose that I and AI [ are postliminal and we show that A is also. Let (H, 'P) be a non-zero irreducible representation of A. If ker( 'P) contains I, then there is a *-homomorphism "p: AI I -+ B(H) such that 'P == "p1'(, where 1'( is the quotient map from A to AI I. Clearly, (H, 1/J) is a non-zero irreducible representation of AI I. Since AI I is postliminal, [{(H) C "p(AII); that is, 1{(H) C 'P(A). Suppose now that ker('P) does not contain I. Then the representation (H', 'P') == (H, 'P)I of I is non-zero. It is also irreducible, by Theorem 5.5.2. Since I is postliminal, this implies that l{ (H') c 'P' (I). It is easily checked that for b E I the operator 'P( b) is compact if and only if 'P' (b) is compact. Hence, there is an element b of I such that 'P( b) is non-zero and compact. Thus, whether ker( 'P) does or does not contain I, there is an element a of A such that 'P(a) is non-zero and compact. This shows that A is postliminal. 0 
170 5. Representations of C*-Algebras A consequence of this _result is that a C*-algebra A is postliminal if and only if its unitisation A is postliminal. 5.6.4. Ezample. Let A denote the Toeplitz algebra (Section 3.5). From Theorem 3.5.10, its commutator ideal is K(H 2 ), and this is liminal as we saw in Example 5.6.3. Moreover, the quotient A/K(H2) is *-isomorphic to G(T) (Theorem 3.5.11), so it is abelian and therefore liminal (cf Ex- ample 5.6.1). Hence, A is postliminal by Theorem 5.6.2. However, A is not liminal, as is seen by observing that the identity representation of A on H2 is irreducible (by Theorem 3.5.5) and not finite-dimensional. A C*-algebra is said to be elementary if it is *-isomorphic to K(H) for some Hilbert space H. It is easily checked that a simple postliminal C*-algebra is elementary. An elementary C*-algebra is unital if and only if it is finite-dimensional, so an infinite-dimensional unital simple C* -algebra is not postliminal. In particular, it follows from Theorem 4.1.16 that if H is a separable infinite-dimensional Hilbert space, then the correspond- ing Calkin algebra is a simple C*-algebra which is not postliminal. An application of Theorem 5.6.2 shows that B( H) therefore cannot be postlim- inal, either. 5.6.3. Theorem. Let (H, cp) and (H', cp') be non-zero irreducible repre- sentations of a postliminal C*-algebra A. Then they are unitarily equivalent if and only if their kernels are the same. Proof. We show only the backward implication because the forward im- plication is trivial. Suppose then that ker( <p) = ker( cp'). Observe that the map 1/;:cp(A)  cp'(A), cp(a)  cp'(a), is well-defined and a *-isomorphism. Since A is postliminal, I{(H) C <p(A). We show that 1/;(I«H)) C 1«H'): Let p be a rank-one projection in B(H). Then pB(H)p = Cp, so if q = 1/;(p), we have ql{(H')q C 1/;(Cp) = Cq, since K(H') C cp'(A). From this it is easily verified that q is a rank-one projection on H'. Since the rank-one projections in B( H) have closed linear span K(H), we have 1/;(K(H)) C K(H'), and the reverse inclusion holds by symmetry, so the restriction 1/;: 1{(H)  I{(H') is a *-isomorphism. It follows from Theorem 2.4.8 that there exists a unitary u: H  H' such that 1/; ( v) = u vu * for all v E I< ( H ). Let a E A and w E 1«H'). Then there exists v E K(H) such that w = uvu*, and there exists b E A such that v = cp(b). Since <p( a)v E K( H), we have 1/;( cp( a) )1/;( v) = 1/;( cp( a)v) = 1/;( cp( ab)) = ucp( ab )u* = (u<p(a)u*)(ucp(b)u*). Hence, 1/;(<p(a))w = (ucp(a)u*)w. Now I«H') is an essential ideal in B(H') (Example 3.1.2), so this argument shows that 1/;(cp(a)) = u<p(a)u*; that is, cp'(a) = ucp(a)u*. Thus, u implements a unit- ary equivalence between the representations (H, cp) and (H', cp'). 0 
5. Exercises 171 5.6.4. Theorem. If A is a non-zero postliminal C*-algebra, then the canonical map A  Prim(A) is a homeomorphism. Proof. This is immediate from Theorems 5.4.9 and 5.6.3. 0 5. Exercises 1. Let T be a pure state on a C*-algebra A, and y a unit vector in H T such that T( a) = (CPT( a )(y), y) for all a E A. Show that there is a scalar A of modulus one such that y = Ax T . 2. Let H be a Hilbert space and x a unit vector of H. Show that the functional w x : B(H)  C, u  (u(x), x), is a pure state of B(H). Show that not all pure states of B(H) are of this form if H is separable and infinite-dimensional. 3. Give an example to show that a quotient C*-algebra of a primitive C*-algebra need not be primitive. 4. If I is a primitive ideal of a C* -algebra A, show that M n (I) is a primitive ideal of Mn(A). (Thus, if A is primitive, so is Mn(A).) 5. Let A be a C*-algebra. Show the following conditions are equivalent: (a) A is prime. (b) If aAb = 0, then a or b = 0 (a, b E A). 6. Let S be a set of C*-subalgebras of a C*-algebra A that is upwards- directed, that is, if B, C E S, then there exists DES such that B, C C D. Show that (US)- is a C*-subalgebra of A. Suppose that all the algebras in S are prime and that A = (US)-. Show that A is prime. 7. If A is a C*-algebra, its centre C is the set of elements of A commuting with every element of A. Show that C is a C*-subalgebra of A. Show that if A is simple, then C = 0 if A is non-unital and C = C1 if A is unital. 8. Let S be an upwards-directed set of closed ideals in a C*-algebra A (cf. Exercise 5.6 for the term upwards-directed). Suppose that A = (US)-, and that all of the algebras in S are postliminal. Show that A is postliminal. 9. Let A be a C*-algebra. If I, J are postliminal ideals in A (that is, closed ideals that are postliminal C*-algebras), show that I + J is postliminal also. Deduce from this and Exercise 5.8 that there is a largest postliminal ideal I in A (which may, of course, be the zero ideal). Show that A/I has no non-zero postliminal ideals. 
172 5. Representations of C*-Algebras 5. Addenda If G is an ordered group, then the C* -algebra T( G) generated by all generalised Toeplitz operators with symbol in C( G) is primitive [Mur] (cf. Addenda 3). A representation (H,cp) of a C*-algebra A is said to be Type I if the von Neumann algebra c.p(A)" is Type I (cf. Addenda 4). We say A itself is Type I (as a C*-algebra) if all its representations are Type I. A C*-algebra is Type I if and only if it is postliminal. If A is a C*-algebra, a composition series for A is a family (lfJ )fJ5:a of closed ideals I fJ of A indexed by the ordinals (3 less than or equal to a fixed ordinal a, and such that (a) 10 = 0, Ia = A; (b) 11' is contained in 1fJ if 'Y < (3 < a; (c) if (3 is a limit ordinal, (3 < a, we have If3 = (U1'<f3I-y)- . The following conditions are equivalent: (a) A is a postliminal C*-algebra. (b) A has a composition series (lf3)f3$;a such that 113+1/113 is postliminal for all (3. (c) A has a composition series (IfJ)f3<a such that IfJ+l/IfJ is liminal for all (3. A C* -subalgebra of a postliminal C* -algebra is postliminal also. References: [Dix 2], [Ped]. 
CHAPTER 6 Direct Limits and Tensor Products This chapter is concerned with a number of techniques for constructing new C* -algebras from old. In Section 6.1 we introduce direct limits, and in Section 6.2 we use them to exhibit examples of AF-algebras, particularly examples of simple AF-algebras. The AF-algebras form a large class, which is relatively easy to analyse in that it is closely associated with the class of finite-dimensional C*-algebras, but which is nevertheless highly non-trivial. Some of these algebras play an important role in mathematical physics. The second fundamental construction of this chapter is the tensor prod- uct. Again, this is a device for getting new C*-algebras, but is also a powerful tool in the general theory. Finally, we introduce the nuclear C*-algebras, whose distinguishing feature is that they behave very well with respect to tensor products. A key result is a theorem of Takesaki, which asserts that abelian C*-algebras are nuclear. 6.1. Direct Limits of C*-Algebras Although our principal aim in this section is to construct direct limits of C*-algebras, we begin with direct limits of groups, because these will be needed in Chapter 7 in connection with K-theory. If (Gn)=l is a sequence of groups, and if for each n we have a homo- morphism Tn: G n  G n + 1 , then we call (G n , Tn)=l a direct sequence of groups. Given such a sequence and positive integers n < m, we set Tnn = id Gn and we define Tnm: G n  G m inductively on m by setting Tn,m+l = TmTnm. If n < m < k, we have Tnk = TmkTnm. If G' is a group and we have homomorphisms pn:  G' such that the diagram G n Tn G n + 1 ---+ pn ! pn+l G' 173 
174 6. Direct Limits and Tensor Products commutes for each n, that is, pn = pn+1Tn' then pn = pmTnm for all m > n. The product II  l Gk is a group with the pointwise-defined operation, and if we let G' be the set of all elements (Xk)k in IIl Gk such that there exists N for which Xk+1 = Tk(Xk) for all k > N, then G' is a subgroup of II  1 G k . Let ek be the unit of G k . The set F of all (Xk)k E IIl G k such that there exists N for which Xk = ek for all k > N is a normal subgroup of G', and we denote the quotient group G' / F by G. We call G the direct limit of the sequence (Gn,Tn) - l' and where no ambiguity can result we sometimes write limGn for G. --+ If x E G n , define fn(x) to be the sequence (Xk) where Xk = ek for k < n, and Xn+k = Tn,n+k(X) for all k > o. Then fn(x) E G', and the map Tn: G n -+ G, x'-"'+ f n ( X ) F, is a homomorphism, called the natural homomorphism from G n to G. It is straightforward to check that the diagram G n Tn G n + 1 ----+ Tn 1 Tn+1 G commutes for each n, and that G is the union of the increasing sequence ( T n ( G n ) ) - 1 . 6.1.1. Theorem. Let G be the direct limit of the direct sequence ofgroups (G n , Tn)  1, and let Tn: G n -+ G be the natural map for each n. (1) lfx E G n and y E G m and Tn(X) = Tm(y), then there exists k > n,m such that T nk( x) = T mk(Y). (2) If G' is a group, and for each n there is a homomorphism pn: G n -+ G' such that the diagram G n Tn G n + 1 ----+ pn 1 pn+1 G' commutes, then there is a unique homomorphism p: G -+ G' such that for each n the diagram G n Tn G  pn lp G' commutes. 
6.1. Direct Limits of C*-Algebras 175 Proof. Condition (1) follows directly from the definitions. Assume we have G' and pn as in Condition (2). If x E G n and y E G m and T n ( x) = Tm(y), then by Condition (1) there exists k > n, m such that Tnk(X) = Tmk(Y). Hence, pn(x) = pkTnk(X) = pkTmk(Y) = pm(y). Thus, we can well-define a map p: G  G' by setting p( T n ( x)) = pn( x). It is easily checked that p is a homomorphism, and by definition pT n = pn for all n. Uniqueness of p is clear. 0 6.1.1. Remark. From Condition (1) of the preceding theorem, if Tn(X) = e, then there exists k > n such that Tnk(X) = e, a result we shall be using frequently. (The symbol e denotes the unit of the relevant group.) Let A be a *-algebra. A C*-3eminorm on A is a semi norm p on A such that for all a, b E A we have p(ab) < p(a)p(b), p(a*) = p(a) and p( a* a) = p( a)2. If in addition p is in fact a norm, we call p a C*-norm. If cp: A  B is a *-homomorphism from a *-algebra into a C*-algebra, then the function p: A  R+, a  IIcp(a)lI, is a C* -seminorm on A, and if cp is injective, p is a C* -norm. If p is any C*-seminorm on a *-algebra A, the set N = p-l {OJ is a self-adjoint ideal of A, and we get a C*-norm on the quotient *-algebra A/N by setting lIa+NII = pea). If B denotes the Banach space completion of A/ N with this norm, it is easily checked that the multiplication and involution operations extend uniquely to operations of the same type on B so as to make B a C*-algebra. We call B the enveloping C*-algebra of the pair (A, p), and the map i:AB, aa+N, the canonical map from A to B. Of course, i(A) is a dense *-subalgebra of B. If p is a C*-nonn, we refer to B more simply as the C*-completion of A. In this case A is a dense *-subalgebra of B. Let (An)=l be a sequence of C*-algebras and suppose that for each n we have a *-homomorphism CPn:An  An+1. Then we call (An,<Pn)  1 a direct 3equence of C*-algebra3. The product rrl Ak is a *-algebra with the pointwise-defined operations, and if A' denotes the set of all elements a = (ak)k of rrl Ak such that there is an integer N for which ak+l = <Pk(ak) for all k > N, then A' is a *-subalgebra of rrl Ak. Note that lIak+lll < Ilak II if k > N (since CPk is norm-decreasing), so the sequence (1lakll)k is eventually decreasing (and of course bounded below). It therefore converges, and we set p(a) = limk--+oo Ilakll. It is straightforward to verify that p:A'  R+, a  pea), 
176 6. Direct Limits and Tensor Products is a C*-seminorm on A'. We denote the enveloping C*-algebra of (A' ,p) by A, and call it the direct limit of the sequence (An, CPn)=l. If no ambiguity results, we sometimes write limAn for A. ---+ Similar to the group case, if a E An, we define <pn(a) in A'to be the sequence (ak)k such that a},.. . , an-l are zero and an+k = 4'n,n+k( a) for all k > o. If i: A' -+ A is the canonical map, then the map cpn: An  A, a  i( <p n ( a)), is a *-homomorphism, called the natural map from An to A. A routine argument shows that for all n the diagram An CPn An+l  cpn ! cpn+l A commutes, and that the union of the increasing sequence of C*-subalgebras (cpn(An))n is a dense *-subalgebra of A. Also, lI<.pn( a) II = lim II4'n,n+k( a) II k-.oo (1) if a E An. 6.1.2. Theorem. Let A be the direct limit of the direct sequence of C*-algebras (An, CPn)=l' and suppose that cpn: An  A is the natural map for each n. (1) If a E An' bEAm, £ > 0 and cpn(a) = cpm(b), then there exists k > n,m such that IICPnk(a) - CPmk(b)11 < c. (2) If B is a C*-algebra and there is a *-homomorphism 1jJn: An  B for each n such that the diagram An cpn  An+l 1jJn 11jJn+l B commutes, then there is a unique *-homomorphism 'ljJ: A  B such that for each n the diagram An cpn  A 'ljJ n  ! 'ljJ B commutes. 
6.1. Direct Limits of C*-Algebras 177 Proof. Condition (1) follows from Eq. (1) above. Suppose that Band 1/Jn are as in Condition (2). Let a E An and bEAm, and suppose that cpn(a) = cpm(b). If £ > 0, then by Condition (1) there exists k > n, m such that IICPnk(a) - CPmk(b)11 < c. Consequently, lIn(a) - m(b)1I = IIk(CPnk(a) - CPmk(b))11 < IICPnk(a) - CPmk(b)11 < c. Let- ting £  0, we therefore have 1/Jn(a) = m(b). This shows that we can well- define a map 1/J from C = U=lcpn(An) to B by setting 1/J(cpn(a)) = n(a). If k is any integer, then lIn(a)11 = l11/Jn+kcpn,n+k(a)11 < IICPn,n+k(a)lI, and therefore 111/J(cpn(a)) II = l11/J n (a)1I < limk--+oo IICPn,n+k(a)1I = IIcpn(a)lI. Thus, 1/J is norm-decreasing, and it is easily seen to be a *-homomorphism. Since C is a dense *-subalgebra of A, we can extend t/J to a *-homomorphism : A  B, and cpn = n for all n. Uniqueness of  follows from density of C in A. 0 6.1.2. Remark. Retaining the notation of the preceding theorem, if a E An and cpn(a) = 0 and c > 0, then by Condition (1) there exists k > n such that IICPnk( a) II < c. 6.1.3. Remark. Let A be a C*-algebra and let (An) - l be an increas- ing sequence of C*-subalgebras of A whose union is dense in A. Let CPn: An  An+1 be the inclusion map. A straightforward application of Theorem 6.1.2, Condition (2), shows that A is (*-isomorphic to) the direct limit of the direct sequence (An, CPn)=l. 6.1.3. Theorem. Let S be a non-empty set of simple C*-subalgebras of a C*-algebra A. Suppose that S is upwards-directed (that is, if B, C E S, then there exists DES such that D contains B and C), and uS is dense in A. Then A is simple also. Proof. To show that A is simple it suffices to show that if 7r: A  B is a surjective *-homomorphism onto a non-zero C*-algebra B, then it is injective. If C E S, then the restriction of 7r to C is either zero, or it is injective, and therefore isometric. Since 7r is not the zero map on US, it follows easily from the upwards-directed property of S that 7r is not the zero map on any non-zero C E S. Hence, 7r is isometric on US, and therefore, by continuity, 7r is isometric on A. 0 6.1.4. Theorem. Suppose that (An, CPn)=l is a direct sequence of simple C*-algebras. Then the direct limit limAn is simple, also. --+ Proof. Let cpn: An  A be the natural map, where A = limAn. Then --+ the set S = {cpn(A n ) I n > I} is an upwards-directed family of simple C*-subalgebras of A whose union is dense in A, so by Theorem 6.1.3 A is simple. 0 
178 6. Direct Limits and Tensor Products 6.2. Uniformly Hyperfinite Algebras The C* -algebras of the title form an interesting class, since they are highly non-trivial yet accessible to detailed analysis. Before introducing them, however, we shall need to consider some preliminary material. We begin by characterising the finite-dimensional simple C*-algebras, since uniformly hyperfinite algebras are defined in terms of these algebras. 6.2.1. Remark. A non-zero finite-dimensional C*-algebra is simple if and only if it is of the form Mn(C) for some n. To see this, suppose that A is a non-zero simple finite-dimensional C*-algebra. By Example 5.6.2 A is liminal. Hence, if (H, cp) is any non-zero irreducible representation of A, then c.p(A) = I{(H) and H is therefore finite-dimensional. Moreover, since ker( c.p) is a proper closed ideal of A, it is the zero ideal. Therefore, if n = dim(H), then A is *-isomorphic to K(H) = B(H), which in turn is *-isomorphic to Mn(C). We shall need the following lemmas in this section and also at various points in the sequel. 6.2.1. Lemma. Let p, q be projections in a unital C*-algebra A and suppose that IIq - pll < 1. Then there exists a unitary u in A such that q = upu. and 1I1-ull < J2llq-pll, namely, u = vlvl- 1 , where v = 1-p-q+2qp. Proof. If v = 1 - p - q + 2qp, then v.v = 1 - (q - p)2 by computation. Because v. = 1 - p - q + 2pq, we have also vv. = 1 - (p - q)2, so vv. = v*v; that is, v is normal. Now IIq - pll < 1 implies that II( q - p )211 = IIq - pll2 < 1, so 1 - (q - p)2 = v*v is invertible. Hence, v is invertible by normality. Consequently, u = vlvl- 1 is a unitary. Now vp = qp = qv, so pv* = v*q, and pv*v = v.qv = v*vp. It follows that p commutes with Ivl and therefore with lvi-I. Hence, up = qu; that is, q = upu.. Since Re(v) = 1 - (q - p)2 = Iv1 2 , we have Re(u) = Re(v)lvl- 1 = Ivl. Therefore, 111 - ull 2 = 11(1 - u*)(l - u)1I = 2111 - Re(u)11 = 2111 - Ivlll < 2111-lv1 2 11 (because 1-t < 1-t 2 for all t E [0,1]). Since I-Iv 1 2 = (q- p)2, therefore 111- ull 2 < 211q - p1l2, so 111- ull < J2l1q - pll. 0 If n is a locally cmpact Hausdorff space and I E Co(), extend I to a continuous function I on the one-point compactification n of n by setting j( 00) = 0, where 00 is the "point at infinity." If 6 > I/(w)1 for all wEn, then 6 > 11/1100, since 11/1100 = lIilloo, and 6 > lIilloo because the continuous function w .-...+ Ij(w)1 attains its upper bound on the compact space n. 6.2.2. Lemma. Let a be a self-adjoint element of a C*-algebra A such that Iia - a 2 11 < 1/4. Then there is a projection pEA such that Iia - pll < 1/2. 
6.2. Uniformly Hyperfinite Algebras 179 Proof. We may suppose that A is abelian and may therefore suppose that A = C o (f2) for some locally compact Hausdorff space f2. The hy- pothesis implies that 1/2 is not in the range of lal, and therefore the set S = lal- 1 (1/2,00) is open and compact in n (it is compact since it is equal to {w E n I la(w)1 > 1/2}). Hence, p = Xs is a projection in A. Since la(w) - xs(w)1 < 1/2 for all wEn, therefore by the observation preceding this lemma, lIa - piloc> < 1/2. 0 A positive linear functional on a C*-algebra A is tracial if r(a*a) = r( aa*) for all a E A. Equivalently, r( ab) = r( ba) for all a, b E A. To see the equivalence, let r be tracial and let b, e be self-adjoint elements of A. Then if a = b+ie, we have a*a = b 2 +c 2 +i(be-cb) and aa* = b 2 +e 2 +i(eb- be). Since r(a*a) = r(aa*), we get r(be - eb) = r(eb - be) = -r(be - eb), so r(bc- cb) = o. Thus, r(be) = r(eb) for all b,e E Asa and, therefore, for all b, e E A. A tracial positive linear functional T is faithful if r( a* a) = 0 implies that a = O. 6.2.1. Ezample. The function n tr: Mn(C)  C, (Aij)   L Aii, i=l is a faithful tracial state on Mn(C). In fact, this is the only tracial state on Mn(C). To show this, we need only show that all tracial states take the same value on the rank-one projections, since these span Mn(C). But this will follow easily if we show that all rank-one projections are unitarily equivalent. Supposing then that p and q are rank-one projections, there exist unit vectors e and f such that p = e Q9 e and q = f Q9 f. Since there exists a unitary u E Mn(C) such that u(e) = f, we have q = u(e) Q9u(e) = u(e Q9 e)u* = upu*. 6.2.2. Remark. If H is an infinite-dimensional Hilbert space, then K(H) does not admit a tracial state. Suppose the contrary, and let r be a tracial state on K(H). Observe that all the rank-one projections on H are unitarily equivalent (same proof as in Example 6.2.1), and therefore r takes on the same (positive) value, t say, on all rank-one projections. Now let E be an orthonormal basis for H. If el,..., en E E and p is the projection l:1 ei Q9 ei, then r(p) < 1 and r(p) = nt, so n < l/t. Thus, l/t is an upper bound for the integers, an absurdity. This shows that K(H) has no tracial state, as claimed. 6.2.3. Remark. If r is a tracial positive linear functional of a C*-algebra A, then N r is an ideal of A. For, we know that N r is a left ideal, because 
180 6. Direct Limits and Tensor Products r is a positive functional, and the tracial condition implies that N r = N:, from which N r is an ideal as claimed. If A is simple, then any non-zero tracial positive linear functional r on A is faithful, since in this case the proper closed ideal N r of A is the zero ideal. 6.2.4. Remark. Let (An)=l be an increasing sequence of C*-subalgebras of a C*-algebra A such that A = (U=l A n )-. Suppose that A is unital and that all the algebras An contain the unit of A. Then if each algebra An admits a unique tracial state, r n say, A also admits a unique tracial state. We show this: The restriction of r n+ 1 to An is a tracial state (r n+ 1 (1) = 1, so the restriction has norm one). Therefore, by the uniqueness assumption, r n+l is equal to r n on An. Define r on the *-subalgebra U=l An of A by setting r( a) == r n (a) if a E An. It is easily checked that this gives a norm-decreasing linear function from UnAn to C, so we can extend to get a bounded linear functional r: A  C. It is clear that T is a tracial state on A; that it is the unique such state follows from the uniqueness assumption on the algebras An. A uniformly hyperjinite algebra or UHF algebra is a unital C*-algebra A which has an increasing s'equence (An)l of finite-dimensional simple C*-subalgebras each containing the unit of A such that U=l An is dense in A. Since An is simple and finite-dimensional, it is *-isomorphic to some Mk( C), so it admits a unique tracial state. It follows from Remark 6.2.4 that A has a unique tracial state, r say. By Theorem 6.1.3 A is simple. Let n, d be positive integers. We call the unital *-homomorphism tp:Mn(C) -4 Mdn(C), a  (: '. :), the canonical map from Mn(C) to Mdn(C) (cp(a) has d blocks of a down the main diagonal, and everywhere else it is zero). Denote by S the set of all functions s: N \ {OJ --+ N \ {OJ. If s E S, define s! E S by set ting s!(n) == s(1)s(2)... s(n) (n > 1). Let cpn: Ms!(n)( C)  M s !(n+l)( C) be the canonical map (obviously, s!( n) divides s!( n + 1)). We denote by Ms the direct limit of the direct sequence (Ms!(n)(C), CPn)=l. Since the C*-algebras Ms!(n)(C) are finite-dimensional simple C*-algebras, it is clear that Ms is a UHF algebra. 
6.2. Uniformly Hyperfinite Algebras 181 Define a function Cs from the set of prime numbers to N U {+oo} by setting, for each prime r, cs(r) = sup{m E N I r m divides some s!(n)}. 6.2.3. Theorem. Let s, s' E S and suppose that Ms, Ms' are *-isomoIphic. Then Cs = £s,. Proof. Let 7r: Ms  Ms' be a *-isomorphism, let T and T' be the unique tracial states of Ms and Ms" respectively, and let cpn: Ms!(n)(C)  Ms and 'ljJn: MS'!(n)(C)  Ms' be the natural *-homomorphisms. Clearly, T'7r is a tracial state on Ms, so by uniqueness of the tracial state, T = r' 7r . To prove the theorem, it suffices to show that Cs < Cs', since the reverse inequality will then follow by symmetry. Therefore, it is enough to show that for each positive integer n there is a positive integer m such that s!( n) divides s'!(m). (For then if r is a prime and k a positive integer such that r k divides s!( n), then r k divides s'!( m), and this shows that €s( r) < Cs' (r).) Suppose then that n is a positive integer. Let p be a rank-one projec- tion in Ms!(n)(C). Since T'{)n is the unique tracial state on Ms!(n)(C), we have T( cpn(p)) = 1/ s!( n). Now 7r( cpn(p)) is a projection in Ms" so there is a positive integer m and a self-adjoint element a E MS'!(m)(C) such that 117r ( '() n (p )) - 'ljJ m ( a ) II < 1/8 and 117r( cpn(p)) - m( a 2 )11 < 1/8 (this uses the density of Ul k(Ms'!(k)(C)) in Ms'). Hence, Iia - a 2 11 = lIm( a) _ m( a 2 )11 < II  m ( a) - 7r ( cp n (p ) ) II + 117r ( cp n (p )) - 'ljJ m ( a 2 ) II < 1/4. It follows from Lemma 6.2.2 that there is a projection q in MS'!(m)(C) such that Iia - qll < 1/2. Hence, 117r( cpn(p)) - m( q)1I < 117r( cpn(p)) - m( a)1I + II1/J m ( a) _ m( q )11 < 1/8 + 1/2 < 1. Applying Lemma 6.2.1, the projections 7r('{)n(p)) and m(q) are unitarily equivalent in Ms" and therefore T' (m( q)) = T' (7r( cpn(p))) = T( cpn(p)) = l/s!(n). But T''ljJm is the unique tracial state on MS'!(m)(C), so T''ljJm(q) must be of the form d/s'!(m) for some positive integer d. Therefore, s'!(m) = ds!(n), so s!(n) divides s'!(m). 0 6.2.4. Corollary. There exists an uncountable number of UHF algebras that are not *-isomorphic. 
182 6. Direct Limits and Tensor Products Proof. Let (Tn) be the sequence of prime numbers. If s E S, let s E S be defined by s( n) = Tn. Then cs( Tn) = Sn. It follows that if s, s' E S and C8 = Cs' then s = s'. Since the set S is uncountable, therefore by Theorem 6.2.3, (M 8 )sES is an uncountable family of UHF algebras that are not *-isomorphic. 0 Now we present an application of these algebras to the theory of von Neumann algebras. A factor on the Hilbert space H is a von Neumann algebra A on H such that A n A' = C id H (cf Addenda 4). If H is a Hilbert space, then B(H) is a factor. It is harder to give other examples (although examples exist in abundance). We shall presently exhibit an example of an infinite- dimensional factor not *-isomorphic to any B(H). A von Neumann algebra on a Hilbert space H is hyperfinite if it has a weakly dense C*-subalgebra that is a UHF algebra and whose unit is idH. 6.2.2. Ezample. If H is a separable Hilbert space, then B(H) is hyper- finite. This is clear if H is finite-dimensional. To prove it for the infinite- dimensional case, let A be an infinite-dimensional UHF algebra, and let (H, cp) be a non-zero irreducible representation of A. Since A is simple, cp is a *-isomorphism of A onto cp(A), so cp(A) is a UHF algebra. Now if x is any non-zero vector of H, then it is cyclic for (H,cp) (by irreducibility), so H = [cp(A)x], and since A is separable, this shows that H is separable. Clearly, H is infinite-dimensional since A is. Because (H, cp) is irreducible, cp(A)' = C, so cp(A)" = B(H). Therefore, cp(A) is a UHF subalgebra of B(H) containing the identity map id H and is weakly dense, so B(H) is a hyperfinite algebra. 6.2.5. Theorem. Let A be a UHF algebra, and let r be its unique tracial state. Then the von Neumann algebra CPr(A)" is a hyperfinite factor ad- mitting a faithful tracial state. Proof. Let B = CPr(A). Since A is unital and the representation (Hr,CPr) is non-degenerate (it is cyclic), we have cp(l) = idHr. The representation (H r, CPr) is faithful because A is simple, so B is *-isomorphic to A and is therefore a UHF algebra. Hence, the von Neumann algebra B" is hyper- finite. The tracial condition gives the equation (UU'(Xr), x r ) = (U'U(Xr), x r ) for all u, u' E B, and weak density of B in B" implies that this equation holds for all u, U I E B" also. Hence, the function w: B" ---+ C, U  (u(xr), x r ), 
6.2. Uniformly Hyperfinite Algebras 183 is a tracial state on B". To show that w is faithful, we show that the vector X T is separat- ing for B". Suppose that U E B" and u(x T ) = o. If v E B, then Iluv(x T )1I 2 = w(v*u*uv) = w(vv*u*u) by the tracial condition, so Iluv(x T )11 2 = (vv*u*u(XT),X T ) = o. Hence, u[Bx T ] = 0, and since X T is cyclic for B, this shows that U = o. Thus, x T is a separating vector for B", as claimed. Now let P be a projection in B' n B". The function w': B"  C, U t--+ w(pu), is a weakly continuous tracial positive linear functional on B". Hence, when we restrict it to the UHF algebra B, it is a constant t times the unique tracial state w B of B; that is, w' (v) = tw( v) for all v E B. Therefore, by weak density of B in B" and weak continuity of w' and w, we have w' = two Hence, t = w(p), so 0 = w'(l - p) = w(p)w(l - p). Since p and 1 - pare positive and w is faithful, this implies that either p or 1 - p is zero. We have therefore shown that the only projections in the von Neumann algebra B' n B" are the trivial ones. Since a von Neumann algebra is the closed linear span of its projections, this shows that B' n B" = C, and therefore B" is a factor. 0 If we suppose in the preceding theorem that A is an infinite-dimensional UHF algebra, then the yon Neumann algebra cpT(A)" is a factor that is not *-isomorphic to B(H) for any Hilbert space H. (By Remark 6.2.2 B(H) has no faithful tracial state if H is infinite-dimensional.) A more general class than the UHF algebras, but which is similarly defined in terms of finite-dimensional algebras, is the class of AF -algebras. An AF-algebra is a C*-algebra that contains an increasing sequence (An)=l of finite-dimensional C*-subalgebras such that U=l An is dense in A. 6.2.3. Ezample. If H is a separable Hilbert space, then I{(H) is an AF-algebra. To show this, we may suppose that H is infinite-dimensional. Let (en)=l be an orthonormal basis for H, and let Pn be the projection 2:::7=1 ei 0 ei. The sequence (Pn) is an approximate unit for ]{(H) (cf. Ex- ample 3.1.1), so I{(H) = (U=lPnI{(H)Pn)-. If U E I{(H), then PnUPn = 2:::j=l(ei 0 ei)u(ej 0 ej) = 2:::j=l(u(ej),ei)ei 0 ej, so the C*-algebra Pn]{(H)pn is finite-dimensional. This shows that I{(H) is an AF-algebra. 6.2.4. Ezample. If A is a direct limit of a direct sequence (An, CPn)=l of C*-algebras, where the An are finite-dimensional, then A is an AF-algebra. For in this case the sequence of algebras (cpn(An))n is increasing, its union is dense in A, and 'P n ( An) is finite-dimensional for each n. 
184 6. Direct Limits and Tensor Products The "converse" is also true. More precisely, if A is an AF -algebra, then it is *-isomorphic to a direct limit of finite-dimensional C*-algebras (by Remark 6.1.3). A finite-dimensional C* -algebra is the linear span of its projections, as we observed in Section 2.4, so an AF -algebra is the closed linear span of its projections. A consequence is that an abelian AF-algebra has totally disconnected spectrum. 6.2.6. Theorem. If I is a closed ideal in an AF-algebra A, then I and AI I are A F-alge bras. Proof. Suppose (An)=l is an increasing sequence of finite-dimensional C*-subalgebras of A whose union is dense in A, and let 7r: A  AI I be the quotient *-homomorphism. Then (7r(An))n is an increasing sequence of finite-dimensional C*-subalgebras of AI I and the union of these algebras is dense in AI I, so AI I is an AF-algebra. Set In = InA n , and let J = (UnIn)-. Then J is a closed ideal of A con- tained in I, and since (In)n is an increasing sequence of finite-dimensional C*-subalgebras of I, we shall have shown that I is an AF-algebra if we show that I = J. To prove this consider the well-defined *-homomorphism c.p: AI J  AI I, a + J t-+ a + I. We shall prove that I = J by showing that c.p is isometric, and to see this it suffices to show that c.p is isometric on the C*-subalgebras (An + J)I J, since these form an increasing sequence whose union is a dense *-subalgebra of AIJ. Denote by 'ljJ:(An + J)IJ  Anl(An n J) and (}:(An + I)II  Anl(An n I) the canonical *-isomorphisms (cf. Remark 3.1.3), and by i: (An + I)II  All the inclusion. Since An n I = An n J, and since the restriction of c.p to (An + J) I J is the composition i{}-l1/J of isometric maps, c.p is isometric on (An + J)I J. 0 It is evident from Theorem 6.2.6 that a C*-algebra A is an AF-algebra if and only if A is an AF -algebra. We shall have more to say about AF -algebras in later sections. 6.3. Tensor Products of C*-Algebras If Hand I{ are vector spaces, we denote by H 0 K their algebraic tensor product. This is linearly spanned by the elements x 0 Y (x E H, y E [{). (There is a conflict between the tensor notation and our use of x 0 y to denote a rank-one operator on a Hilbert space, but the context will always resolve the ambiguity.) One reason why tensor products are useful is that they turn bilinear maps into linear maps. More precisely, if c.p: H x I{  L is a bilinear map, 
6.3. Tensor Products of C*-Algebras 185 where H, K and L are vector spaces, then there is a unique linear map c.p/: H fl:) K -+ L such that c.p'(X fl:) y) = c.p(x, y) for all x E Hand y E K. If T, P are linear functionals on the vector spaces H, K, respectively, then there is a unique linear functional T fl:) P on H fl:) K such that (T fl:) p)( x fl:) y) = T( X )p(y) (x E H, y E K), since the function H x l{ -+ C, (x, y)  T(X)p(y), is bilinear. Suppose that Ej=l x j fl:)Yj = 0, where x j E Hand Yj E K. If YI,. . . , Yn are linearly independent, then Xl = ... = X n = o. For, in this case, there exist linear functionals Pj: K -+ C such that Pj(Yi) = 6ij (Kronecker delta). If T:H -+ C is linear, we have 0 = (T 0 pj)(E=1 Xi 0 Yi) = E7:1 T(Xi)Pj(Yi) = E7:1 T(Xi)6 ij = T(Xj). Thus, T(Xj) = 0 for arbitrary T, and this shows that Xl = . . . = X n = O. Similarly, if Ej=l x j 0 Yj = 0 and Xl,. . . , X n are linearly independent, then YI = . . . = Yn = o. If u: H -+ H' and v: I{ -+ l{' are linear maps between vector spaces, then by elementary linear algebra there exists a unique linear map u 0 v: H 01{ -+ H' 0 K ' such that (u 0 v)(x fl:) y) = u(x) 0 v(y) for all x E H and all Y E K. The map (u, v)  u fl:) v is bilinear. If Hand K are normed, then there are in general many possible norms on H fl:) K which are related in a suitable manner to the norms on Hand K, and indeed it is this very lack of uniqueness that creates the difficulties of the theory, as we shall see in the case that Hand K are C*-algebras. When the spaces are Hilbert spaces, however, matters are simple. 6.3.1. Theorem. Let Hand I{ be Hilbert spaces. Then there is a unique inner product (. , .) on H 0 K such that (x 0 Y, x' 0 y') = (x, x') (y, y') (X, x' E H, y,y' E K). Proof. If T and P are conjugate-linear maps from Hand K, respectively, to C, then there is a unique conjugate-linear map T 0 P from H 0 K to C such that (T fl:) p)(x fl:) y) = T(X)p(y) for x E Hand y E K. (Observe that f and p are linear, and set T 0 P = (f 0 p)-.) If x is an element of a Hilbert space, let T x be the conjugate-linear functional defined by set ting T x (y) = (x, y) . 
186 6. Direct Limits and Tensor Products Let X be the vector space of all conjugate-linear functionals on H 0K. The map H x I<  X, (x, y) .-...+ Tx 0 Ty, is bilinear, so there is a unique linear map M: H 0 K  X such that M(x 0 y) = Tx 0 Ty for all x and y. The map (., .): (H 0 K)2  C, (z, z') .-...+ M(z)(z'), is a sesquilinear form on H 0 K such that (x 0 y, x' 0 y') = (x, x')(y, y') (x, x' E H, y, y' E 1<). That it is the unique such sesquilinear form is clear. If z E H 0 1<, then z = L: j = 1 X j 0 Y j for some Xl,. . . , X n E Hand Yl , . . . , Yn E K. Let e},..., em be an orthonormal basis for the linear span OfYl,...,Yn. Then z = L:j:1xj 0ej for some x,...,x E H, and therefore, m (z,z) = L (x 0 ei,xj 0 ej) i,j=l m = L (x, xj)(ei, ej) i,j=l m = Lllxjl12. j=l Thus, (.,.) is positive, and if (z, z) = 0, then xj = 0 for j = 1,... , m, so z = o. Therefore, (. , .) is an inner product. 0 If H and K are as in Theorem 6.3.1, we shall always regard H 0 K as a pre-Hilbert space with the above inner product. The Hilbert space completion of H 0 K is denoted by H 0 I{, and called the Hilbert space tensor product of Hand I<. Note that IIx 0yII = IlxlIIlYII. It is an elementary exercise to show that if El and E 2 are orthonormal bases for Hand K, respectively, then El 0 E 2 = {x 0 Y I x E E 1 , Y E E 2 } is an orthonormal basis for H 0 K. If H', K' are closed vector subspaces of H, K, respectively, then the inclusion map H' 0 I{'  H 0 I{ is isometric when H' 0 K' has its canonical inner product. It follows that we may regard H' 0 I{' as a closed vector subspace of H 0 I{. 
6.3. Tensor Products of C*-Algebras 187 6.3.2. Lemma. Let H, K be Hilbert spaces and suppose that u E B(H) and v E B(I{). Then there is a unique operator u  v E B(H  K) such that (u  v)(x Q9 y) = u(x) Q9 v(y) Moreover, Ilu 0 vii = Ilullllvll. Proof. The map (u, v) .-...+ U Q9 v is bilinear, so to show that u Q9 v: H Q9 K  H Q9 K is bounded, we may assume that u and v are unitaries, since the unitaries span the C*-algebras B( H) and B( K). If z E H Q9 I{, then we may write z = 2:7=1 X j Q9 Yj, where Y1, . . . , Yn are orthogonal. Hence, (x E H, Y E I{). n II(u Q9 v)(z)11 2 = II L u(Xj) Q9 v(Yj)1I2 j=1 n = L lIu(x j) Q9 v(Yj )11 2 j=1 (since V(Yl), . . . , V(Yn) are orthogonal) n = L Ilu(xj)112I1v(Yj)1I2 j=1 n = L IIx j 11211Yj 11 2 j=l = Il z 11 2 . Consequently, Ilu Q9 vii = 1. Thus, for all operators u, v on H, I{, respectively, the linear map u Q9 v is bounded on H Q9 1< and hence has an extension to a bounded linear map u 0 v on H 0 K. It is easily verified that the maps B(H)  B(H 0I{), U'-"'+ u 0 idK, and B(I<)  B(H 0 I{), V'-"'+ idH 0 v, are injective *-homomorphisms and therefore isometric. Hence, Ilu 0 id II = lIull and II id vll = IIvll, so Ilu  vii = II( u  id)(id v )11 < lIu IIlIvll. If c: is a sufficiently small positive number, and if u, v =I 0, then there are unit vectors x and Y such that Ilu(x)11 > lIull- £ > 0 and Ilv(Y)11 > IIvll- £ > o. Hence, II(u 0 v)(x Q9 y)1I = Ilu(x)llllv(Y)11 > (Ilull - c:)(llvll - c:), so Ilu 0 vII > (Ilull- £)(llvll- c:). Letting c:  0, we get Ilu 0 vii > Ilullllvll.o 
188 6. Direct Limits and Tensor Products 6.3.1. Remark. Let Hand K be Hilbert spaces and suppose that u, u' E B(H) and v, v' E B(K). It is routine to show that (u  v)(u'  v') = UU'  VV' and ( " ) * *" * uQ)v =u Q)v. If Ul, . . . , Un are operators on H and VI,. . . , V n are linearly independent operators on K such that E j- l Uj  Vj = 0, then UI = ... = Un = o. For if x E H, choose orthonormal vectors el,. . . , em in H such that CUI(X) + ... + Cun(x) C Cel +... + Cem. Then there are scalars Aij such that Uj(x) = E:I Aijei for 1 < j < n. If Y E K, we have n m m n o = I:(I: Aijei) Q) Vj(Y) = I: ei 0 I: AijVj(Y), j=l i=l i=l j=l so E j- l AijVj(Y) = O. This shows that Ej=l AijVj = 0, and therefore, by linear independence of VI, . . . , V n , we get Aij = 0 for all i and j. It follows that UI (x) = . . . = u n ( x) = o. If A and B are algebras, there is a unique multiplication on A 0 B such that (a 0 b)( a' 0 b') = aa' 0 bb' for all a, a' E A and b, b' E B. We show this: Let La denote left multipli- cation by a, and let X be the vector space of all linear maps on A 0 B. If a E A and b E B, then La Q) Lb EX, and the map A x B  X, (a, b)  La 0 Lb, is bilinear. Hence, there is a unique linear map M: A 0 B  X such that M(a 0 b) = La 0 Lb for all a and b. The bilinear map (A 0 B)2  A 0 B, (c, d)  cd = M(c)(d), is readily seen to be the required unique multiplication on A 0 B. We call A 0 B endowed with this multiplication the algebra tensor product of the algebras A and B. If A and Bare *-algebras, then there is a unique involution on A 0 B such that (a 0 b) * = a * Q) b* for all a and b. The existence of such an involution is easily seen if we show that n n I: aj 0 b j = 0 =? I: aj Q) bi = O. j=l j=l 
6.3. Tensor Products of C*-Algebras 189 Choose linearly independent elements C1, . . . , C m in B having the same lin- ear span as b 1 ,...,b n . Then b j = Ll AijCi for unique scalars Aij. Since L}=l aj 0bj = 0, we have Li,j Aijaj Ci = 0, and therefore L;=l Aijaj = 0 for i = 1,..., m, because C1,..., C m are linearly independent. Hence, n - * Lj=l Aijaj = 0, and therefore, n n m L aj 0 bj = L L aj 0 Xijci j=l j=l i=l m n = L(L Xijaj) 0 ci i=l j=l m = L 0 0 ci i=l = o. We call A 0 B with the above involution the *-algebra tensor product of A and B. If A, Bare *-subalgebras of *-algebras A', B', respectively, we may clearly regard A 0 B as a *-subalgebra of A' 0 B'. 6.3.2. Remark. Let c.p: A  C and 'ljJ: B  C be *-homomorphisms from *-algebras A and B into a *-algebra C such that every element of c.p(A) com- mutes with every element of 'ljJ( B). Then there is a unique *-homomorphism 7r: A 0 B  C such that 7r( a 0 b) = c.p( a )'ljJ( b) (a E A, b E B). This follows from the observation that the map A x B -+ C, (a, b)  c.p(a)'ljJ(b), is bilinear and so induces a linear map 7r: A 0 B  C, which is easily seen to have the required properties. If A, A', B, B' are *-algebras and c.p: A  A' and 'ljJ: B  B' are *-homomorphisms, then c.p 0 'ljJ: A 0 B  A' 0 B' is a *-homomorphism. The proof is a routine exercise. We shall use the next result to show that there is at least one C*-norm on A 0 B if A and Bare C*-algebras. 6.3.3. Theorem. Suppose that (H, c.p) and (K, 'ljJ) are representations of the C*-algebras A and B, respectively. Then there exists a unique *-homomorphism 7r: A  B  B(H  f<) such that 7r(a 0 b) = c.p(a) @ 'ljJ(b) (a E A, b E B). Moreover, if c.p and 'ljJ are injective, so is 7r. 
190 6. Direct Limits and Tensor Products Proof. The maps <p': A -+ B( H &J K), a...... <p( a) &J id K , and 1/;': B -+ B( H &J I{), b...... id H 01/;( b), are *-homomorphisms, and the elements of <p'(A) commute with the ele- ments of'ljJ'(B). Hence, by Remark 6.3.2, there is a unique *-homomorphism 11": A 0 B -+ B(H &J K) such that 1r(a 0 b) = <p'(a)'ljJ'(b) = <p(a) 0'ljJ(b) (a E A, b E B). N ow suppose that the representations (H, c.p) and (I{, 1/;) are faithful, and let c E ker( 1r). We can write c in the form c = Ej=I aj 0 b j , where the elements b I , . . . , b n are linearly independent. Then 1/;( b I ), . . . , 1/;( b n ) are linearly independent, because 'ljJ is injective, and Ej=I c.p(aj) 0'ljJ(b j ) = O. Hence, cp( aI) = ... = c.p( an) = 0 (Remark 6.3.1). Since 'P is injective, al = . . . = an = 0, so c = o. Thus, ker( 7r) = o. 0 We denote 1r in Theorem 6.3.3 by c.p 01/;. Let A and B be C*-algebras with universal representations (H, 'P) and (K, 'ljJ), respectively. By Theorem 6.3.3 there is a unique injective *-homomorphism 1r: A 0 B -+ B(H 0 [<) such that 1r(a 0 b) = c.p(a) 0'ljJ(b) for all a and b. Hence, the function 11.11.: A 0 B -+ R + , c...... 111r( c) II, is a C*-norm on A  B, called the 3patial C*-norm. Note that IIa  bll. = lIallllbli. We call the C*-completion of A0B with respect to 11.11. the 3patial ten30r product of A and B, and denote it by A 0. B. In general, there may be more than one C*-norm on A 0 B. If , is a C*-nonn on A 0 B, we denote the C*-completion of A 0 B with respect to , by A 01' B. 6.3.4. Lemma. Let A, B be C*-algebras and let, be a C*-norm on A0B. Then for a' E A and b' E B the maps c.p:A-+A01'B, a-+a0b', and 1/;: B -+ A 0...,. B, b...... a' 0 b, are continuous. 
6.3. Tensor Products of C*-Algebras 191 Proof. Since cp is a linear map between Banach spaces, we may invoke the closed graph theorem. Thus, to show that <p is continuous we need only show that if a sequence (an) converges to 0 in A and the sequence (cp(a n )) converges to C in A 01' B, then c = O. We may suppose that an and b ' are positive (by replacing an by a an and b ' by b ' * b ' if necessary). Hence, c > o. If r is a positive linear functional on A 01' B, then the linear functional p: A --+ C, a  r( a 0 b'), is positive, hence continuous. Consequently, r( c) = lim n -+ oo r( an 0 b ' ) = lim n -. oo p( an) = 0, since lim n -. oo an = o. Since r is an arbitrary positive linear functional on A 0...,. B, Theorem 3.3.6 implies that c = O. Therefore, cp is continuous. Similar reasoning shows that 'ljJ is continuous. 0 6.3.5. Theorem. Let A, B be non-zero C*-algebras and suppose that "Y is a C*-nonn on A 0 B. Let (H,7r) be a non-degenerate representation of A 01' B. Then there exist unique *-homomorphisms cp: A --+ B(H) and 'ljJ: B --+ B(H) such that 7r(a 0 b) = cp(a)'ljJ(b) = 1jJ(b)cp(a) (a E A, b E B). Moreover, the representations (H, cp) and (H, 1jJ) are non-degenerate. Proof. Let Ho = 7r(A 0 B)H. If z E Ho, it can be written in the form n z = L 7r(a l 0 bi)(xi). i=l Suppose we have two such expressions; that is, z can be written n m z = L 7r(ai 0 bi)(Xi) = L 7r(Cj Q9 dj)(Yj), i=1 j=1 (1) where ai, Cj E A, b i , d j E B, and Xi, Yj E H. If (Vp,)p,EM is an approximate unit for B and a E A, then n m 7r(a 0 vp,)(z) = L 7r(aai 0 Vp,bi)(Xi) = L 7r(acj 0 vp,dj)(Yj), i=1 j=1 so in the limit (using Lemma 6.3.4), n m lim7r(a 0 vp,)(z) = L 7r(aai 0 bi)(Xi) = L 7r(acj 0 dj)(Yj). p, . . 1=1 }=1 
192 6. Direct Limits and Tensor Products We can therefore well-define a map cp(a): Ho -+ Ho by setting cp(a)(z) = E  1 7r(aai Q9b i )(Xi), if z is as in Eq. (1). Since cp(a)(z) = lim ll 7r(aQ9v ll )(z), it is clear that cp( a) is linear. By Lemma 6.3.4 there exists a positive number M (depending on a) such that 117r(a Q9 b)1I < Mllbll (b E B), so IIcp(a)(z)1I < Mllzil. Hence, cp(a) is bounded. Since Ho is dense in H (because (H,7r) is non-degenerate), we can therefore extend cp(a) uniquely to a bounded linear map on H, also denoted by cp(a). Suppose now that (u A ) is an approximate unit for A. Reasoning as before, for all b E B we have limA 7r(U A Q9 b)(z) = E - l 7r(ai 0 bbi)(Xi), if z is as in Eq. (1). We can therefore well-define a map "p(b): Ho -+ Ho by setting 1jJ(b)(z) = El 7r(ai Q9bb i )(xi) = limA 7r(U A Q9b)(z). The linear map 1jJ(b) is bounded and extends uniquely to a bounded linear map on H, also denoted by "p( b). A routine verification shows that the maps cp: A -+ B(H), a  cp(a), and "p: B -+ B(H), b  "p(b), are *-homomorphisrns, and that 7r(a Q9 b) = c.p(a)"p(b) = 'ljJ(b)cp(a). Now suppose that cp': A -+ B(H) and "p': B -+ B(H) are another pair of *-homomorphisrns such that 7r(a Q9 b) = cp'(a)"p'(b) = 1jJ'(b)cp'(a) for all a and b. Suppose that z E H is such that cp'(a)(z) = 0 (a E A). Then 7r( a Q9 b)(z) = 0 for all a E A and b E B, and therefore 7r( c)(z) = 0 for all c E A Q9-y B. By non-degeneracy of (H, 7r), we have z = O. Thus, cp' is non-degenerate. Similarly, 1jJ' is non-degenerate. In particular, c.p and "p are non-degenerate. If ( U A) and (V II) are approximate units as above, then the nets (cp' ( U A)) and (1jJ'( VII)) converge strongly to id H (by non-degeneracy of cp' and 1/;'). Hence, (7r(aQ9v ll ))1l converges strongly to both cp'(a) and cp(a) for all a E A, so cp' = cp. Similarly, (7r(u A 0b))A converges strongly to both "p'(b) and "p(b) for all b E B, so "p' = "p. 0 We denote the maps cp and "p in the preceding theorem by 7rA and 7rB, respectively. 6.3.6. Corollary. Let A and B be C*-algebras and let, be a C*-seminorm on A 0 B. Then ,( a 0 b) < lIalill bll (aEA, bE B). Proof. Let 6 = max(" 11.11.), so 6 is a C*-norm on A 0 B. Let (H,7r) be the universal representation of A06 B. This is faithful and non-degenerate, so Theorem 6.3.5 applies. If a E A and b E B, then 7r(a 0 b) = 7rA(a)7rB(b). 
6.3. Tensor Products of C*-Algebras 193 Hence, 6(a 0 b) = 117r(a 0 b)11 < II7rA(a)IIII7rB(b)1I < lIalillbll, so ,(a 0 b) < 6(a0b) < Ilallllbli. 0 Let A and B be C*-algebras. Denote by r the set of all C*-norms , on A 0 B. We define lIeli max = sUP-yEr ,(e) for each e E A 0 B. If e = Ej=l aj0 b j with aj E A and b j E B, then for any, E r we have ,(e) < Ej=l,(aj 0 b j ) < Ej=l lIajllllbjll by Corollary 6.3.6. Hence, IIcll max < 00. I t is readily verified that 1I.lImax: A 0 B  R+, c  IIcll max , is a C*-norm, called the maximal C*-norm. We denote by A 0max B the C*-completion of A 0 B under this norm, and call A 0max B the maximal ten30r product of A and B. If , is a C*-seminorm, then, < 1I.lImax (because max( ,,11.11.) < 11.11 max, since max(" 11.11.) is a C* -norm). The maximal tensor product has a very useful universal property: 6.3.7. Theorem. Let A, B, and C be C*-algebras and suppose that cp: A  C and 'ljJ: B  Care *-homomorphisms such that every element of cp(A) commutes with every element of'ljJ(B). Then there is a unique *-homomorphism 7r: A 0max B  C such that 7r(a 0 b) = cp(a)'ljJ(b) (a E A, b E B). Proof. Uniqueness is clear. By Remark 6.3.2 there is a *-homomorphism 7r: A0B  C satisfying the equation in the statement of the theorem. The function ,: A 0 B  R + , c  117r( c) II, is a C* -seminorm. Hence, ,(c) < II climax for all c E A 0 B. Therefore, 7r is a norm-decreasing *-homomorphism, and so extends to a norm-decreasing *-homomorphism on A 0max B. 0 We say a C*-algebra A is nuclear if, for each C*-algebra B, there is only one C* -norm on A 0 B. 6.3.3. Remark. If a *-algebra A admits a complete C*-norm 11.11, then it is the only C*-norm on A. For if , is another C*-norm on A, and B denotes the C*-completion of A with respect to " then the inclusion (A, 11.11)  (B,,) is an injective *-homomorphism and therefore isometric, so , = 11.11. 
194 6. Direct Limits and Tensor Products 6.3.1. Ezample. For each n > 1, the C*-algebra Mn(C) is nuclear. The reason is that for each C*-algebra A, the *-algebra Mn(C) Q9 A admits a complete C*-norm. This is seen by showing that the unique linear map 7r: Mn(C) Q9 A  Mn(A), such that 7r((Aij)ij 0 a) = (Aija)ij for (Aij) E Mn(C) and a E A, is a *-isomorphism (this is a routine exercise). We are going to show that all finite-dimensional C*-algebras are nuclear and for this we shall need to determine the structure theory for such alge- bras. This is given in the following. 6.3.8. Theorem. If A is a non-zero finite-dimensional C*-algebra, it is *-isomoIphic to M n1 (C) EB... EB Mnlc(C) for some integers nl,...' nk. Proof. If A is simple, the result is immediate from Remark 6.2.1. We prove the general result by induction on the dimension m of A. The case m = 1 is obvious. Suppose the result holds for all dimensions less than m. We may suppose that A is not simple, and so contains a non-zero proper closed ideal I, and we may take I to be of mininum dimension. In this case I has no non-trivial ideals, so I is *-isomorphic to M n1 (C) for some integer n1. Hence, I has a unit p, so I = Ap and p commutes with all the elementS- of A. Also, A( 1 - p) is a C* -subalgebra of A and the map A  ApEBA(l- p), a  (ap,a(l- p)), is a *-isomorphism. Since the algebra A(l - p) has dimension less than m, it is *-isomorphic to M n2 (C) EB. . . EB Mnlc (C) for some n2,. . . , nk (inductive hypothesis). Hence, A is *-isomorphic to M n1 (C) EB... EB Mnlc(C). 0 6.3.9. Theorem. A finite-dimensional C*-algebra is nuclear. Proof. Let A be a finite-dimensional C*-algebra, which we may suppose to be the direct sum A = M n1 (C) EB ... EB Mnlc(C). Let B be an arbitrary C*-algebra. A routine verification shows that the unique linear map 7r: A  B  (M n1 (C) Q9 B) EB . . . EB (M nlc (C) Q9 B), such that 7r((al,...' ak)  b) = (al  b,..., ak  b) for all aj E M nj (C) and b E B, is a *-isomorphism. Hence, A  B admits a complete C*-norm, so it admits only one C*-norm. This shows that A is nuclear. 0 The next result suggests that nuclear algebras exist in abundance. 
6.3. Tensor Products of C*-Algebras 195 6.3.10. Theorem. Let S be a non-empty set of C*-subalgebras of a C*-algebra A which is upwards-directed (that is, if B, C E S, then there exists DES such that B, C C D). Suppose that uS is dense in A and that all the algebras in S are nuclear. Then A is nuclear. Proof. Let B be an arbitrary C*-algebra and suppose {3, I are C*-norms on A 0 B. Set C = U DESD 0 B (we may regard D 0 B as a *-subalgebra of AQ9B for each DES). Then C is a *-subalgebra of AQ9B and it is clear that C is dense in the C*-algebras AQ9pB and A0-rB. Now {3 = ,on D0B for each DES, by nuclearity of D, so (3 = I on C, and therefore the identity map on C extends to a *-isomorphism 7r: A Q9p B -+ A Q9-r B. If a E A and b E B, then there is a sequence (an) in uS such that a = lim n -+ oo an in A, so a 0 b = lim n -+ oo an 0 b in A Q9p B and in A Q9-r B, and therefore 7r( a Q9 b) = lim n -+ oo 7r( an 0 b) = lim n -+ oo an 0 b = a 0 b, where convergence is with respect to 'Y. Therefore, 7r = id on A 0 B. Hence, for any e E A 0 B, we have I(e) = I(7r(C)) = (3(e), so I = (3 on A 0 B. Therefore, the algebra A is nuclear. 0 6.3.2. Ezample. If H is a Hilbert space, K(H) is a nuclear C*-algebra. To see this, suppose that e1,..., en are orthonormal vectors in H. If p = Ej=l ej 0 ej, then p is a projection, p E K(H), and the map n c.p: Mn(C) -+ pI{(H)p, (Aij)  L Aijei 0 ej, i,j=l is a *-isomorphism. We show surjectivity only: If u E pI«H)p, then u = pup n = L(ei0 e i)u(ej0 e j) i,j=l n = L (u( ej), ei)ei 0 ej i,j=l = c.p(((u(ej), ei))ij). Since pK(H)p is finite-dimensional, it is nuclear (Theorem 6.3.9). If E is an orthonormal basis for H, let I be the set of all finite non- empty subsets of E made into an upwards-directed poset by setting i < i if i C j. For i E I, let Pi = EXEix 0 x and Ai = PiK(H)Pi. Each Ai is therefore a nuclear C*-algebra, and the set S = {Ai liE I} is upwards- directed. If u is a finite-rank operator on H, we can write u = El Xk 0 Yk for some x k , Y k E H. Therefore, m UPi = L Xk 0 Pi(Yk), k=l 
196 6. Direct Limits and Tensor Products so limi UPi = 2: ;;- 1 xk 0 Yk = u, since limi Pi(Y) = Y (y E H). From this it follows by norm-density of the finite-rank operators in K(H) that limi UPi = U for all U E K(H). Therefore, limiPiuPi = u. Hence, uS is dense in K(H). By Theorem 6.3.10 K(H) is a nuclear C*-algebra. 6.3.11. Theorem. All AF-algebras are nuclear. Proof. If A is an AF-algebra, then it has an increasing sequence (An)=1 of finite-dimensional C*-subalgebras such that UnAn is dense in A. Each An is nuclear by Theorem 6.3.9, so A is nuclear by Theorem 6.3.10. 0 6.4. Minimality of the Spatial C*-Norm As the title of the section suggests, we show here that the spatial C*-norm on a tensor product of C*-algebras is the least C*-norm. Along the way we shall also show the important result-due to Takesaki-that abelian C*-algebras are nuclear. We begin with a result on approximate units that will be used at a ntUl1ber of points in the sequel. 6.4.1. Lemma. Let A and B be C*-algebras and suppose that "Y is a C*-norm on A 0 B. Then A 01' B admits an approximate unit of the form (u..x 0 V..x)..xEA, where (U..x)..xEA and (V..x)..xEA are approximate units for A and B, respectively. Proof. Let (U..x)..xEA and (VIl)IlEM be approximate units for A and B, respectively. Write (A,J.L) < (A',J.L') in A x M if A < A' and J.L < /-l'. The relation < is reflexive, transitive, and upwards-directed, and if we set U(..x,Il) = U..x and V(..x,Il) = v ll ' then a routine argument shows that the net (u' ( ..x ) 0v ( '..x } )(..x Il)EAxM is an approximate unit for A01'B of the required ,Il ,Il' type. 0 6.4.1. Remark. Suppose that A and Bare C*-algebras and suppose also that (H..x, cp ..x) ..xEA and (I{ Il' 'l/J Il ) IlE M are families of representations of A and B, respectively. Set (H,cp) = fB..xEA(H..x,cp..x) and (I{,'l/J) = fBIlEM(I<Il,'l/JIl). It is readily verified that there is a unique unitary u:HI{ EBH..xKIl' ..xEA IlEM such that for all x = (X..x)..xEA E Hand Y = (YIl)IlEM E 1<, u(x 0 y) = (x..x 0 YIl)..x,Il. 
6.4. Minimality of the Spatial C*-Norm 197 For each element e E A 0 B , (cp  1/J)(e) = u.( EB (CPA  1/J#l)(e))u. AEA #lEM (To see this, show it first for e = a 0 b.) It follows that lI(cp  )(e)1I = sup II(CPA  #l)(e)lI. AEA #lEM 6.4.2. Theorem. If A, B are non-zero C*-algebras and e E A 0 B, then lie II. = sup II(CPT  cpp)(e)lI. TES(A) pES(B) Proof. If (H, cp) and (K, ) are the universal representations of A and B, respectively, then lie II. = II(cpQ$)1/J)(e)11 by definition of the norm 11.11.. Since (H, cp) = E1JTES(A)( H T, CPT) and (Ii, ) = E1J pES(B) (H p, c.p p), the theorem follows from Remark 6.4.1. 0 6.4.3. Corollary. If r, p are states on C*-algebras A, B, respectively, then r  p is continuous on A 0 B with respect to the spatial C*-norm. Proof. If e E A 0 B, then (r 0 p)(e) = ((CPT  cpp)(e)(xT 0 xp),x T 0 x p ) (1) (to see this, first show it for e = a 0 b). Since II(CPT Q$) cpp)(e)1I < Ileli. by Theorem 6.4.2, we have I(r 0 p)(c)1 < Ileli. by Eq. (1). Thus, r 0 p is continuous with respect to 11.11.. 0 6.4.2. Remark. If (H, cp) and (I{,) are representations of C*-algebras A and B that are unitarily equivalent to representations (H', cp') and (K', 1/J'), respectively, then there exists a unitary u: H  K -+ H'  K ' such that for all e E A 0 B we have (cp'  /)( e) = u( cp  1/J)( e)u.. (The proof is routine.) 6.4.4. Theorem. If (H, cp) and (J{, 1/J) are arbitrary representations of C*-algebras A and B, respectively, and e E A 0 B, then II(cp Q$) )(e)11 < lIell.. Proof. Let H' = [cp(A)H] and J{' = [1/J(B)J<]. Then H'  J{' = [(c.p  )(A 0 B)(H  I{)], 
198 6. Direct Limits and Tensor Products and a routine verification shows that (cp  1jJ )( c) H' (sK' = (cp H'  1jJ K' )( C) for all e E A 0 B. Therefore, 1I(<p  1jJ)(e) II = lI(cp  )(e)H'(sK,1I = II(CPH'  K' )(e)lI. Thus to prove the theorem we may suppose that <p and 1jJ are non-degenerate (replacing them with the non-degenerate representations <PH' and K' if necessary). Hence, we may write cP and  as direct sums of cyclic repre- sentations. By Theorem 5.1.7 each non-zero cyclic representation of a C*-algebra is unitarily equivalent to a representation of the form (H T , <PT) for some state r of the algebra. Replacing (H,cp) and (K,) by unitar- ily equivalent representations if necessary, we may suppose that (H, cp) = ffi )..EA (H T" , CPT" ) for some index set A with r).. E S( A) for all A, and likewise we may suppose that (I{,1jJ) = ffip,EM(HpJj,cppJj) for an index set M with pp, E S(B) for all I" (we can do this by Remark 6.4.2). For all e E A 0 B, II(cp  )(e)1I = sup II(CPr"  CPPJj )(e)1I )..EA p,EM by Remark 6.4.1, and therefore, II(cp  1jJ)(e)11 < sup II(CPr  cpp)(e)11 = lIell., TES(A) pE S( B) by Theorem 6.4.2. o We shall use the following elementary observation In the proof of Theorem 6.4.5. 6.4.3. Remark. If p is a rank-one projection of Mn(C), then there exist scalars AI'...' An such that p = (Ai.x j )ij. To see this, write p = x 0 x for some x E c n . If el, . . . , en is the canonical orthonormal basis of C n , then x = 2:  1 Aiei for some scalars AI, . . . , An. Since ei 0 ej is the matrix with all entries zero except for the (i,j)-entry, which is 1, and since p = 2:j=1 Ai.xjei 0 ej, we have p = (Ai.xj)ij. 6.4.5. Theorem. Let r, p be positive linear funetionals on C*-algebras A, B, respectively. Tben tbe linear functional r 0 p on A 0 B is positive. Proof. If c E A 0 B, then we have to show (r 0 p)(c.c) > O. We write c = 2: j- 1 aj 0 b j , where a1, . . . , an E A and b 1 , . . . , b n E B. Then n n (r 0 p)(c.c) = (r 0 p)( L aiaj 0 bibj) = L r(a;aj)p(bibj). i,j=l i,j=l 
6.4. Minimality of the Spatial C*-Norm 199 Now if AI, . . . , An E C, then n n n L p(bibj)iAj = p((L Aibi)*(L Ajb j )) > 0, i,j=1 i=1 j=1 since p is positive. Hence, the matrix u == (p( bi b j ) )i,j is a positive element of Mn(C), so it can be diagonalised, and therefore, it can be written in the form u == E=1 tjpj with t},..., t n E R+ and PI,... ,Pn rank-one projections in M n ( C). Thus, to show that (r 0 p ) ( c* c) > 0, it is sufficient to show that Ej=1 r(aiaj)Pij > 0 for each rank-one projection P == (Pij)ij in Mn(C). By Remark 6.4.3 any such projection P is of the form P == (iAj)ij for some scalars AI, . . . , An E C. Hence, n n L r(aiaj)Pij == L r(aiaj)iAj i,j=1 i,j=1 n n == r((L Aiai)*(L Ajaj)) i=1 j=1 > 0, since r is positive. o 6.4.6. Theorem. Let A and B be C*-algebras and suppose that, is a C*-norm on A 0 B. If r, p are states on A, B, respectively, and r  p is continuous with respect to " then r 0 p extends uniquely to a state w on A 0')' B. Proof. Since A 0 B is a dense vector subspace of A 0')' B, it is clear that r 0 p extends uniquely to a continuous linear functional w on A 0')' B-the point of the theorem is that w is positive and of norm 1. If c E A 0')' B, then there is a sequence (c n ) of elements of A 0 B converging in the norm, to c. Hence, c*c = limnoo c;c n , so w(c*c) == limnoo(r 0 p)(c;c n ). Since (r 0 p)(c;c n ) > 0 for all n by Theorem 6.4.5, we have w( c*c) > o. Thus, w is positive. By Lemma 6.4.1 we may choose an approximate unit for A 0')' B of the form (u..x 0 V..x)..xEA, where (U..x)..xEA and (V..x)..xEA are approximate units for A and B, respectively. Applying Theorem 3.3.3, Ilwll = lim..xw(u..x 0v..x) = lim..x r( u..x)p( v..x) == 1, since lim..x r( u..x) == IIrll = 1 and lim..x p( v..x) == Ilpll = 1. Therefore, w is a state of A 0')' B. 0 We denote the state w in Theorem 6.4.6 by r 0')' p. 6.4.7. Theorem. Let A, B be C*-algebras and suppose, is a C*-nonn on A 0 B. Let r, p be states on A, B, respectively, such that r 0 p is continuous 
200 6. Direct Limits and Tensor Products with respect to "y. Then there exists a unitary u: H T 0 Hp -+ HTfg)..,p such that for all c E A 0 B , <PT..,p(C) = U(<PT 0 <pp)(c)u*. Proof. Let w = r 01' p and 1r = CPT  <pp, and let y be the unit vector X T 0 X p in H T  Hp. We claim that for all c E A 0 B we have (<Pw(c)(Xw),Xw) = (1r(c)(y),y). (2) To show this, we may suppose that c = a 0 b, where a E A and b E B. Then (<Pw(c)(Xw), xw) = w(a 0 b) = r(a)p(b) = (<PT(a)(xT),XT)(<pp(b)(xp),x p ) = ((<PT(a) 0 <pp(b))(x T 0 x p ), X T 0 x p ) = (1r(c)(y), y). Let Ho = <Pw(A 0 B)x w and I(o = 7r(A 0 B)y. The map Uo: I(o -t Ho, 1r(c)(y)  <Pw(c)xw, is well-defined (by Eq. (2)), linear, and isometric (again by Eq. (2)). Hence, by density of Ko in H T 0 Hp and Ho in Hw, we can extend Uo uniquely to a unitary u: H T 0 Hp -+ Hw. A routine verification shows that <Pw(c) = u1r(c)u* for all c E A 0 B. 0 6.4.8. Theorem. Let A and B be non-zero C*-algebras, and suppose that c E A 0 B. Then Ilcll = sup sup rES(A) dEAB pES(B) (Tp)(d d»O (r 0 p)(d*c*cd) (r 0 p)( d*d) . Proof. If w is a state of A 0* B, then lI<Pw(c)112 = sup dEAfg)B w(d- d»O w( d*c*cd) w(d*d) , because lid + N w l1 2 = w(d*d) and <Pw(A 0 B)x w is dense in Hw. By Theorem 6.4.2 we have IIcll = sup II(<PT 0 <pp)(c*c)II. TES(A) pES(B) 
6.4. Minimality of the Spatial C*-Norm 201 Applying Corollary 6.4.3 and Theorems 6.4.6 and 6.4.7, we have IICPrII.II. p(d)11 = II(CPr  cpp)(d)lI, for all T E SeA), p E S(B), and d E A 0 B. Putting these equations together, we get 2 * (T 0p)(d*e*cd) lIell* = sup l!'Pr@uu.p(e e)1I = sup sup ( )(d*d) . rES(A) rES(A) dEAB T 0 P pES(B) pES(B) o 6.4.9. Theorem. If A and Bare C*-algebras, the restriction to A 0 B of the spatial C*-nonn on A 0 iJ is the spatial C*-nonn on A 0 B. Proof. Let, be the restriction to A 0 B of the spatial C* -norm on A 0 iJ. Applying Theorem 6.4.8, we get for e E A 0 B, ,(C)2 = sup sup rES(A) dEAB pES(B) (rp)(d d»O (T 0 p)(d*e*ed) (T 0 p)( d* d) , and also, II 11 2 (T 0 p)( d*e*ed) e * = sup sup . rES(A) dEAB (T 0 p)( d*d) pES(B) (rp)(d d»O Using the fact that each p E S(B) has a unique extension p in S(B) (Theorem 3.3.9), we therefore have ,(c) > lIell*" Now let (H, cp) and (I<,,,p) be the universal representations of A and iJ, respectively. Let 1/J B denote the restriction of 1/J to B. If c E A 0 B, then (cp 0 "p )( c) = (cp @ "p B)( c). Since ,(c) = II (<p  "p)( c) II (by definition of the spatial C*-norm on A 0 B), and since II (cp  "p B)( c) II < II ell * by Theorem 6.4.4, we have ,(c) < lIell*. Therefore, , = 11.11*. 0 6.4.10. Theorem. If A and Bare C*-algebras, if B is non-unital, and if , is a C*-norm on A 0 B, then there is a C*-norm on A 0 B extending ,. Proof. Let (H, 7r) be a faithful non-degenerate representation of A 01' B. Since 7r(a 0 b) = 7rA(a)7rB(b) = 7rB(b)7r A(a) (3) for all a and b, it is clear that 7r A and 7r B are injective, because 7r is. Extend 7rB to a unital *-homomorphism 7rB: iJ  B(H). We claim that 7rB is injective. To see this, let b E B and A E C and suppose that 7rn(b + A) = o. If A i= 0 this implies that 7rB(b') = 1 for b' = -bl A E B. Since 
202 6. Direct Limits and Tensor Products 7r B is injective, the element b' is therefore a unit for B, contradicting our assumption that B is non-unital. Hence, ,,\ = 0 and therefore, b = o. Thus, 7r is injective as claimed. By Eq. (3), the elements of 7rA(A) commute with the elements of 7r(iJ). It follows from Remark 6.3.2 that we have a *-homomorphism 7r': A 0 B  B(H) extending 7r. Since ,(e) = 1/7r(e)1I for all c E A 0')' B by injectivity of 7r, to prove the theorem we have only to show that 7r' is injective (since in this case e  117r'(c)1I is a C*-norm on A0B extending ,). Suppose that d E ker(7r'). If e E A0B, then de E A0B and 7r(dc) = 0, so dc = 0, since 7r is injective. Let B = 7r A 0 7rB. Then B( d)B( c) = o. This shows that B( d) is equal to zero on I<o = B( A 0 B)( H  H). But it is easily verified that [<0 is dense in H  H (use the non-degeneracy of 7r A and 7rB). Hence, B( d) = o. Since 7r A and 7r are injective, so is B by Theorem 6.3.3. Therefore, d = 0, and 7r' is injective. 0 6.4.11. Lemma. Let A and B be C *-alge bras, and suppose that U and v are unitaries in A and B, respectively. Then the unique *-isomorphism 7r on A 0 B, such that 7r(a 0 b) = uau* 0 vbv* for all a E A and b E B, is an isometry for any C*-norm , on A 0 B. Proof. Since 7r has inverse the unique *-isomorphism 7r' on A 0 B such that 7r' (a 0 b) = U * au 0 v* bv for all a and b, it suffices by symmetry to show that 7r is norm-decreasing. Applying Lemma 6.4.1, we may choose an approximate unit for A 0')' B of the form (u A 0 v A)AEA, where (UA)AEA and ( v A) AEA are approximate units for A and B, respectively. Let W = U 0 v and W A = U A 0 VA' and observe that 7r(e) = WCw* for all e E A 0 B (we are regarding A 0 B as a *-subalgebra of A 0 B). If a E A and b E B, then uau* = limA uuAauAu* in A and vbv* = limA vvAbvAV* in B, so by continuity of the bilinear map A x B  A 0')' B, (a', b')  a' 0 b', (this is given by Corollary 6.3.6), we have w(a 0 b)w* = uau* 0 vbv* = lim uuAauAu* 0 vvAbvAv* A = lim wwA(a 0 b)wAw* A in A 0')' B. It follows that wew* = limA wwAewAw* for all e E A 0 B; that is, 7r( c) = limA 7r( w A ew A). Hence, ,(7r(e)) = lim,(7r(w A ew A )) A = lim,( ww Aew..\ w*) A < sup,( ww A),( e),( w AW*) AEA < sup lIuuAllllvvAII,(c)lIuAu*IIIlvAv*1I AEA < ,(e). 
6.4. Minimality of the Spatial C*-Norm 203 This proves the lemma. o If A, B are C*-algebras and, is a C*-norm on A 0 B, we denote by S.., the set of all pairs (T, p) E PS( A) x PS( B) such that T 0 P is continuous on A 0 B with respect to ,. The set S.., plays a fundamental role in the proof that abelian C*-algebras are nuclear and that the spatial C*-norm is minimal. 6.4.12. Theorem. Let A, B be C*-algebras and let, be a C*-norm on A 0 B. Then S.., is closed in PS(A) x PS(B) (where the sets PS(A), PS(B) are endowed with the weak* topologies). Moreover, if u, v are unitaries in A,.8, respectively, and (T, p) E S.." then (T'\ pV) E S..,. Proof. If 1r: A0B -+ A0B is the unique *-isomorphism such that 1r( a 0 b) = uau* 0 vbv* for all a and b, then T U 0 pV = (T 0 p)1r, so continuity of T U 0 pV with respect to , follows from Lemma 6.4.11 and the continuity of T 0 P with respect to ,. The proof that S"( is closed is a routine argument (use Theorem 6.4.6 to show that I(T 0 p)(c)1 < ,(c) (c E A 0 B) if (T,p) E S..,). 0 Let A and B be C*-algebras and, a C*-norm on A 0 B. If W is a state of A 0.., Band 1r = 'Pw, we define states WA and WB on A and B, respectively, by setting WA(a) = (1rA(a)(xw), xw) and wB(b) = (1rB(b)(xw), xw). If (U.x).xEA is an approximate unit for A, then Xw = lim.x1rA(U.x)(Xw) (by non-degeneracy of (H, 1rA)), so for all b E B, wB(b) = limw(u 0 b), .x sInce wB(b) = (1rB(b)(xw),xw) = lim(1rB(b)1r A( u.x)( x w ), xw) .x = lim(7r( U.x 0 b)( xw), xw) .x = limw( U.x 0 b). .x Similarly, if (VP.)P.EM is an approximate unit for B, then for all a E A, WA(a) = limw(a 0 Vp.). #l (In particular, if B is unital, then wA(a) = w(a 01).) If (T,p) E S.., and w = T 0"( p, then T = WA and p = wB. We prove this for T (the proof for pis similar): wA(a) = limp.w(a0vll) = lim ll T(a)p(v ll ) = T( a), because limp. p( v II) = 1, since p is a state on B. 
204 6. Direct Limits and Tensor Products 6.4.13. Theorem. Let A and B be C*-algebras and suppose that A or B is abelian. Suppose that 'Y is a C*-nonn on A Q9 B and let (r, p) E S-y. Then r -y p is a pure state of A 0-y B. Proof. We show this in the case that A is abelian. Let W = r Q9-y p, and (H,7r) = (Hw,CPw). Let I{ = [7rA(A)x w ]. Then K is a closed vector subspace of H invariant for 7rA(A), the map 'ljJ: A  B( K), a....... 7r A (a)K, is a *-homomorphism, and the vector Xw is a unit cyclic vector for the representation (K,'ljJ) of A. Since ((a)(xw),xw) = r(a) = (CPr(a)(xr),xr) for all a E A, the representations (I{, 'ljJ) and (Hr, CPr) are unitarily equiva- lent, by Theorem 5.1.4. Because r is a pure state, (H r, CPr) is an irreducible representation, and therefore (K, ) is also irreducible. Hence, (A)' = C1 (Theorem 5.1.5). Since A is abelian, 'ljJ(A) C 'ljJ(A)', so if a E A, there is a scalar A such that (a) = AI. Hence, r(a) = ((a)(xw),xw) = (Axw,xw) = A. Therefore, (a) = r(a)l. We claim now that 7rA(a)7rB(b) = r(a)7rB(b), for all a E A and b E B. To see this it suffices to show that 7rA(a)7rB(b)(x) = r(a)7rB(b)(x), (4) for all x E H of the form x = 7r A (a ' )7r B (b ' )( xw) (since the set of such elements has dense linear span H). However, 7r A ( a ) 7r B ( b ) 7r A ( a I ) 7r B ( b ' ) ( X w) = 7r B ( bb ' )  ( a ) 7r A ( a ') ( X w ) = 7rB(bb ' )r(a)7rA(a ' )(xw) = r(a)7rB(b)7rA(a / )7rB(b ' )(xw), so Eq. (4) holds, and the claim is proved. It follows directly that 7r(A 0-y B) = 7rB(B). Hence, Xw is a cyclic vector for (H,7rB). Since (7rB(b)(x w ), xw) = pCb) = (cpp(b)(xp),x p ) for all b E B, it follows from Theorem 5.1.4 that (H,7rB) and (Hp, cpp) are unitarily equivalent representations of B. Since p is pure, (H p, cP p) is irreducible, so (H, 7r B) is irreducible. Hence, 7r(A 0-y B)' = 7rB(B)' = C1, by Theorem 5.1.5, so (H,7r) is an irreducible representation of A 0-y B. Therefore W = r 0-y p is a pure state of A 0-y B.O 6.4.14. Lemma. Let A and B be C*-algebras and suppose that 'Y is a C*-norm on A 0 B, that W is a pure state of A 0...,. B, and that W A is a pure state of A. Then (WA,WB) E S...,. and W = WA 0...,. WB. Proof. Let (H,7r) = (Hw, CPw) and r = WA, P = WB. Let I{ = [7rA(A)x w ] and let  denote the *-homomorphism A  B(K), a....... 7rA(a)K. 
6.4. Minimality of the Spatial C*-Norm 205 The vector Xw is cyclic for the representation (K, 1/;). Since (1jJ(a)(xw), xw} = r( a) for all a E A, the representations (K, 1jJ) and (H T, CPT) are unitarily equivalent, and since r is pure by hypothesis, (H T, CPT) is irreducible, and therefore so is (K,,,p). Let p be the projection of H onto K. Then p E 7r A (A)' as K is invariant for 71'" A (A). If q is a projection in p7r A (A)' p, then q(H) is a closed vector subspace of K invariant for (K, 1jJ), so q(H) = o or K by irreducibility of (K, 1jJ ). Hence, q = 0 or p. Thus, the von Neumann algebra p7rA(A)'p contains only scalar projections, and since a von Neumann algebra is the closed linear span of its projections, it follows that p7rA(A)'p = Cpo Now 7rB(B) C 7rA(A)', so if b E B, then there exists a scalar A such that p7r B (b)p = Ap. Hence, pCb) == (7rB(b)(x w ),x w ) == (7r B ( b ) p( x w ), p( x w ) ) = (p7r B ( b) p( x w ), x w ) == (Ax w , xw) == A. Thus, p7rB(b)p == p(b)p. If a E A, then w(a 0 b) = (7r(a 0 b)(xw), xw) == (7r A ( a )7r B (b )p( x w ), p( x w )} == (7r A (a )p7rB(b )p( xw), xw) == (7r A( a )p(b )xw, xw) == p(b)(7rA(a)(xw),xw) == p(b)r(a) ==(r0p)(a0 b ). Hence, w extends r 0 p to A 0...,. B, so w == r 0...,. p. By Theorem 6.4.7 there is a unitary u: H T 0 H p  H such that 7r( c) == U(CPT cpp)(c)u* for all c E A0B. Suppose that p is not pure (and we shall get a contradiction). In this case there exists a non-trivial closed vector subspace L of Hp invariant for cpp(B). Set L' == H T  L, so L' is a non- tri vial closed vector su bspace of H T 0 H p invariant for (cp T 0 cP p) ( c) for all c E A 0 B. Hence, L" = u(L') is a non-trivial closed vector subspace of H invariant for 71'"( c) for all c E A 0 B, and therefore for all c E A 0...,. B. This is impossible, because (H,7r) is irreducible (since w is a pure state). Thus, to avoid contradiction we conclude that p is pure. Hence, (WA,WB) E 51'.0 6.4.15. Theorem (Takesaki). Everyabelian C*-algebra is nuclear. Proof. Let A, B be C*-algebras where A is abelian, and suppose that, is a C*-norm on A0B. Let W E PS(A1'B) and set (H,7r) = (Hw, CPw) and 
206 6. Direct Limits and Tensor Products r = WA, P = WB. Since A is abelian and (H,7r') is an irreducible representa- tion of A0..,B, we have 7rA(A) C 7r(A0.., B)' = C1. Hence, if a E A there is a scalar A such that 7rA(a) = AI. Consequently, r(a) = (7rA(a)(xw),xw) = (AXw,X w ) = A, so 7rA(a) = r(a)l. Therefore, r is a multiplicative state on A, so, by Theorem 5.1.6, a pure state. By Lemma 6.4.14 p is a pure state of Band (r,p) E S.., and W = r 0.., p. By Theorem 6.4.7, if c E A 0 B, we have 117r(c)II = II(CPr  cpp)(c)l/. Since ,(c) = sUPwEPS(A@ B) I/CPw(c)1/ (this is got by combining Theorems 5.1. 7 and 5.1.12), therefor; ,(c) = sup II(CPr @ cpp)(c)II. ( r,p)ES.., Thus, if we show that S.., = PS(A) x PS(B), we shall have ,(c) = sup II(CPr @ cpp)(c)II, rEPS(A) pEPS( B) and since the right-hand side of this equation is completely independent of the norm " we shall have shown that A 0 B has a unique C*-norm. Suppose then that S, :/= PS(A) x PS(B). We shall derive a con- tradiction and thus prove the theorem. Since S.., is relatively closed in PS(A) x PS(B) (Theorem 6.4.12), there exist a pair of non-empty, rela- tively weak* open sets U and V in PS(A) and PS(B), respectively, such that S.., n (U x V) = 0. Using Theorem 6.4.12 again, we may and _do sup- pose that U and V are unitarily invariant (if for each unitary u E A we set UU = {T U I T E U} and similary define VV for each unitary v E iJ, then U' = U U UU and V' = U v VV are relatively weak* open unitarily invariant non-empty sets in PS(A) and PS(B), respectively, such that S.., n (U' x V') is empty-thus, we may replace U, V by U', V' if necessary). The sets SA = PS(A) \ U and SB = PS(B) \ V are relatively weak* closed unitarily invariant sets in PS(A) and PS(B), respectively, and since SA =I PS(A) and SB # PS(B), it follows from Theorem 5.4.10 that the closed ideals Si and Sii in A and B, respectively, are non-zero, and therefore con- tain non-zero positive elements a and b, respectively. If (T, p) E S.." then (T 0.., p)(a 0 b) = T(a)p(b) = 0, since either T  U or p  V. How- ever, by Theorem 5.1.11, there is a pure state W E PS(A 0.., B) such that ,(a 0 b) = w(a 0 b), and by the first part of the proof of this theorem, W = T 0.., P for some (T,p) E S'Y' so ,(a 0 b) = 0, and therefore a 0 b = o. Hence, either a or b is zero, a contradiction. 0 We shall need the following elementary topological fact: Suppose that n is a compact Hausdorff space, and U I ,..., Un are open sets such that n = U I U. . . U Un. Then there exist continuous functions hI,. . . , h n from n to [0,1] such that Uj contains the support of h j for allj and hI +.. .+h n = 1 [Rud 1, Theorem 2.13]. 
6.4. Minimality of the Spatial C*-Norm 207 If n is a locally compact Hausdorff space and X a Banach space, Co(n, X) denotes the Banach space of all continuous functions 9 from n to X that vanish at infinity (this means that the function w  IIg(w)1I vanishes at infinity). The operations on Co(n, X) are the pointwise-defined ones and the norm is the supremum norm. (We are particularly interested in Co(n, X) when X is a C*-algebra, in which case Co(n, X) is a C*-algebra also, with the pointwise-defined multiplication and involution.) If f E Co(n), and x E X, denote by fx the element of Co(n, X) defined by setting fx(w) = f(w)x. 6.4.16. Lemma. Let n be a locally compact Hausdorff space and X a Banach space. Then Co(n,X) is the closed linear span of the functions Ix (I E Co(n), x EX). Proof. Let 9 E Co(n, X). Define an extension g of 9 to the one-point compactification n of n, by setting g( 00) = 0 where 00 is the point at infinity. Since 9 is continuous and vanishes at infinity, the function 9 is con tinuous. Let c > o. The set yen) is compact and therefore totally bounded, so there exist elements Xl, . . . , X n E g(n) such that if U j = {w E n IlIg( w) - x j II < £}, then n = U 1 U ... U Un. The sets U 1 ,..., Un are open in n, so, by the ele- mentary topological fact q,uoted before this lemma, there exist continuous functions hI, . . . , h n from n to [0, 1] such that the support of h j is contained in U j for j = 1,. . . , n and hI + . . . + h n = 1. Hence, n n Ilg(w) - L hj(w)xjll = II L hj(w)(g(w) - xj)1I j=l j=l n < Lhj(w)llg(w) -Xjll j=l n < L hj(w)c j=l = £. In particular, II Ej=l h j ( oo)x j II < £, since g( 00) = O. Let fj be the restric- tion to n of hj - hj( 00). Then fj E Co(n) and n n n IIg - Lljxjlloo < Ilg - Lhjxjlloo + II Lhj(oo)xjll < 2£. j=l j=l j=1 
208 6. Direct Limits and Tensor Products This proves the lemma. o Let r! be a locally compact Hausdorff space and A a C*-algebra. Since the map Co(r!) x A  Co(r!, A), (f, a)  fa, is bilinear, it induces a unique linear map 7r: Co(r!) 0 A  Co(r!, A) such that 7r(f 0 a) = fa for all f E Co(r!) and a E A. We call 7r the canonical map from Co(r!) 0 A to Co(r!, A). 6.4.17. Theorem. If r! is a locally compact Hausdorff space and A a C*-algebra, then the canonical map from Co(r!) 0 A to Co(r!, A) extends uniquely to a *-isomorphism from Co(r!) 0* A to Co(r!, A). Proof. Let 7r be the canonical map. It is readily verified that 7f is a *-homomorphism. If e E ker( 7r), write e = Ej=1 fj 0 aj, where f1, . . . , f n E Co (r!) and aI, . . . , an are linearly independent elements of A. Then 7r( e) = 0 implies that Ej=1 fj(w)aj = 0 for all w in r!. By linear independence of al,... ,an, therefore, f1(W) = ... = fn(w) = o. Hence, f1 = ... = fn = 0, so e = o. Therefore, 7r is injective. The function Co(r!)0AR+, eII7r(e)lI, is a C*-norm on Co(r!) 0 A, and by Theorem 6.4.15 it is the only C*-norm on this algebra, so 117r(e)1I = lIell* for all e E Co(r!) 0 A. Hence, 7r extends uniquely to an isometric *-homomorphism 7r': Co(r!) 0* A  Co(r!, A). Since the range of 7r' contains the elements fa for all f E Co(r!) and a E A, it follows from Lemma 6.4.16 that 7f' is surjective. 0 If A and Bare *-algebras, then the unique linear map 8: A0B  B0A, such that 8( a 0 b) = b 0 a for all a E A and b E B, is a *-isomorphism. Hence, if A 0 B admits a unique C*-norm, so does B 0 A. This simple observation is used in the proof of the following theorem. 6.4.18. Theorem. For any C*-algebras A and B, the spatial C*-nonn is the least C*-nonn on A 0 B. Proof. Let "y be a C*-norm on A 0 B. If B is non-unital, then we can ex- tend"Y to a C*-norm on A0B by Theorem 6.4.10, and the spatial C*-norm on A 0 iJ extends the spatial C*-norm of A 0 B by Theorem 6.4.9. Thus, it suffices to prove the theorem in the case that B is unital, and therefore we assume B is unital. We show first that S-y = PS(A) x PS(B). Suppose the contrary (and we shall get a contradiction). As in the proof of Theorem 6.4.15, we can get relatively weak* closed unitarily invariant proper subsets S A and S B of PS(A) and PS(B), respectively, such that S-y C SA x PS(B) U PS(A) x SB 
6.4. Minimality of the Spatial C*-Norm 209 and the ideals Si and Sj; contain non-zero positive elements ao and b o , respectively. Thus, for all (T, p) E S...p ( T 0')' P )( ao 0 b o ) = T( ao )p( b o ) = o. (5) Now let C be the C*-subalgebra of B generated by b o and 1. This is abelian, and therefore nuclear, by Theorem 6.4.15, so A 0 C has a unique C*-norm, and therefore, = 11.11. on A 0 C. Thus, we may regard A 0. C as a C* -subalgebra of A 0')' B. Choose pure states T on A and p on C such that T ( ao) = II ao II > 0 and p( b o ) = II b o II > 0 (this is possible by Theorem 5.1.11 ). Then T 0 P extends to a pure state Wi on A 0. C, by Corollary 6.4.3 and Theorem 6.4.13. It follows from Theorem 5.1.13 that Wi can be extended in turn to a pure state W on A 0...,. B. For each a E A, wA(a) = w(a 0 1), so wA(a) = T(a)p(l) = T(a), and therefore WA = T is a pure state on A. It follows from Lemma 6.4.14 that (WA,WB) E S')' and W = WA 0')' WB. Hence, w(ao 0 b o ) = 0 by Eq. (5), and yet w(ao 0 b o ) = (T 0 p)(ao 0 b o ) = T(ao)p(b o ) > o. This contradiction shows that S')' = PS(A) x PS(B). The states of a C* -algebra are weak* limits of nets of convex com- binations of the zero functional and the pure states (by Theorems 5.1.8 and A.14). Let the positive functionals T, p on A, B, respectively, be con- vex combinations of the zero functional and pure states. Hence, there exist T1, . . . , Tn E {OJ U PS(A) and PI,. . . , pm E {OJ U PS(B) and non-negative numbers t},... , tn, 81,. . . ,8m such that E7:I ti = 1, E  I 8j = 1, and T = E7:I tiTi and p = Ej:1 8jPj. Therefore, the functional n m T0p= LL t i 8 j T i0pj i=1 j=1 is continuous with respect to" because this is the case for each Ti0pj (since S...,. = PS(A) x PS(B)). Hence, if we suppose now that T, p are arbitrary states of A, B, respectively, then there exist nets (T).).EA and (PP,)p,EM of positive linear functionals on A and B, respectively, converging weak* to T and p, respectively, such that IIT).II,lIpp,11 < 1 and T). 0 PP, is continuous with respect to , for all ,,\ E A and I" EM. Reasoning as in the proof of Theorem 6.4.6, T). 0 PP, has a unique extension to A 0')' B which is a positive linear functional of norm IIT).llllpp,ll. Therefore, for all c E A 0 B, I( T). 0 pp,)( c)1 < ,( c). Since (T 0 p)(c) = lim).,p,(T). 0 pp,)(c), therefore I(T 0 p)(c)1 < ,(c) for all c E A 0 B. Hence, T 0 P is continuous with respect to ,. Let D be the unitisation of A 0')' B, let T, p be states of A, B, respec- tively, and let W be the unique state on D extending T 0')' p. If d ED, then the linear functional W d : D -+ C, c  w(d.cd), 
210 6. Direct Limits and Tensor Products is obviously positive. Since ,(e*e)l-e*e > 0, we havew d (,(e*e)l-e*e) > 0; that is, ,(e*e)w d (l) > wd(e*e). Hence, if w(d*d) > 0, we have ,(e)2 > w(d*e*ed)jw(d*d), and therefore, by Theorem 6.4.8, for all e E A Q9 B we have lIell; < ,(e)2. Thus, 11.11* is the least C*-nonn on A Q9 B. 0 6.4.4. Remark. Let A and B be C*-algebras and let, be a C*-nonn on A Q9 B. Then ,(a Q9 b) = lIalillbll. This is true, since lIallllbll = lIa Q9 bll* < ,(a Q9 b) (by Theorem 6.4.18) < II all II bll (by Corollary 6.3.6). 6.4.19. Theorem. If (H,c.p) and (I<,'lj;) are faithful representations of C*-algebras A and B, respectively, then lI(cp  'lj;)(e) II = lIell* (e E A Q9 B). Proof. The function ,: A Q9 B  R+, e  II(cp  1/J )(e)lI, is a C*-norm, since cp'lj;: AQ9B  B(H I<) is injective by Theorem 6.3.3. Hence, ,(c) > Ilell* for all e E A Q9 B by Theorem 6.4.18. However, by Theorem 6.4.4, the reverse inequality also holds, so , = 11.11*. 0 6.5. Nuclear C*-Algebras and Short Exact Sequences We continue our investigation of nuclear C*-algebras. The principal result of this section is Theorem 6.5.3, which asserts that extensions of nuclear C*-algebras by nuclear C*-algebras are themselves nuclear. 6.5.1. Theorem. Let A, B, A', B' be C*-algebras and let c.p: A  A' and 'lj;: B  B' be *-homomorpmsms. Then there is a unique *-homomorpmsm 11": A Q9* B  A' Q9* B' such that 11"( a Q9 b) = cp( a) Q9 'lj;( b) (a E A, b E B). Moreover, if c.p and 'lj; are injective, so is 7r. Proof. Let (H',cp') and (I{','lj;') be faithful representations of A' and B', respectively. Then cp'  'lj;' is isometric on A' Q9 B' for the spatial C* -norm by Theorem 6.4.19. Let 7r = cp Q9 'lj;. Then (cp'  'lj;')7r = cp'cp  'lj;'1/J, so by Theorem 6.4.4, 1I(c.p'1/J')7r(e)1I < lIell*. Hence, 117r(e)lI* = 1I(c.p'1/J')7r(e)1I < lIell* for all e E A Q9 B. If the *-homomorphisms cp and 'lj; are injective, so are cp'cp and 'lj;''lj;. Therefore, by Theorem 6.4.19 II(cp'cp  'lj;''lj;)(e)1I = IIcll* for all e E A Q9 B, so 7r is isometric for the spatial C*-norms. The theorem now follows, since we can extend 7r to a *-homomorphism from A Q9* B to A' Q9* B'. 0 We denote the *-homomorphism 7r in Theorem 6.5.1 by cp Q9* 'lj;. 
6.5. Nuclear C*-Algebras and Short Exact Sequences 211 6.5.1. Remark. If A and Bare C*-subalgebras of the C*-algebras A' and B', respectively, and i and j are the inclusion *-homomorphisms, then by Theorem 6.5.1 the *-homomorphism i 0. j: A 0. B -+ A' 0. B' is injective. Thus, we may regard A 0. B as a C*-subalgebra of A' 0. B'. Let J,A, and B be C*-algebras. Suppose that j: J -+ A is an injective *-homomorphism and 7r: A -+ B is a surjective *-homomorphism, and that im(j) = ker(7r). Then we say that the sequence J 7r O-+J-+A-+B-+O is a short exact sequence of C* -algebras. In this case we also say that A is an extension of B by J. If 7r: A -+ B is a surjective *-homomorphism of C*-algebras, J = ker( 7r), and j: J -+ A is the inclusion map, then J 1r O-+J-+A-+B-+O is a short exact sequence of C*-algebras. 6.5.2. Theorem. Let J, A, B, and D be C*-algebras and suppose that J 7r O-+J-+A-+B-+O is a short exact sequence of C*-algebras. Suppose also that B  D has a unique C*-norm (this is the case if B or D is nuclear). Then o -+ J 0. D j id A 0. D 7r id B 0. D -+ 0 is a short exact sequence of C*-algebras. Proof. Let J = j. id D and 7f = 7r* id D . That J is injective follows from Theorem 6.5.1. The map 7r is surjective, since im(7r) contains 7f(A 0 D) = 7r( A) 0 D = B 0 D. The C* -subalgebra im(J) = im(j) 0. D of A * D is an ideal, since im(j) is a closed ideal of A. Let Q be the quotient C* -algebra (A 0. D)/ im(J), and 'l/J the quotient map from A 0* D to Q. Clearly, 7r(im(J)) = 0, so there exists a unique *-homomorphism 1r': Q -+ B . D such that 7r'1/J = 7f. We shall show that 1r' is a *-isomorphism, and this will imply that ker( 7f) = ker( 'l/J) = im(J), thus proving the theorem. That 7r' is surjective is immediate from the surjectivity of 7f. We show injectivity of 7r' by constructing a left inverse. The map B x D -+ Q, (1r(a), d)  a  d + im(J), 
212 6. Direct Limits and Tensor Products is well-defined, and it is readily verified that it is bilinear, so it induces a unique linear map <p: B Q9 D -+ Q such that <p( 7r( a) Q9 d) = a 0 d + im(J) for all a E A and d E D. It is easily checked that c.p is a *-homomorphism. The function B Q9 D -+ R+, e  max(II<p(e)lI, lIell.), is a C*-norm, so, by the assumption that B Q9 D has a unique C*-norm, we have max(II<p(e)lI, lIell.) = lIell., and therefore, 1I<p(e)1I < lIell., for all e E B 0 D. Hence, <p extends to a *-homomorphism from B 0. D to Q which we shall also denote by <po Now <p7r' = id Q , since for all a E A and d E D we have c.p7r'(aQ9d+im(J)) = <p7f(aQ9d) = <p(7r(a)0d) = aQ9d+im(J). Therefore, 7r' is injective, and the theorem is proved. 0 6.5.3. Theorem. An extension of a nuclear C*-algebra by a nuclear C*-algebra is itself nuclear. Proof. Let J, A, and B be C*-algebras, and suppose that J 7r O-+J-+A-+B-+O is a short exact sequence of C*-algebras, and that J and B are nuclear. We prove that A also is nuclear. Let D be an arbitrary C*-algebra. Since B Q9 D has a unique C*-norm, it follows from Theorem 6.5.2 that the following is a short exact sequence: j 7f o -+ J Q9. D -+ A Q9. D -+ B Q9. D -+ o. We are using j and 7f to denote j Q9.id and 7rQ9.id, respectively. The identity map on A 0 D extends to a *-homomorphism c.p: A 0max D -+ A 0. D, since 11.11. < 11.llmax. We shall have proved the theorem if we show that 1I.lImax = 11.11. on A 0 D, since any C*-norm must lie between 11.11. and 1I.llmax (Theorem 6.4.18). Therefore, we need only show that <p is injective. Let j' denote the unique *-homomorphism from J 0 D to A 0max D such that j'(a 0 d) = j(a) 0 d for all a E J and d E D. By nuclearity of J, the C*-norm J 0 D -+ R+, e  max(llj'(e)lImax, Ilcll.), is the same as the spatial C*-norm 11.11. on J Q9 D. Hence, j' is norm- decreasing for 11.11. and so extends to a *-homomorphism from J 0. D to A 0max D which we shall also denote by j'. Clearly, j = c.pj'. There is a unique *-homomorphism 7r': A 0 D -+ B 0. D such that 7r'(a 0 d) = 7r(a) 0 d for all a E A and d E D. The function A 0 D -+ R+, e  max(II7r'(e)II., lIell max ), 
6. Exercises 213 is a C*-nonn and, therefore, it is dominated by the maximal C*-norm 11.11 max. Hence, 7r' is norm-decreasing for 11.11 max, so 7r' can be extended to a *-homomorphism from A 0max D to B 0* D which we shall also denote by 7r'. Let Q be the quotient algebra of A 0max D by the closed ideal im(j'), and let 1fJ: A 0max D  Q be the quotient map. By a construction sim- ilar to that carried out in the proof of Theorem 6.5.2, there is a unique *-homomorphism 8: B 0* D  Q such that 8(7r(a) ° d) = a ° d + im(j') for all a E A and d E D (this uses nuclearity of B). We therefore get a commutative diagram: J0*D j 0* id A0*D 7r 0* id B0*D ---+ ---+ j' Tep /' 7r' !8 A 0max D 1/; Q.  Now suppose that c E ker( ep). Then 0 = 7fep( c) = 7r' (c), so 0 = 87r'(c) = "p(c). Hence, c = j'(co) for some element Co E J * D, and therefore j(co) = epj'(co) = ep(e) = O. Since j is injective by Theorem 6.5.1, we have Co = 0 and therefore c = j'( co) = o. Thus, c.p is injective and the theorem is proved. 0 6.5.1. Ezample. Let A denote the Toeplitz algebra (the C*-algebra gen- erated by all Toeplitz operators on the Hardy space H 2 having continu- ous symbol). This algebra was investigated in Section 3.5, where it was shown that its commutator ideal is I«H2) (Theorem 3.5.10). The algebras K(H 2 ) and A/I{(H2) are nuclear (by Example 6.3.2 and Theorem 6.4.15, respectively), so by Theorem 6.5.3, A is nuclear. 6. Exercises 1. Let (An, epn)=l and (Bn, "pn)=l be direct sequences of C*-algebras with direct limits A and B, respectively. Let epn: An  A and "pn: Bn  B be the natural maps. Suppose there are *-homomorphisms 7r n : An  Bn such that for each n the following diagram commutes: An epn An+l ---+ ! 7r n ! 7r n+ 1 Bn 1fJn Bn+l. ---+ 
214 6. Direct Limits and Tensor Products Show that there exists a unique *-homomorphism 7r: A -+ B such that for each n the following diagram commutes: An cpn A ----+ ! 7r n !7r Bn 'ljJn B. ----+ Show that if all the 7r n are *-isomorphisms, then 7r is a *-isomorphism. 2. Show that every non-zero finite-dimensional C*-algebra admits a faithful tracial state. Give an example of a unital simple C*-algebra not having a tracial state. 3. Let A be a C*-algebra. A trace on A is a function r: A+ -+ [0, +00] such that r(a + b) = Tea) + r(b) r(ta) = tr(a) r(c*c) = r(cc*) for all a, b E A +, c E A, and all t E R + . We use the convention that 0.( +00) = O. The motivating example is the usual trace function on B(H). Another example is got on Co(R) by setting r(f) = J f dm where f E Co(R)+ and m is ordinary Lebesgue measure on R. Traces (and their generalisation, weights) play a fundamental role, especially in von Neumann algebra theory ([Ped], [Tak]). Let A; = {a E A I r(a*a) < oo}. Show that (a + b)*(a + b) < 2a*a + 2b*b and (ab)* ab < IIal12 b* b, and deduce that A; is a self-adjoint ideal of A. Let AT be the linear span of all products ab, where a, b E A;. Show that AT is a self-adjoint ideal of A. Show that for arbitrary a, b E A, 3 a*b= t 2: ik (b+i k a)*(b+i k a), k=O 
6. Exercises 215 and if a* b is self-adjoint, a*b == [(b + a)*(b + a) - (b - a)*(b - a)]. Let A == {a E A+ I rea) < oo}. Show that Ar is the linear span of A; and A; == Ar n A + . Show that there is a unique positive linear extension (also denoted r) of r to Ar. Show that r ( ab) == r ( ba ) for all a, b E A;, and deduce that this equation also holds for all a E A and bEAr. 4. Show that an AF-algebra admits a sequential approximate unit consist- ing of projections. 5. Let u be a normal operator on a Hilbert space H. Show that there is a commuting sequence of projections on H such that the C*-algebra that they generate contains u. Use this to construct an example of a C*-subalgebra of an AF-algebra which is not an AF -algebra. 6. If A is an AF-algebra, show that Mn(A) is one also. 7. Show that if A and Bare AF-algebras, then A 0* B is an AF-algebra. 8. Show that if A, B, and Care *-algebras, then the unique linear map <.p: (A 0 B) 0 C  B 0 (A Q9 C), such that <.p((a 0 b) 0 c) == b 0 (a 0 c) for all a E A, b E B, and c E C, is a *-isomorphism. Deduce that if A is a nuclear C*-algebra, so is Mn(A). 9. If HI, H 2 , and H3 are Hilbert spaces, show that there exists a unique unitary u: (HI 0 H 2 ) 0 H3  HI 0 (H 2 0 H 3 ) such that U((XI 0 X 2)Q9 X 3) == Xl Q9(X2 Q9 X 3) Show that (X j E Hj, j == 1,2,3). "'" "'" * " " u( (VI Q9 V2) Q9 V3)U == VI 0 (V2 Q9 V3) (Vj E B(Hj), j == 1,2,3). Deduce that if A I ,A 2 , and A3 are C*-algebras, then there exists a unique *-isomorphism 8: (AI 0* A 2 ) 0* A3  Al 0* (A 2 0* A 3 ) such that 8( (al 0 a2) 0 a3) == al Q9 (a2 Q9 a3) (aj E Aj, j == 1,2,3). 10. If A, Bare C*-algebras, show that there exists a unique *-isomorphism (): A 0. B  B Q9* A such that ()( a Q9 b) == b 0 a (a E A, b E B). 
216 6. Direct Limits and Tensor Products 6. Addenda The hyperfinite factor exhibited in Theorem 6.2.5 is of Type 111 if the UHF algebra A is infinite-dimensional. If A is a separable C* -algebra, then A is an AF -algebra if and only if for any aI, . . . , an in A and £ > 0 there is a finite-dimensional C* -subalgebra B of A and there exist b l ,..., b n in B such that Ilaj - bjll < £ for 1 < j < n. A hereditary C*-subalgebra of an AF-algebra is an AF-algebra. If I is a closed ideal in a C*-algebra A such that I and AI I are AF-algebras, then A is an AF-algebra. This was first proved by L. Brown using K-theory. Reference: [Eff]. If A and B are simple C*-algebras, then A * B is simple. Postliminal C*-algebras are nuclear. If I is a closed ideal in a nuclear C*-algebra A, then I and AI I are nuclear. It follows from this and Theorem 6.3.10 that the direct limit of a sequence of nuclear C*-algebras is nuclear. A hereditary C*-subalgebra of a nuclear C*-algebra is nuclear. An arbitrary C*-subalgebra of a nuclear C*-algebra need not be nuclear. If H is an infinite-dimensional Hilbert space, then B(H) is non-nuclear. References: [Lan], [Sak], [Tak]. 
CHAPTER 7 K-Theory of C*-Algebras One of the most important recent developments in C*-algebra theory has been the introduction of homological algebraic methods. Specifically, the theory we investigate in this chapter, the K-theory of C*-algebras, has had some spectacular successes in solving long-open problems. The basic idea of this theory is to associate with each C*-algebra A two abelian groups I{o(A) and K 1 (A), which reflect some of the properties of A. In the first sec- tion of this chapter, we present some elementary results concerning Ko(A), and in the second we use J<o(A) to show how AF-algebras can be classi- fied. In Sections 3, 4, and 5 we establish the basic properties of K-theory, including Bott periodicity. 7.1. Elements of K-Theory Let A be a *-algebra. If a = (aij) and b = (b j k) are, respectively, an m x n matrix and n x p matrix with entries in A, then the product c = ab is an m x p matrix with (i,k)-entry given by Cik = 2: j- l aijb jk . Also, the adjoint a* is the n x m matrix with (j, i)-entry aij. We shall have frequent need to use block matrices, so we shall state here a few elementary results concerning them. Let r = (rl,. . . , r m) and C = (Cl,...' cn) be tuples of positive integers, and suppose that for each integer i, such that 1 < i < m, and j, such that 1 < j < n, we have an ri X Cj matrix A ij with entries in A. The r x C block matrix All A l2 A 1n A 21 A 22 A 2n a= (1) AmI A m2 Amn 217 
218 7. K-Theory of C*-Algebras is regarded in an obvious fashion as an (rl + . .. + r m ) x (Cl + ... + cn) matrix with entries in A. The matrix a* is the C x r block matrix Ail A 2l a* = Ai2 A 22 Ain A 2n A:nl A:n2 A:n n If b is a C x d block matrix, where d = (d l , . . . , d p ), and b is given in block form by Bll B 12 Blp B 2l B 22 B 2p b= Bnl B n2 B np then the product ab is the r x d block matrix G ll G 12 G lp G 2l G 22 G 2p ab = G ml G m2 G mp where G ik = 2::.7=1 AijBjk. In words, to multiply two block matrices, perform the usual matrix multiplication on the corresponding blocks. The proofs of these results are elementary exercises. If the block matrix a in Eq. (1) has m = n and zero off-diagonal entries-that is, A ij is the zero matrix for i =I j-we shall denote a by All ED A 22 ED... ED Amm. We denote by On the n x n matrix all of whose entries are 0, and if A is unital, we denote by In the n x n matrix all of whose entries are zero, except for those on the main diagonal, all of which equal 1. Let A be an arbitrary *-algebra, and set P[A] = U=l {p E Mn(A) I p is a projection}. If p, q E P[A] we say p and q are equivalent, and write p I'V q, if there is a rectangular matrix u with entries in A such that p = u*u and q = uu*. If this is the case, we may suppose that u = uu*u, by replacing u by qup if necessary. It is a straightforward exercise to check that I'V is an equivalence relation on P[A]. If p and q are contained in the same algebra Mn(A), then p I'V q if and only if p and q are Murray-von Neumann equivalent in the sense that we defined this in Section 4.1. 
7.1. Elements of K-Theory 219 7.1.1. Theorem. Let A be a *-algebra, and suppose that p,q,p',q' are projections in P[A]. (1) If p I'V p' and q I'V q', then p EB q I'V p' EI1 q'. (2) p EI1 q I'V q EB p. (3) If p, q E Mn(A) and pq = 0, then p + q I'V P EB q. Proof. Suppose that p I'V p' and q I'V q'. If p = u.u, p' = uu., q = v.v, and q' = vv., then p EB q = w.w and p' EI1 q' = ww., where w = u EI1 v. Therefore, p EI1 q I'V p' EI1 q' . To see that p EB q I'V q EB p, set u = ( 6)' Then p E9 q = u*u and q EB p = uu*. Observe that if p E Mn(A), then p I'V P EB Om, for if u = (p, Onm) where Onm is the n x m matrix all of whose entries are zero, then u*u = p EI1 Om and uu* = p. Suppose now that p, q E M n (A) and pq = o. Set u=(t n oqn)' Then u*u = p EI1 q and uu* = (p + q) EI1 On. (p + q) EB 0 n I'V P EB q. Hence, we have p + q I'V o Let A be a unital *-algebra. We say elements p, q of P[A] are stably equivalent, and we write p  q, if there is a positive integer n such that In EB p I'V In EI1 q. It is easy to check that  is an equivalence relation on P[A]. Observe that if p  p' and q  q', then p EI1 q  p' EI1 q'. For p E P[A], let [P] denote its stable equivalence class, and denote by Ko(A)+ the set of all these equivalence classes. For [P], [q] E Ko(A)+, define (P] + [q] = [pEI1 q]. If there is a possibility of ambiguity, we write [P]A for the stable equiva- lence class relative to the algebra A. 7.1.2. Theorem. If A is a unital *-algebra, then Ko(A)+ is a cancellative abelian semigroup with zero element [0]. Proof. Associativity and commutativity are immediate. It is also clear that [On] is the zero element of 1(0 (A)+ . To show that l{o(A)+ is cancellative, suppose that [P] + [q] = [P] + [r], and we shall show that [q] = [r]. For some integer m, we have p EI1 q EI11 m I'V p E9 r EB 1m. If we suppose that p E Mn(A), then (In - p) EI1 p EB q EI1 1 m I'V (I n -p)EBpEBrE91 m . But (I n -p)EI1p "'-J In, by Theorem 7.1.1, Condition (3), so In EB q EB 1m I'V In E9 r E91m, and therefore 1n+m EB q I'V 1n+m EI1 r. Hence, [q] = [r]. 0 
220 7. K-Theory of C*-Algebras Let N be a cancellative abelian semigroup with a zero element. We define an equivalence relation rv on N x N by setting (x, y) rv (z, t) if x + t = Y + z. Denote by [x, y] the equivalence class of (x, y). The set G(N) of equivalence classes is an abelian group under the well-defined operation [x,y] + [z,t] = [x + z,y + t]. (The inverse of [x, y] is [y, x].) We call G(N) the enveloping or Grothendieck group of N. The map c.p: N  G( N), x...... [x, 0] is a homomorphism (that is, c.p(x + y) = c.p(x) + c.p(y) for all x, yEN), and since it is also injective, we can and do identify N as a subsemigroup of G(N) by identifying x with [x,O]. Hence, G(N) = {x - y I x,y EN}. If 'ljJ: N  G is a homomorphism, where G is an abelian group, then there is a unique homomorphism ,(f: G(N)  G extending "p. The elementary proofs of these results are left as exercises. If A is a unital *-algebra, we define J<o(A) to be the Grothendieck group of I<o(A)+. If c.p: A  B is a *-homomorphism of *-algebras and a = (aij) is an m X n matrix with entries in A, set c.p( a) = (c.p( aij)), so c.p( a) is an m x n matrix with entries in B. If b is an n x p matrix with entries in A, then c.p( ab) = c.p( a )c.p( b). Observe also that c.p( a*) = (c.p( a)) *. If p rv q, then c.p(p) rv c.p(q) and if c.p:A  B is a unital *-homomorphism of unital *-algebras, then p  q =:} c.p(p)  c.p( q). Hence, there is a well- defined map C{)*: J<o(A)+  J<o(B)+ given by setting c.p*[p] = [c.p(p)]. Since c.p(p EB q) = c.p(p) EB c.p(q), we have c.p*([p] + [q]) = c.p*([p]) + c.p*([q]); that is, c.p* is a homomorphism. Hence, there is a unique homomorphism c.p*: J<o(A)  I<o(B) such that C{)*([p]) = [c.p(p)]. If c.p: A  Band 'ljJ: B  C are unital *-homomorphisms of unital *-algebras, then ("pC{))* = "p*C{)*. Also, (id A )* = idKo(A). Thus, we have a covariant functor A ...... J< 0 ( A ) , C{) ...... c.p * , from the category of all unital *-algebras to the category of abelian groups. 7.1.1. Eample. It is easy to check that projections p, q E P[C] are equivalent if and only if they have the same rank. Also, for any p, q E P[C], we have rank(p EB q) = rank(p) + rank(q). Thus, we may define a homomorphism rank:I{o(C)+  Z by setting rank([p]) = rank(p). Hence, we can extend uniquely to get a homomorphism rank: J<o(C)  Z. This function is an isomorphism: It is surjective, since 1 = rank([1 1 ]), and it is injective, since if x E ker(rank) we can write x = [P] - [q], where p, q E P[C] 
7.2. The K-Theory of AF-Algebras 221 are projections of the same rank, and therefore equivalent, from which [P] = [q], and x = o. Motivated by this example, one should think of Ko(A) as a "dimension" group, and think of [P] as the "generalised dimension" of p. 7.1.2. Ezample. A non-trivial algebra may have trivial Ko-group. For instance, if H is a separable infinite-dimensional Hilbert space, then for any pair of infinite-rank projections p, q on H we have p I'V q (cf. Remark 4.1.5). Since Mn(B(H)) = B(H(n»), it follows that for any p E P[B(H)], we have 11 EB P I'V 11, so [P] = o. Hence, Ko(B(H)) = o. 7.1.1. Remark. It is easy to verify that for any positive integer n and any p, q E P[Mn(C)], we have p  q if and only if p I'V q. From this it follows easily that for any finite-dimensional C*-algebra A and p,q E P[A], p and q are stably equivalent if and only if they are equivalent. 7.2. The K-Theory of AF-Algebras We show in this section that the Ko-group of a unital AF-algebra A, endowed with some additional structure, is a complete isomorphism invariant of A (Elliott's theorem). The additional structure on Ko(A) is a naturally defined partial ordering (together with the "base point" [11]). The C* -algebraic concept that enables us to get the ordering is stable finiteness: A unital C* -algebra A is stably finite if, for every positive integer n and u E Mn(A) such that u*u = 1, we have uu* = 1. 7.2.1. Theorem. If A is a unital AF-algebra, then it is stably finite. Proof. First observe that if A = Mn(C) and u is an element of A having a left inverse, that is, there is an element v E A such that vu = 1, then by elementary linear algebra uv = 1. Since a finite-dimensional C*-algebra A is a direct sum of a finite number of such matrix algebras Mn(C), it follows in this case that an element of A that is left invertible is invertible. Now suppose that A is a unital AF-algebra. To prove the theorem, we have to show that if u E Mn(A) and u*u = 1, then uu* = 1. Since Mn(A) is a unital AF-algebra, it suffices to show the result in the case n = 1. Suppose then u E A and u*u = 1. There is a sequence (un) in A converging to u such that each Un is contained in a finite-dimensional C*-subalgebra An of A containing the unit of A. Since 1 = u*u = lim n -+ oo uun' we may suppose that 111 - u: Un II < 1 (by going to a subsequence if necessary) and therefore u:u n is invertible in An. It follows that Un is left invertible in An, and therefore (since An is finite-dimensional), Un is invertible in An. If V n is the inverse of u, then u = lim n -+ oo vnuun, so u = lim n -. oo v n . Hence, uu* = lim n -. oo vnu:, and since vnu = 1, therefore uu* = 1. 0 
222 7. K-Theory of C*-Algebras A partially ordered group is a pair (G, < ) consisting of an abelian group G and a partial order < on G such that if G+ = {x E G I 0 < x}, then G = G+ - G+, and if x < y, then x + z < y + z, for all x, y, z E G. If G is an abelian group and N is a subset such that N + N e N, G = N - N, and N n (-N) = {OJ, we call N a cone on G. If G is a partially ordered group, then G+ is a cone on G. If G is an abelian group and N is a cone on G, then we define a partial order < on G by setting x < y if y - x E N. Clearly, (G, < ) is a partially ordered group with G+ = N. We say < is the partial order induced by N. Partially ordered groups exist in great abundance. For instance, every subgroup of the additive group R is a partially ordered group with the order induced from R. An important example is given by Zk. This is a partially ordered group with the partial order induced by the cone N k . 7.2.2. Theorem. If A is a stably finite unital C*-algebra, then Ko(A)+ is a cone on Ko(A). Proof. The only thing not obvious is that x E Ko(A)+ n (-Ko(A)+) => x = o. Suppose that x = [P] = -[q] for some P, q E P[A]. Then [p EB q] = 0, so if r = P EB q, and r E Mn(A), then [In] = [In - r], and therefore, for some positive integer m, we have 1m EB (In - r) rv 1m EB In = 1m+n. Thus, there exists u E Mn+m(A) such that u*u = 1m+n and uu* = 1m EB (In - r). Since A is stably finite, uu* = 1m+n' so r = O. Hence, P and q are zero projections, and therefore, x = o. 0 If A is as in Theorem 7.2.2, then Ko(A) is a partially ordered group with partial order induced by J<o(A)+. 7.2.3. Lemma. Let A be a C*-algebra and PI,... ,Pn projections in P[A], and q a projection in A such that q rv PI EB. . .EBPn. Then there exist pairwise orthogonal projections qI, . . . , qn in A such that qi rv Pi (i = 1,. . . , n), and q = ql + . . . + q n . Proof. There exists a rectangular matrix W with entries in A such that w*w = q and ww* = PI EB . . . EB Pn. Write W as a block matrix, w = (::) , where each Wi is an mi x 1 matrix. We have q = w*w = El W:Wi, and ww. = (::) (w;,..., w) = * WIW I * W2 W I * WIW n * W2 W n * WnWI * WnW n 
7.2. The K-Theory of AF-Algebras 223 so Wi wi = Pi (i = 1, . . . , n) and Wi W; = 0 for i =I j (i, j = 1, . . . , n ). Set qi = WiWi. Each element qi is a self-adjoint element of A and q = q;, so by the functional calculus, qi is a projection. Also, ql + . . . + qn = q and qi  Pi. The qi are pairwise orthogonal, since qiqj = Wi(Wiwj)Wj = WiOwj=Oifii=j. 0 If 1 < i,j < n, define the element eij E Mn(C) to be the matrix with all its entries zero except for the (i,j) entry, which is 1. The matrices eij (i, j = 1,..., n) form a linear basis for M n ( C), called the canonical basis. We shall make frequent use of the following elementary facts: eijekl = 6jk e il and . eij = eji. Since Mn(C) = B(cn), and Mm(Mn(C)) = Mm(B(cn)) = B(c mn ), it is clear that every projection in Mm(Mn(C)) is unitarily equivalent to a diagonal matrix with only zeros and ones on the diagonal, so if P E P[Mn(C)], then [P] = k[ell] for some integer k. Now suppose that A is the C*-algebra Mn1(C) EB... EB Mnjc(C). We regard Mn,(C) as a C*-subalgebra of A in an obvious way. We denote by ej (i,j = 1,..., n,) the canonical basis of Mn,(C), and call the elements ej (1 = 1,..., k; i,j = 1,..., n,) the canonical basis of A. The homomorphism k T: Zk  J<o(A), (ml,..., mk) ....... L m,(el]' 1=1 is the canonical map from Zk to Ko(A). If cp: G l  G 2 is a group homomorphism between partially ordered groups Gland G 2 , we say cp is p03i tive if c.p( Gt) C Gt. If, in addition, c.p is bijective and cp-1 is also positive, we call cp an order i30morphi3m, and we say G l and G 2 are order i30morphic if such an order isomorphism exiss. 7.2.4. Theorem. If A = M n1 (C) EB... EBMnk(C), then the canonical map T: Zk  Ko(A) is an order isomorphism. Proof. If P is a projection in P[A], then P = (PI,. .. ,pk) = PI + . . . + Pk, where each PI is a projection in P[Mn,(C)]. Each PI is equivalent to a direct sum eil EB ... EB el in P[Mn,(C)], and therefore in P[A]. This shows that the elements [etl]'...' [el] generate the group Ko(A), so T is surjective, and this also shows that T((Zk)+) = Ko(A)+. Let 7r/: A  Mn,(C) be the projection *-homomorphism. Suppose that T(ml,.. ., mk) = 0; that is, E=l m,(eil] = o. Then, for each 1', we have k I I' I' I' · ( 7r 1 1 ).(L.JI=l m,(el l ]) = mi' [ell] = 0, so ell EB . . . EB ell (1m I' I-summands) IS equivalent to zero. Hence, mi' = o. Thus, T is injective. 0 
224 7. K-Theory of C*-Algebras 7.2.5. Corollary. If A is a non-zero finite-dimensional C*-algebra, then Ko(A) is a free abelian group with a basis Xl,. . . , Xk such that I<o(A)+ = NXI +... + NXk. Proof. This follows from Theorems 6.3.8 and 7.2.4. 0 If A and B are unital C*-algebras and r: Ko(A) -+ Ko(B), we say r is unital if r([1 1 ]) = [11]. 7.2.6. Theorem. Let A and B be non-zero finite-dimensional C*-algebras. (1) Suppose that r: I<o(A) -+ I<o(B) is a unital positive homomorphism. Then there is a unital *-homomorphism cp: A -+ B such that cp* = r. (2) If cp, 1/;: A -+ B are unital *-homomorpmsms, then cp* = 1/;* if and only if1/;= (Ad u)cp for some unitary U E B. Proof. Since there exists a *-isomorphism 7r from A to a direct sum of matrix algebras Mn(C), and this induces the isomorphism 7r* between the corresponding I<o-groups, we may suppose A = M n1 (C) EB. . . EB M nk (C) for some positive integers nl, . . . , nk. For I = 1,. . . , k and (i, j = 1, . . . , n,), let ej denote the canonical basis elements of A. Let e, be the unit of M n, (C) and let lA, 1B be the units of A, B, respectively. Recall that, over a finite- dimensional C*-algebra, stable equivalence and equivalence are the same ( Remark 7.1.1). We have r[e,] = [P,) for some projection PI E P[B], since r is positive. Hence, k [PI EB . . . EB Pk] = r(L:[e,)) = r[lA] = [lB], 1=1 so PI EB . . . EB Pk f'V 1 B. Therefore, by Lemma 7.2.3, there exists pairwise orthogon,al projections ql,... , q k E B such that ql + . . . + q k = 1 Band q, f'V PI for I = 1, . . . , k. Note that r[e,] = [q,]. Now r[ei1] = [Pil] for some projection pil in P[B], so nl(PI] = r(n,[el]) = r[e,] = [q,]. Hence, P 1 EB . . . EB P 1 (nl summands) f'V q,. Since q, is a projection in B, it follows from Lemma 7.2.3 that there exist .. th I . t . I I I. B h th t "",n, I paIrWIse or ogona proJec Ions ql1, Q22,. . . , qn"n, In suc a L..Jj=l qjj = q" and qj f'V pil for j = 1,. . . , n,. Note that r[eil] = [qj]. Since qj f'V q 1 for all j, there exist partial isometries U  E B such th t I - I ( I ) * d I - ( I ) * I S t I - ' ( I ) * £ .. - 1 a qjj - Uj Uj an qll - Uj Uj. e qij - ui Uj or Z,J - ,...,n" and note that this is consistent with our previous use of the symbols qj. 
7.2. The K-Theory of AF-Algebras 225 Elementary computations show that (qL)* = qi and q!jqn = bjmq!n, where i, j, m, n = 1,..., n,. It is straightforward to show from this that the unique linear map c.p: A  B, such that c.p( e1 j ) = q!j for 1 = 1,..., k and i, j = 1,..., n" is a unital *-homomorphism. Since the elements [el]'. · . , [el] generate the group Ko(A) (Theorem 7.2.4) and c.p*[e{l] = [ql] = T[e{l] for 1 = 1,. .., k, we have c.p* = T. This proves Condition (1). Suppose now that c.p, 'l/J are arbitrary unital *-homomorphisms from A to B. It is easily checked that if u is a unitary of B, then (Adu)* = ide Therefore, if 'l/J = (Ad u )c.p, then 'l/J* = c.p* Suppose conversely that 'l/J* = c.p*. Set p1 j = c.p( ej) and q!j = 'l/J( ej) for all 1 = 1,..., k and i,j = 1 ..., n,. Then [pj] = c.p*[e1 j ] = 'l/J*[ej] = [qL], so Pj f'V qL. Hence, there exist partial isometries u, E B such that Pl = uju, and qi1 = u'u;. Set k n, U = L L q:1 U'Pi. '=1 i=l A direct computation shows that u is a unitary in B and that UPj = q!jU for all 1, i, j. Hence, 'l/J( ej) = (Ad U )c.p( ej) for all 1, i, j, so 'l/J = (Ad u )c.p. Therefore, Condition (2) holds. 0 7.2.7. Lemma. Let p, q be projections in a C*-algebra A and suppose that there is an element u E A such that lip - u*ull and IIq - uu*11 are less than one and u = qup. Then p f'V q. Proof. The inequality lip - u*ull < 1 implies that u*u is invertible in the C*-algebra pAp and, similarly, the inequality Ilq - uu*1I < 1 implies that uu* is invertible in qAq. Let z be the inverse of lul in pAp, and put w = uz. Then w*w = zu*uz = zlul 2 z = p. Also uu*ww* = uu*z2u* = ulul 2 z 2 u* = uu*, so ww* = q by invertibility of uu* in qAq. Thus, p f'V q.O 7.2.8. Lemma. Suppose that A is a unital C*-algebra and (An)=l is an increasing sequence of C*-subalgebras of A containing the unit of A. Suppose also that U=lAn is dense in A. (1) If p E P[A], then there exists q E P[Ak] for some integer k such that [P]A = [q]A. (2) If p, q E P[Ak] for some k and [P]A = [q]A, then there exists an integer m > k such that [P]A m = [q]Am. Proof. For each integer 1, the sequence of C*-algebras (M,(An))=l is increasing, and the union U=lM,(An) is dense in M,(A). Moreover, each M,(An) contains the unit of M,(A). Thus, to prove the theorem, it suffices to show that if p is a projection in A, then it is equivalent to a projection in some Ak, and to show that if p, q are projections in some Ak equivalent in A, then they are equivalent in Am for some m > k. 
226 7. K-Theory of C*-Algebras Suppose first that p is a projection in A. Then there is a sequence (un) of elements in UnAn converging to p, and by replacing Un by Re(u n ) if necessary, we may suppose that the Un are self-adjoint. Since (u;)n also converges to p, there exists n such that lip - unll < 1/2 and lIu n - u1I < 1/4. Hence, by Lemma 6.2.2, there exists a projection q in the C*-algebra generated by Un such that Ilu n -qll < 1/2. Consequently, q belongs to some Ak, and since lip - qll < 1, there is a unitary U in A such that q = upu* (Lemma 6.2.1), and therefore, q I'V p. Now suppose that p, q are projections in some Ak equivalent in A. There exists u E A such that p = u*u and q = uu* and u = qup. Hence, there is a sequence (un) in UnAn converging to u, and we may suppose that Un = qunP for all n (replace Un by qunP if necessary). For sufficiently large n, we have lip - uunll < 1 and IIq - unu1I < 1, and we may choose such an n so that Un belongs to some Am with m > k. By Lemma 7.2.7, p and q are equivalent in Am. 0 The following elementary lemma will be used in the proof of Elliott's theorem. 7.2.9. Lemma. Let A, B, and C be unital stably finite C*-algebras. Suppose that A is finite-dimensional and that r: Ko(A)  Ko(C) and p: Ko(B)  Ko(C) are positive homomorphisms such that r(Ko(A)+) C p(Ko(B)+). Then there is a positive homomorphism r': Ko(A)  Ko(B) such that pr' = r. Proof. By Corollary 7.2.5, there is a basis Xl,. . . , Xk of Ko(A) as a free abelian group such that Ko(A)+ = NXI + .. . + NXk. The assumption that r(Ko(A)+) C p(I<o(B)+) implies that there are elements Y1,... ,Yk in Ko(B)+ such that r(xj) = p(Yj) for 1 < j < k. Let r' be the unique homomorphism from Ko(A) to Ko(B) such that r'(xj) = Yj (1 < j < k). Clearly, pr' = r. Moreover, r'(I<o(A)+) = NYI +... + NYk C Ko(B)+, so r' is positive. 0 7.2.10. Theorem (Elliott). Let A and B be unital AF-algebras and r a unital order isomorphism from Ko(A) to Ko(B). Then there is a *-isomorphism 'P from A to B such that 'P* = r. Proof. There exist increasing sequences (An)=l and (Bn)=l of finite- dimensional C*-subalgebras of A and B, respectively, such that (UnAn)- = A and (UnBn)- = B, and we may suppose that each An and Bn contains the unit of A and B, respectively. Denote by cpn: An  A and 'ljJn: Bn  B the inclusion *-homomorphisms. Let p be the inverse of r, so p is also a unital order isomorphism. It is immediate from Lemma 7.2.8 that Ko(A)+ is the union of the in- creasing sequence (cpZ(([<o(An))+))=I' so [<o(A) = UncpZ(Ko(An)). Simi- 
7.2. The K-Theory of AF-Algebras 227 larly, Ko(B)+ is the union of the increasing sequence (1P:((Ko(Bn))+))  1, and Ko(B) = U n 1/1:(K o (B n )). Set nl = 1. By Corollary 7.2.5, there is a basis Xl, . . . , Xk for the free abelian group Ko(Anl) such that Ko(Ant)+ = NXI + ... + NXk. Hence, rcp:l(Ko(Ant)+) = NTcpl(Xl) + ... + Nrcp:l(xk). Since Ko(B)+ is the increasing union of the sets 1/1;:( Ko (Bm)+) (m = 1,2, . . .), it follows that the elements rep:l(xl)'...' rcp:l(xk) belong to 1/1':(Ko(Bm)+) for some m > nl. Hence, rcp:l(Ko(A nl )+) C 1P':(Ko(Bm)+), so by Lemma 7.2.9 there is a positive homomorphism f: Ko(Anl) -+ Ko(Bm) such that the following diagram commutes: Ko(Anl) ep:t Ko(A)  !f !r I<o(Bm) 1P': [<0 (B).  Now f[lA] = [e]Bm say, and therefore [e]B = 1P,:[e]Bm = r'P: l [lA] = [lB]B. By Lemma 7.2.8, Condition (2), there exists ml > m such that [e]Bml = [lB]B ml . Let ,(f: Bm -+ Bml be the inclusion, and set r 1 = ,(f*f. Then r l is a unital positive homomorphism and the diagram I<0(A n1 ) epl Ko(A)  ! r l !r Ko(Bml) 1P;:l Ko(B)  commutes. By a similar argument applied to p1P':l, there exists an integer n > ml and a positive homomorphism p: 1<0 (Bml ) -+ Ko(An) such that the diagram I<o(Bml) 1/1 ;: 1 Ko(B)  !p !p Ko(An) ep: Ko(A)  commutes. We can write Xj = [Pj]A nl for projections PI,... ,Pk E P[Anl], and sim- ilarly, prl(xj) = [qj]An for projections ql,... ,qk E P[An]. We have [Pj]A = [qj]A, since 'P: pr l = P1P':l r l = prep:l = ep:l. Applying Lemma 7.2.8, Condition (2), there exists n2 > n such that P[A n2 ] contains PI, . . . , Pk and ql , · . . , q k, and [Pj]A n2 = [qj]A n2 (j = 1, . . . , k ) . (1) 
228 7. K-Theory of C*-Algebras Set pI = *P, where : An  A n2 is the inclusion. Then the following diagram commutes: I<o(Bml) !pl 1/J ': 1 ---+ Ko(B) !p 1<0(An2 ) c.p :2 ---+ Ko(A). Moreover, plrl = c.p1*, where c.pl: Anl  A n2 is the inclusion, since for each j, we have plrl(xj) = *pr1(xj) = *[qj]An = [qj]A n2 = [Pj]A n2 (by Eq. (1)) = c.p 1 * [P j ] An 1 = c.p 1 * ( X j ). Continuing in the above fashion, we inductively construct two sequences of integers such that nl < ml < n2 < m2 < ..., and positive homo- morphisms rk:J<o(A nk )  J<o(B mk ) and pk:I<o(Bmk)  Ko(A nk + l ) such that the diagrams J<o(A nk ) c.p :k J<o(A) ---+ ! r k !r (2) I<o(Bmle ) 1/J ': k I<o(B) ---+ and J<o(B mk ) 1/J ': k I<o(B) ---+ !pk !p (3) c.p:k+l Ko(Ank+l) ---+ 1<0 (A) commute, and pkrk = c.pk* and rk+l pk = 1/Jk*, where c.pk: Ank  Ank+l and 1/Jk: Bmk  Bmk+l are the inclusions. Note that r k and pk are necessarily unital (because r l is unital and pkrk = c.pk*, and rk+1 pk = 1/Jk* for all k, so by induction r k and pk are unital for all k). By Theorem 7.2.6, there are unital *-homomorphisms a l : Anl  Bml and (31: Bml  A n2 such that a = r 1 and (3! = pl. We have ((31 a l )* = plrl = c.pl*, so by Theorem 7.2.6 again, there is a unitary u in A n2 such that (Adu)(3la l = c.pl. Since (Adu)* = id, we may suppose that (3la l = c.pl (replacing (31 by (Ad U )(31 if necessary). Continuing in this fashion, we construct by induction unital *-homomorphisms a k : Anle  Bmk and (3k: Bmk  Ank+l such that for all k we have a = r k , (3 = pk, (3ka k = c.pk, and a k + l (3k = 1/Jk. If a E A nk , then ak(a) = ak+l(a), because ak(a) = 1/Jkak(a) = a k + l (3k a k ( a) = a k + 1 c.pk( a) = a k + l (a). Hence, if A' = UkAnk' we may well-define a map c.p: A'  B by set- ting c.p( a) = a k (a) if a E Anle. Since the maps a k are norm-decreasing 
7.3. Three Fundamental Results in K-Theory 229 *-homomorphisms, so is cpo Hence, it extends from the dense *-subalgebra A' of A to a *-homomorphism on A, again denoted by cpo In like manner we get a *-homomorphism "p:B -+ A such that if bE B rnle , then "p(b) = {3k(b). If a E A nle , then "pcp(a) = {3ka k (a) = CPk(a) = a. This shows that"pc.p = id, and similarly, one shows that c.p"p = ide Thus, cp is a *-isomorphism. Now suppose that P E P[A nle ]. Then T([P]A) = TcpIe([P]Anle) = "p:n le a([P]Anle) (the last equality follows from commutativity of Diagram (2) and the fact that T k = a). Hence, T([P]A) = 1/J:n Ie ([a k (p)]B mle ) = [c.p(p)]B = CP.([P]A). This shows that T = cp. on the sets c.pIe(Ko(Anle)+) for all k, and since these sets have union [{ 0 ( A ) +, and this generates K 0 ( A ), it follows that T = cp.. 0 7.2.11. Corollary. Two unital AF-algebras are *-isomorphic if and only if there is a unital order isomorphism between their Ko-groups. Proof. If cp is a *-isomorphism of unital AF-algebras, then cp. is a unital order isomorphism. This gives the forward implication. The reverse impli- cation is given by Theorem 7.2.10. 0 7.3. Three Fundamental Results in K-Theory The three results referred to in the title of this section are weak exact- ness, homotopy invariance, and continuity of the functor Ko, or more pre- cisely, of i<o. The latte: is the extension of Ko to the class of all C*-algebras. For technical reasons [{o is defined differently than Ko, but if an algebra A is unital, then Ko(A) and [{o(A) are isomorphic groups. Let A be a C*-algebra, which may be unital or non-unital. If T: A -+ C is the canonical *-homomorphism, we set i<o(A) = ker(T.), so Ko(A) is a subgroup of Ko(A.). If cp: A -+ B is a *-homomorphism of C*-algebras and (j;: A -+ iJ is the unique unital *-homomorphism extending cp, then <p.(I{o(A)) C I{o(B). Hence, we get a homomorphism cp.: I{o(A) -+ Ko(B) by restricting t{;.. It is straightforward to show that these constructions give a covariant functor A -+ Ko(A), cp  cp. from the category of all C*-algebras and *-homomorphisms to the category of all abelian groups and homomorphisms. Suppose now A is a unital C*-algebra, and e denotes its unit. If P E P[A], then T(p) = 0, so [P]A E J{o(A). Moreover, if P, q E P[A] and P  q relative to A, then for some integer n we have en EB P  en EB q relative to A, and therefore relative to A. Hence, [en]A: + [P]A: = [en EB p]A: = [en EB q]A: = [en]A: + [q]A' so [P]A = [q]A. Thus, we get a well-defined map I{o(A)+ -+ Ko(A), [P]A  [P]A' which is clearly a homomorphism, 
230 7. K-Theory of C*-Algebras apd which therefore extends uniquely to a homomorphism jA: KolA) -+ Ko(A). We call jA the natural homomorphism from Ko(A) to Ko(A). In the language of category theory, the following theorem asserts that j A implements a natural isomorphism between the two functors Ko and Ko on the category of all unital C*-algebras and unital *-homomorphisms. 7.3.1. Theorem. H A is a unital C *-alge bra, then the natural map jA: Ko(A) -+ Ko(A) is an isomorphism. Moreover, if <p: A -+ B is a unital *-homomorphism into a unital C*-algebra B, then the diagram Ko(A) <p. Ko(B) ---+ !jA !jB I{o(A) 'P. Ko(B) ---+ commutes. Proof. If e is the unit of A, then the map 'ljJ: A -+ A, a  eae, is a unital *-homomorphism. If j: Ko(A) -+ Ko(A) is the inclusion, then it is easily checked that 'l/J.j j A = id and j A 'l/J.j = ide Thus, j A is a bijection. Commutativity of the diagram is trivial. 0 We now need some elementary results on the unitary group of a unital C*-algebra. These results will be used in connection with weak exactness, which we shall be looking at presently. If A is a unital C*-algebra, we denote by Un (A) the group of unitaries of Mn(A). We let U(A) denote the connected component of the unit in Un (A). We shall also write U(A), UO(A) for U1(A), Uf(A), respectively. 7.3.2. Theorem. Let A be a unital C*-algebra. Then UO(A) is a normal subgroup of U(A). If u E A, then u E UO(A) if and only if there exist elements aI, . . . , an E Asa such that u = e ia1 . . . e ian . Proof. Let V be the set of all elements u of A which can be written in the form u = e ia1 . . . e ian for some n and some al,. . . , an E Asa. Any such element u belongs to UO(A), since the function [0, 1] -+ U(A), t  e ita1 . . . e itan , is a continuous path in U(A) from 1 to u. Obviously, V is a subgroup of U(A). By Theorem 2.1.12, V is a neighbourhood of 1 in U(A), and therefore (since V is a subgroup), V is a neighbourhood of all its points; that is, V is open in U(A). Since this implies that the cosets of V are also open, and 
7.3. Three Fundamental Results in K-Theory 231 since the complement of V in U(A) is a union of cosets, it follows that V is closed in U(A). Thus, V is a non-empty clopen subset of the connected set UO(A), and therefore V = UO(A). If u E U(A), then uUO(A)u- 1 is a connected set of unitaries containing 1, so uUO(A)u- 1 C UO(A). Hence, UO(A) is normal in U(A). 0 7.3.3. Corollary. Suppose that c.p is a unital surjective *-homomorpmsm from a unital C*-algebra A to a unital C*-algebra B. If v E UO (B), then there exists u E UO(A) such that v = c.p( u). Proof. If b E B sa, then b = c.p( a) for some a E A, and therefore b = c.p(Re(a)); that is, we may suppose that a E Asa. If v E UO(B), then v = e ib1 · . . e ibn for some bj E B sa. Hence, there exists al,..., an E Asa such that c.p(aj) = b j for all j, so u = e ia1 ... e ian E UO(A) and c.p(u) = v. 0 7.3.1. Remark. If u is a symmetry, that is, a self-adjoint unitary, in a C*-algebra A, then u E UO(A). This follows from the computation e i7ru / 2 = En even (i7r I)n In ! + u En odd (i7r /2)n /n! = cos( 7r /2) + iu sin( 7r /2) = iu. Thus, u = e l7ra , where a = (u - 1 )/2. 7.3.4. Lemma. Let p, q be equivalent projections in a unital C*-algebra A. Then there exists u E U(A) such that u(p EB O)u* = q EB 0 in M2(A). Proof. Let v be a partial isometry in A such that p = v*v and q = vv*, and set ( V u- - 1 - v*v 1 - vv* ) * . v Simple computations show that u is a unitary, and that if w=( ), then wand wu are symmetries. Hence, u is the product u = w( wu) of symmetries, and therefore u E U(A) (c/. Remark 7.3.1). That we have ( q 0 ) ( p 0 ) * o 0 =u 0 0 u again follows by direct computation. o If G .!: G'  G" is a sequence of homomorphisms of abelian groups, it is exact if im( T) = ker(p). The sequence o  G  G'  G"  0 (1) 
232 7. K - Theory of C * - Algebras is a 3hort exact 3equence of group3 if r is injective, p is surjective, and the sequence G  G' .t G" is exact. The short exact sequence (1) is said to 3plit if there is a homomorphism p': G"  G' such that pp' = ide In this case there is a unique homomorphism r': G'  G such that r'r = id and rr' + p' p = id a ,. Hence, the map G'  G ffi G", x  (r'(x), p(x)), is an isomorphism. The proofs of these observations are elementary exer- C1ses. A sequence of homomorphisms between groups G CPn G CPn+l G . .. n  n+l  n+2 . . . (where the sequence may be finite or extend infinitely in either direction) is said to be exact if im( CPn) == ker( CPn+l) for all relevant n. 7.3.2. Remark. Let A be a C*-algebra and let x E Ko(A). If 1 is the unit of A, then there exists an integer n and a projection p E P[A] such that x == [P] - [In]. To see this, observe that x == [r] - [q] for some projections r, q, both of which we may suppose to be elements of Mn(A) for some n. Then x = [r] + [In - q] - [In] = [p] - [In], where p = r E9 (In - q). The property of the functor J{o asserted in the following theorem is referred to as weak exactne33. 7.3.5. Theorem. Let J 'P OJABO be a short exact sequence of C*-algebras and *-homomorphisms. Then the sequence ,." J *"" CP*"" J<o(J)  J<o(A)  J<o(B) is exact. Proof. We may assume that J is an ideal in A and that j is the inclusion map. Since cpj = 0, we have cP*j* == 0* = 0, so im(j*) C ker( cP*). To show the reverse inclusion, let x E ker( CP*). Then we can write x = [P] - [In] for p a projection in some M m ( A) and for some integer n < m. (We may suppose the unit of J and that of A are the same and we shall use 1 to denote the unit for J, A, and B.) Now [cp(p)] = [In] in J<o(B), so there is an integer k such that 1k EBcp(p) f'V 1k EB In EBOm-n = 1k+n ffiOm-n in Mk+m(B). It follows from Lemma 7.3.4 that there exists v E Uk+2m(B) such that 1k+n EB Ok+2m-n = v(lk EB cp(p) EB Ok+m)V*. 
7.3. Three Fundamental Results in K- Theory 233 By Corollary 7.3.3, there exists u E Uk+2m(A) such that_<p(u) = v. Let r = u(lk EB p EB Ok+m)U*, so r is a projection in M 2k + 2m (A) equivalent to 1k EB P EB Ok+m. Since <p(r) = v(lk EB <p(p) E9 Ok+m)V. = 1k+n EB Ok+2m-n, it follows that r E M 2k + 2m ( j). It is easily checked that the element [r]-[lk+n] of Ko( j) actually lies in [(0 ( J). Finally, j.([r] - [lk+n]) = [lk EBp] - [lk+n] = [P] - [In] = x, so X E im(j.). Hence, ker(<p.) = im(j.). 0 7.3.6. Theorem. If Al and A 2 are C*-algebras, then Ko(AI E9 A 2 ) is isomorphic to i<o(A I ) EB i<0(A2). Proof. Let <Pi: Ai  Al EB A 2 and 7ri: Al EB A 2  Ai be the inclusion and projection *-homomorphisms, respectively. The sequence <PI 7r2 o  Al  Al EB A 2  A 2  0 is a short exact sequence of C*-algebras, so by Theorem 7.3.5 the sequence ]{o(A I )  ](o(A I EB A 2 )  I{0(A 2 ) is exact. Since 7ri<pz = id, and, therefore, 7ri.<Pi. = id (i = 1,2), the homomorphism <Pl. is injective, the homomorphism 7r2. is surjective, and o  I{o(AI)  I{o(AI E9 A 2 )  !(0(A 2 )  0 is a split short exact sequence. ](o(A I ) E9 i{0(A 2 ). Hence, ]<O(AI E9 A 2 ) is isomorphic to o If <p,,,p: A  Bare *-homomorphisms of C*-algebras A, B, we say <P and "p are homotopic, and write <P  "p, if for each t E [0, 1] there is a *-homomorphism CPt: A  B, where <Po = <P and <PI = "p, and for each a E A we have continuity of the map [0, 1]  B, t  'Pt(a). We then call ('Pt)OtI a homotopy from <P to "p. The relation 'P  "p is an equivalence relation on the *-homomorphisms from A to B. 7.3.1. Ezample. Let J be a closed ideal in a unital C*-algebra A, and let a E Asa. Set U = e ia , and let <p: J  J be the restriction to J of Ad u. Then <p is homotopic to the identity map id J of J. A homotopy (<pt)t from id J to <P is got by letting CPt be the restriction to J of Ad e ita for all t E [0,1]. 7.3.2. Ezample. If H is a Hilbert space and the map <p: I{(H)  K(H) is a *-isomorphism, then <P is homotopic to idK(H). This is immediate from Theorems 2.4.8 and 2.5.8 and from Example 7.3.1. 
234 7. K-Theory of C*-Algebras 7.3.7. Theorem. Ifcp,,,p:A  B are homotopic *-homomorphisms be- tween C*-algebras A, B, then cp. = "p.: Ko(A) -+ Ko(B). Proof. If (<pt)t is a homotopy from cp to "p, then it is easily checked that (t)t is a homotopy from  to ;p in which all t are unital. We may therefore suppose that A, B are unital and that there is a homotopy (<pt)t from cp to "p such that all CPt are unital, and show that cp. = "p.: Ko(A) -+ Ko(B). In this case, if P E Mn(A) is a projection and Pt = <Pt(p), then the map [0,1] -+ Mn(B), t t-+ Pt, is uniformly continuous, so there is a partition 0 = to < t 1 < .. . < t m = 1 of [0,1] such that IIptj - ptj+111 < 1 (0 < j < m). Hence, Ptj and Ptj+1 are unitarily equivalent by Lemma 6.2.1, so Ptj  Ptj +1' and therefore, cp(p)  "p(p). Consequently, 'P.([P]) = "p.([P]), and since P was an arbitrary element of P[A], we have cp. = "p.. 0 The following lemma says if A = An, then Mk(A) = Mk(An). It is of independent interest, but for us its importance is its application to proving "continuity" of the Ko-functor (Theorem 7.3.10). 7.3.8. Lemma. Let A be the direct limit of the sequence of C*-algebras (An,CPn) _ l' and let B be the direct limit of the corresponding sequence (Mk(An), CPn)=l' where k is a fixed integer. Denote by cpn: An  A and "pn: Mk(An)  B the natural maps for each integer n. Then there is a unique *-isomorphism 7r: B  Mk(A) such that for each n the diagram Mk(An) "pn B ---+ cpn !7r Mk(A) commutes. Proof. Since the diagram Mk(An) cpn Mk(An+l) ---+ cpn 1 <P n+ 1 Mk(A) commutes for each n, it follows from Theorem 6.1.2 that there is a unique *-homomorphism 7r: B  Mk(A) such that the diagram "pn ---+ B Mk(An) cpn ! 7r Mk(A) 
7.3. Three Fundamental Results in K-Theory 235 commutes for each n. As Uncpn(An) is dense in A, so Uncpn(Mk(An)) is dense in Mk(A), and it follows that 7r is surjective. To show that 7r is injective, it suffices to show that it is injective when restricted to the C*-subalgebras n(Mk(An)), since it is then iso- metric on these algebras, and therefore, by continuity of 7r and density of Unn(Mk(An)) in B, it follows that 7r is isometric on B. Suppose then that 7r(n(a)) = 0, where a E Mk(An). Let e > o. If bEAn and <pn(b) = 0, then there exists m > n such that lI<Pnm(b)1I < £ (cf. Re- mark 6.1.2). Applying this to the entries aij of the matrix a, since <pn( a) = 0, there exists m > n such that lI<Pnm( aij)1I < £ (1 < i, j < k). Hence, lI<Pnm(a)1I < Ej=ll1<Pnm(aij)1I < k2£ (cf. Remark 3.4.1). Consequently, lIn(a)1I = lIm<pnm(a)11 < IICPnm(a)1I < k 2 e. Letting £ -+ 0 we get lIn(a)1I = 0, so 7r is injective on n(Mk(An)) as required. This proves the theorem. 0 If (An' CPn)=l is a direct sequence of C*-algebras, we say it is unital if the algebras An and the *-homomorphisms CPn are unital. In this case the algebra A = limAn is unital, as are the natural *-homomorphisms  <pn: An -+ A. 7.3.9. Lemma. Let A be the direct limit of a unital sequence (An,CPn)=l of C*-algebras, and for each n let cpn: An -+ A be the natural map. (1) H p is a projection in A, then there is an integer n and a projection q E An such that p is unitarily equivalent to <pn( q) in A. (2) If n is given and p, q are projections in An such that cpn(p) rv <pn( q) in A, then there is an integer m > n such that CPnm(P) rv CPnm(q) in Am. Proof. Let p be a projection in A. Since A = (U=lcpn(An))- there is a sequence (cpnk(ak))l in Uncpn(An) converging to p. As p = p* we may suppose that each ak is self-adjoint (replace ak by Re( ak) if necessary). Since p = p2, the sequence (cpnk(ai))l also converges to p, and therefore (<pnk(ak - a))l converges to o. Hence, there exists an integer m and a self-adjoint element a E Am such that IIp-cpm(a)1I < 1/2 and IIcpm(a - a 2 )11 < 1/4. It follows that there exists n > m such that IICPmn(a - a 2 )11 < 1/4. Set b = <Pmn ( a). Then b is a self-adjoint element of An such that II b - b 2 11 < 1/4, and therefore, by Lemma 6.2.2, there is a projection q E An such that lib - qll < 1/2. Using the equality cpm(a) = <pn(b), we have lip - cpn(q)11 < lip - cpm(a)11 + IIcpn(b) _ <pn(q)11 < lip - cpm(a)1I + lib - qll < 1/2 + 1/2 = 1, so by Lemma 6.2.1 the projections p and <pn(q) are unitarily equivalent in A. This proves Condition (1). 
236 7. K-Theory of C*-Algebras Now suppose that n is a given integer, and that P, q are projections in An such that c.p n (p) f'V c.p n ( q) in A. Then there is a partial isometry u in A such that <pn(p) = u*u and c.pn(q) = uu*. Now u is the limit of a sequence (<pn/c (Vk) )1' where Vk E An/c and nk > n, and since u = uu*u = <pn(q)u = u<pn(p), we may suppose that Vk = <Pnn/c(q)Vk<PnnAl(P) (replace Vk by c.pnn/c(q)Vk<Pnn/c(P) if necessary). Clearly, <pn(p) = limk--+oo c.pn/c(vkVk) and <pn(q) = limk--+oo c.pn/c(Vk v k ). Hence, there exists an integer k > n and v E Ak such that v = c.pnk(q)Vc.pnk(P) and lI<pk(<Pnk(P) - v*v)11 < 1 and lI<pk( 'Pnk( q) - vv*)11 < 1. It follows that there is an integer m > k such that lIc.pkm(<Pnk(P) - v*v)1I < 1 and lIc.pkm(<Pnk(q) - vv*)11 < 1. Therefore, if w = c.pkm(V), then w E Am and we have lI<Pnm(P) - w*wll < 1 and lI<Pnm(q) - ww*1I < 1 and c.pnm(q)wc.pnm(p) = w. Hence, by Lemma 7.2.7, the projections c.pnm(P) and c.pnm(q) are equivalent in Am. This proves Condition (2). 0 The content of the following theorem is the continuity of Ko. It says that if A = An' where (An, c.pn)n is a unital sequence of C*-algebras, then Ko(A) = Ko(An). 7.3.10. Theorem. Let A be the direct limit of a unital sequence of C*-algebras (An, c.pn)=l' and let G be the direct limit of the corresponding sequence of abelian groups (I<o(An), c.pn*)=l. Denote by c.pn: An  A and Tn: KO(An)  G the natural maps. Then there is a unique isomorphism T: G  Ko(A) such that the diagram )  G [<0 (An --r c.p: ! T Ko(A) commutes for each integer n. Proof. For each integer n the diagram I<o(An)  Ko(An+l) <P:  ! c.p:+1 Ko(A) commutes, so there is a unique homomorphism T: G  Ko(A) such that for each n the diagram J{o(An) Tn G  c.p: !T J{o(A) 
7.3. Three Fundamental Results in K-Theory 237 commutes. For each integer k, let Bk be the direct limit of the direct sequence (Mk(An), <Pn)=l' and for each n, let c.pi: Mk(An) -+ Bk be the natural map. By Lemma 7.3.8, there is a unique *-isomorphism 7rk: Bk -+ Mk(A) such that for each n the diagram Mk(An) c.pk  Bk c.pn'\, ! 7rk Mk(A) commutes. We show first that r is surjective. Let p E P[A]. Then p E Mk(A) for some integer k. Hence, p = 7rk(q) for some projection q E Bk. By Lemma 7.3.9, Condition (1), there is a projection r E Mk(An) for some n such that q rv c.pk(r) in Bk. Hence, p rv <pn(r) in Mk(A), since 7rk<Pk = <pn. Consequently, [P]A = <p([r ]A n ) = rrn([r]A n ), since c.p = rr n . This shows that Ko(A)+ C im(r), and, since Ko(A)+ generates Ko(A), we have Ko(A) = im( r); that is, r is surjective. Now we show that r is injective. Suppose that x E ker(r). Since G = U=l im(rn), we can write x = rn([p]A n - [q]An) for some projections p, q in P[An]. We may suppose that p, q E Mk(An). Since r(x) = 0, we have [c.pn(p )]A = [c.pn( q )]A, as rr n = c.p. Thus, c.pn(p)  c.pn( q) relative to A, so there is an integer 1 such that 1,E8<pn(p) rv 1 , EB<pn(q); that is, c.pn(l,E8p) rv c.pn(l, E8 q) in the C*-algebra M'+k(A). Applying the *-isomorphism 7rlJk' we get <p;+k(l, E8 p) rv <p;+k(l, E8 q) in B'+k. Hence, by Lemma 7.3.9, Condition (2), there is an integer m > n such that <Pnm(l,E8P) rv <Pnm(l,E8q) in M,+k(Am). Therefore, <Pnm.([P]An) = c.pnm.([q]An)' and if we apply r m to both sides and observe that rm<pnm. = r n , we get rn([p]A n ) = rn([q]A n ). Hence, x = rn([p]A n - [q]An) = O. This shows that r is injective and completes the proof. 0 7.3.3. Eample. Let s: N \ {O}  N \ {O}, and define s!, Cs, and Ms as in Section 6.2. If S/: N \ {O}  N \ {O}, we saw in Theorem 6.2.3 that Ms and Ms' are *-isomorphic implies Cs = Cs'. We shall now use K-theory to prove that Ms and Ms' are *-isomorphic if Cs = Cs'. It is convenient to "normalise" the sequences s, S/: If oS is the sequence (1, Sl, S2, . . . ,), then Ms and Ms are *-isomorphic and c. = Cs. Thus, to prove our result we may confine ourselves to sequences s such that Sl = 1. Let An = Ms!(n)(C) and let the map <Pn: An -+ An+1 be the canonical *- homomorphism. Denote by Z( s) the additive group of rational numbers r which can be written in the form r = m/s!(n) for m E Z and n > 1. We make Z(s) into an ordered group by endowing it with the usual order from R. 
238 7. K-Theory of C*-Algebras If en is the matrix in An such that all its entries are zero except for the (1, I)-entry, which is 1, then Ko(An)+ = N[en]An' and Ko(An) = Z[en]An (Theorem 7.2.4). Let pn: Ko(An) -+ Z(s) be the unique positive homo- morphism such that pn([en]An) = l/s!(n). Since pn = pn+lCPn* for all n, there is a unique homomorphism p from Ko(An) to Z(s) such that pn = prn for all n, where the Tn: Ko(An) -+ Ko(An) are the nat- ural maps. A routine verification shows that p is an isomorphism. By Theorem 7.3.10, there is a unique isomorphism 0': Ko(An) -+ Ko(MIJ) such that O'T n = cP for all n, where cpn: An -+ MIJ are the natural maps. Hence, T = pa- 1 is an isomorphism from Ko(MIJ) to Z(s). An application of Lemmas 7.3.8 and 7.3.9 shows that Ko(MIJ)+ = U _ ICP:(Ko(An)+). Us- ing this, it is easily checked that T is in fact an order isomorphism. Also, r([ll]M.) = 1 (use the fact that pl([e 1 ]Al) = 1). Now suppose that s' is another function from N \ {o} to itself such that s = 1. By Corollary 7.2.11, the AF-algebras Ms and MIJI are *-isomorphic if and only if there is a unital order isomorphism from Ko(Ms) to Ko(MIJ/), and by our computations above this is equivalent to saying there is an order isomorphism T: Z(s) -+ Z(s') such that T(l) = 1. It is elementary that the latter condition is equivalent to Z( s) = Z( s'), which is in turn easily seen to be equivalent to Cs = Cs'. We want to extend our continuity result to 1<0. First some technical details on unitisations and direct limits are needed. 7.3.3. Remark. Let A be the direct limit of the sequence of C*-algebras (An, CPn)  1. Then A = An. We formulate this more precisely: Let B be the direct limit of the unital sequence (An, CPn)=I' and let cpn: An -+ A and "pn: An  B be the natural maps. Then there is a unique *-isomorphism cp: B -+ A such that for each n the diagram An 'ljJn B  -n !cp cp A commutes. We show only that cp is injective, as the rest is straightforward. It suffices to show for each n that cp is injective on "pn(A n ). Let a E An and ,,\ E C, and suppose that cp("pn(a+"\)) = O. Let c > o. Now (j;n = cp"pn, so cpn(a) = -"\, and therefore"\ = O. Hence, there exists m > n such that IICPnm(a)1I < e, and therefore lI"pn(a+"\)1I = lI"pmcpnm(a+"\)1I < IICPnm(a)1I < c. Letting c -+ 0, this gives "pn(a +,,\) = 0, so cp is injective on "pn(A n ) as required. 7.3.11. Lemma. Let A be the direct limit of the sequence of C*-algebras (An,CPn) _ I' and suppose G is the direct limit of the sequence of abelian 
7.3. Three Fundamental Results in K-Theory 239 groups (Ko(An),CPn.) - l. Denote by cpn:An -+- A and rn:Ko(An2 -+- G the natural maps. Then there is a unique isomorphism r: G -+- Ko(A) such that the diagram Ko{li n ) r n G ---+ -n !r <.p. Ko(A) commutes for all n. Proof. Uniqueness of r is clear. To see existence let B be the direct limit of the sequence (An, <Pn)=l' and let 'ljJn: An -+- B be !he natural *-homomorphism. There is a unique *-isomorphism cp: B -+ A such that for each n the diagram An 'ljJn B  -n !cp cp A commutes (cf. Remark 7.3.3). Also, by Theorem 7.3.10, there is a unique isomorphism p: G -+- J<o(B) such that the diagram K 0 (An) r n ---+ G 1/J: ! p Ko(B) commutes for each n. Set r = 'P.p. Then rr n = 'P.pr n = 'P.tP: = <P: for all n, and the lemma is proved. 0 7.3.12. Theorem. Let A be the direct limit of a sequence of C*-algebras (An, c.pn)=l' and G the direct limit of the corresponding sequence of abelian groups (Ko(An), CPn.) - l. Denote by cpn: An -+ A and rn: Ko{An) -+- G the natural maps. Then there is a unique isomorphism r: G ---+ Ko(A) such that the diagram K 0 (An) r n G ----+ cp: !r Ko(A) commutes for each integer n. Proof. The result follows from earlier results by a diagram chase, so we begin by setting up the diagrams: 
240 7. K-Theory of C*-Algebras Since the diagram [<o(An) CPn. ----+ Ko(An+1) 'P: ! cp:+1 Ko(A) commutes for each n, there is a unique homomorphism T: G  Ko(A) such that the diagram [<o(An) Tn G ----+ CP: !T (1) Ko(A) commutes for each n. Let G be the direct limit of the sequence (Ko(An), <.Pn.)=l' and for each n denote by fn: I<o(An)  G the natural map. By Lemma 7.3.11, there is a unique isomorphism f: G  Ko(A) such that the diagram I<o(An) fn G ----+ -n !f (2) 'P. Ko(A) commutes for all n. With these diagrams in place, we can now show injectivity of T. Sup- pose that x E ker( T ). For some integer n, and some y E K 0 ( An), we have x = Tn(y), so 0 = TTn(y) = cp:(y) by commutativity of Diagram (1). Since <'p:(y) = cp:(y), as y E I<o(An), and <.P: = ffn by commutativ- ity of Diagram (2), we have ffn(y) = 0, and because f is an isomorph- ism this implies that fn(y) = O. Hence, there exists m > n such that 'Pnm.(Y) = <.Pnm.(Y) = o. Therefore, x = Tn(y) = Tmcpnm.(y) = o. Thus, T is injective. Now we show surjectivity of T. Let Pn: An  C and p: A  C be the canonical maps. Suppose that z E I<o(A). Then z E I<o(A), so z = fey) for some y E G, because f is surjective. Since G = Un im(f n ), there exists an integer n, and x E I<o(An), such that y = fn(x). Hence, z = ffn(x) = <'p:(x) by commutativity of Diagram (2). However, x E Ko(An) = ker(Pn.), since z E I{o(A) = ker(p.) and pn.(x) = P.<.P:(x) = P.(z). Therefore, z = cp:(x) = TTn(X) by commutativity of Diagram (1), so z E im(T). 0 
7.4. Stability 241 7.4. Stability The most important result of this section asserts that if H is a separable infinite-dimensional Hilbert space, then Ko(A) = Ko(K(H) 0* A). This is referred to as 3tability of the functor Ko. It is a fundamental result, and will be used in the next section to prove Bott periodicity. Our line of attack is to show firs that Ko(A) = I<o(M 2 (A)), and then derive stability using continuity of Ko. If A is a C*-algebra, the map K: A -+ M 2 (A), a  ( ), is an injective *-homomorphism, which we shall call the inclu3ion of A in M 2 (A). In cases of ambiguity we shall write K A rather than K. 7.4.1. Remark. If A is a unital C*-algebra, then every element x E Ko(A) can be written x = [P]A - [q]A' where P, q belong to P[A]. We may even suppose that q = (lA)n for some integer n, where 1A is the unit of A. This follows from the natural isomorphism of Ko(A) with Ko(A) (Theorem 7.3.1), and the same trick we used in Remark 7.3.2. 7.4.1. Theorem. Suppose A is a unital C*-algebra and K: A -+ M2(A) is the inclusion of A in M 2 (A). Then the map K*: Ko(A) -+ K o (M 2 (A)) is an isomorphism. Proof. Let n be a positive integer. If ()" is a permutation of {I,. . . , n}, let U u be the unitary in Mn(A) defined by setting (Uu)ij = bU(i),j. If P is a projection in Mn(A), a routine verification shows that K(PII) K(PIn) ( P On ) On On = U u * U U , K(Pnl) K(pnn) where ()" is the permutation of {I, . . . , 2n} given by ()" = ( 1 2 3 .. . n n + 1 n + 2 . . . 2n ) 1 3 5 ... 2n - 1 2 4 . . . 2n . Hence, P rv K(p) relative to A. Suppose now that P is a projection in M m (M 2 (A)). Then P rv K(p) relative to M 2 (A), so [P]M 2 (Af = [K(p)]M2(Af = K*([P]A). This shows that K* is surjective. Now we show injectivity of K*. Suppose that x E ker(K*), so that for some positive integer n we have x = [P]A - [q]A for projections P, q in Mn(A). Now [K(p)]M 2 (Af = [K(ql]M 2 (Ar, so [K(p)]M2(A) = [K(q)]M 2 (A) (using the natural isomorphism of Ko(M2(A)) and Ko(M2(A))). Hence, K(p)  K(q) relative to M 2 (A). Therefore, K(p)  K(q) relative to A, so P rv K(p)  K( q) rv q again relative to A, so P  q relative to A. Hence, [P]A = [q]A' so x = o. Thus, K* is injective. 0 
242 7. K-Theory of C*-Algebras 7.4_.2. Remark. If A is a C*-algebra and i: A -+ A is the inclusion and p: A -+ C is the canonical map, then - z - - p. - o -+ Ko(A)  Ko(A)  Ko(C) -+ 0 is a short exact sequence. This is immediate from the natural isomorphism of Ko and Ko on unital C*-algebras (Theorem 7.3.1) and the fact that - J - P. o -+ Ko(A) ---+ I{o(A) ---+ Ko(C) -+ 0 is a short exact sequence (j denotes the inclusion). 7.4.2. Theorem. If A is a C*-algebra and K: A -+ M 2 (A) is the inclusion, then K.: Ko(A) -+ K o (M 2 (A)) is an isomorphism. Proof. If i: A -+ A is the inclusion and p: A -+ C is the canonical map, then we have a commutative diagram A z A P C ---+ ---+ !K A !K A !K C M2(A) z - P M 2 (C) ---+ M 2 (A) ---+ and therefore the corresponding diagram on the Ko-level commutes: Ko(A) z. I{o(A) p. Ko(C) ---+ ---+ !K1 !K1 !K Ko(M2(A)) J Ko(M2(A)) P.) K o (M 2 (C)). ---+ To avoid ambiguity in the argument to follow, we are denoting the map i.: Ko(M2(A)) -+ !(0(M 2 (A)) by j. The top row in the second diagram is a short exact sequence (cf. Remark 7.4.2), so, in particular, i. is injective. Since K1 is an isomorphism by Theorem 7.4.1, and since jK1 = K1i., it is clear that K1 is injective. If we assume that the map j is injective, we can show that K1 is surjective: If x E K o (M 2 (A)), then j(x) = K1(y) for some y E Ko(A), by surjectivity of K A . Since KP*(Y) = P*K(Y) = P.j(x) = 0 and K is injective by Theorem 7.4.1, we have P.(y) = 0, and therefore y E i.(Ko(A)). Thus, y = i.(z) for some z E Ko(A), and j(x) = K1i.(z) = jK1(z). Because j is assumed injective, we have, therefore, x = K1( z). Hence, K1 is surjective. 
7.4. Stability 243 Thus, to prove the theorem we need only show that j is injective. Let B = M 2 (A) + C1 2 , and let k: M 2 (A) -+ Band 1fJ: B -+ M 2 (A) be the inclusion *-homomorphisms. The diagram M2(A) z M 2 (A)  k i1fJ B commutes, so the diagram K o (M 2 (A)) J Ko(M2(A))  k* i * i{o(B) commutes. By Remark 7.4.2, the map k* is injective (identify B with M 2 (A)). Thus, to show that j is injective, it suffices to show that 1fJ* is injective. We need another map: Denote by r the unique unital *-homomorphism from B to C1 2 having kernel M 2 (A). If x E ker(*), then by Remark 7.4.1 there exist an integer k and projections p,q E Mk(B) such that x = [P]jj - [q]jj. We may even suppose q is of the form q = (1 2 )n ED Or. Hence, r(q) = q. Now p  q relative to M 2 (A), and to prove the theorem we need only show this implies that p  q relative to B. There is an integer m such that (1 2 )m ED p  (1 2 )m ED q in M m + k (M 2 (A)). Thus, replacing p and q by (1 2 )m ED p and (1 2 )m ED q if necessary, it suffices to show that if p, q are projections in Mk(B) such that p  q in Mk(M 2 (A)) and r(q) = q, then p  q in Mk(B). Since p is a *-homomorphism, we have p(p)  p(q) in M k (M 2 (C)). Now p(p) = r(p) and p(q) = r(q) belong to the subalgebra M k (C1 2 ), and to show that they are equivalent in this sub algebra we have only to show they have the same rank (c/. Example 7.1.1). But for projections in Mk(C12) the rank is the same as the trace, and the normalised trace on Mk(C1 2 ) is just the restriction of the normalised trace on M k (M 2 (C)), so r(p), r(q) have the same trace in Mk(C1 2 ) because they are equivalent in M k (M 2 (C)). Hence, there is a partial isometry w E M k ( C 1 2 ) such that r (p) = w* wand r(q) = ww*. Since p  q in M k (M 2 (A)), there exists v E Mk(M 2 (A)) such that p = v*v and q = vv*. Set u = wp( v)*v. Then u = Ul + U2, where Ul = wp(v)*(v - p(v)) and U2 = wp(v)*p(v). Because M k (M 2 (A)) is an ideal in Mk(M 2 (A)), and v-p(v) E M k (M 2 (A)), we have Ul E Mk(M 2 (A)), and therefore U1 E Mk(B). As U2 = wp(v*v) = wp(p) = wr(p), so U2 E Mk(C1 2 ), and therefore, U2 E Mk(B). Hence, U E Mk(B). Finally, u*u = v*p(v)w*wp(v)*v = v*p(v)p(p)p(v)*v = v*p(vv*vv*)v = v*qv = p and 
244 7. K-Theory of C*-Algebras UU* = wp(v)*vv*p(v)w* = wp(v)*qp(v)w* = wp(v*vv*v)w* = wr(p)w* = r(q) = q, so p rv q in Mk(B) as required. 0 If A is a C* -algebra, if H is a Hilbert space, and if p is a rank-one projection in K(H), then the map 'P: A  I«H) 0* A, a  pO a, is a *-homomorphism. Write P* = 'P*: 1<0 (A)  Ko(I«H) 0* A). If q is another rank-one projection in I« H), then there is a unitary u E B( H) such that q = upu*. Hence, by Theorem 2.5.8, there is a self-adjoint operator v E B(H) such that u = e iv . For t E [0,1], set Ut = e itv , so Ut is a unitary in B(H). If 'Pt = 'lfJt'P: A  I«H)0*A, where'lfJt = Ad ut0*idA: I«H)0*A  K(H) * A, then it is easy to check that ('Pt)t is a homotopy. Hence, p* = 'Po* = 'PI* = q*. The homomorphism P* is called the canonical map from I{o(A) to Ko(K(H) 0* A). ",The following is a key result of K-theory, and is referred to as stability of 1<0: 7.4.3. Tl1eorem. If A is a C*-algebra and H is a separable infinite- dimensional Hilbert space the canonical map from I<o(A) to I<0(I«H)0*A) is an isomorphism. Proof. Set K = I«H). Let (en)=l be an orthonormal basis of H, and write eij for the operator in B(H) given by eij(x) = (x, ej}ei. Set Pn = Ej=l ejj and note that the map n 'lfJn: Mn(A)  PnI<Pn ° A, (aij)  L eij ° aij, i,j=l is a *- isomorphism. Set Bn = M 2 n-l(A) for n > 1. The map 7r n : Bn  1< 0* A, a  'lfJ2n-1 (a), is an isometric *-homomorphism, and 7r n = 7r n + I K n , where Kn: Bn  Bn+l is the inclusion; that is, Kn(a) = ( ). Let B be the direct limit of the sequence (Bn, Kn)=l and for each n let K n : Bn  B be the natural map. Then there is a unique *-homomorphism 7r: B  K 0* A such that for all n we have 7r n = 7rK n . Since I{ = (U=IP2n-l I<P2 n - 1 )-, 
7.5. Bott Periodicity 245 we have K @. A = (U=IP2n-1I{P2n-l @ A)- = (U=I7rn(Bn))-' and therefore 7r is surjective. Moreover, 7r is isometric on each subalgebra ",n(B n ), and therefore, by density of U  l",n(Bn) in B, it follows that 7r is isometric on B. Thus, 7r is a *-isomorphism and therefore 7r. is an isomorphism. If G is the direct limit of the sequence of abelian groups (Ko(Bn), "'n.), and if for each n we denote by r n the natural map from Ko(Bn) to G, then by Theorem 7.3.12, there is a unique isomorphism r: G  Ko(B) such that "' = rr n for all n. It follows from Theorem 7.4.2 that each map "'n.: Bn  Bn+l is an isomorphism, and therefore each map r n is an isomorphism. Hence, "': = rr n is an isomorphism, and therefore 7r: = 7r ."': is an isomorphism. Since 7r 1 is the map from A = MI(A) to K @. A given by 7r 1 (a) = 1/;1(a) = ell @ a, the map 7r is the canonical map from Ko(A) to Ko(I{ @. A). This proves the theorem. 0 7.4.3. Remark. If cP, 1/;: A  Bare *-homomorphisms of C*-algebras, we say they are orthogonal if cp( a)1/;( a') = 0 (a, a' E A). In this case cp + 1/; is a *-homomorphism. If p E P[A] then cp(p) and 1/;(p) are orthogonal projections, so (cp + 1/;).[P]A = [cp(p) + 1/;(p)] A = [CP(P)]A + [1/;(p)]A = (cp. + 1/;.)[P]A. If A is unital, the elements of [{o(A) are of the form [P]A - [q]A (p, q E P[A]), so clearly, (cp + 1/;). = cp. + 1/;*: [{o (A)  [{o (B). 7.5. Bott Periodicity We shall find it convenient to adopt the following notation: If n is a locally compact Hausdorff space and A is a C*-algebra, we set An = Co(n, A). If cp: A  B is a *-homomorphism between C*-algebras A and B, the map <P: An  Bn, f....... cp 0 f, is a *-homomorphism. 7.5.1. Theorem. Let A, B be C*-algebras and let n be a locally compact Hausdorff space. If(cpt)t is a homotopy of*-homomorphisms from A to B, then (<pt)t is a homotopy of *-homomorphisms from An to Bn. Proof. It is easily checked that if C is the set of all 9 E An such that 1/;g: [0, 1]  Bn, t....... <Pt(g), 
246 7. K-Theory of C*-Algebras is continuous, then C is a C*-subalgebra of Aft If f E Co(n) and a E A, then 'Pt(fa) = f'Pt(a), so the map "pg is continuous in the case that 9 is of the form 9 = fa. Since the elements of An of this form have closed linear span An by Lemma 6.4.16, it follows that C = An and therefore ('Pt)t is a homotopy. 0 If A is a C*-algebra, then the C*-algebra C(A) = {f E A[O, 1] I f(l) = O} is a closed ideal in A[O, 1], called the cone of A. A C*-algebra A is said to be contractible if the identity map id: A -+ A is homotopic to the zero map. In this case I{o(A) = 0, by Theorem 7.3.7. 7.5.2. Theorem. If A is a C*-algebra, then its cone C(A) is contractible. Proof. Let f E C(A) and for t E [0,1] define 'i't(f) E C(A) by setting 'i't(f)( s) = f( 1 - t + st) (0 < s < 1). It is easily checked that the map 'Pt: C(A) -+ C(A), f  'Pt(f), is a *-homomorphism. Since the map h: [0,1]2 -+ A, (s, t)  f(l - t + st), is continuous, and therefore uniformly continuous, it follows that if c > 0 there exists some 6 > 0 such that max(ls - s'l, It - t'l) < 6 => IIh(s, t) - h(s', t')11 < c/2. Thus, if It - t'l < 6, then II'Pt(f) - 'Ptl(f)II < c. Hence, the map [0, 1] -+ C(A), t  'i't(f), is continuous for all f E C(A), and therefore ('Pt)t is a homotopy on C(A). Since 'i'o = 0 and 'PI = id, the zero and identity maps on C(A) are homo- topic; that is, C(A) is contractible. 0 7.5.3. Theorem. If A is a contractible C*-algebra and n is a locally compact Hausdorff space, then An is also contractible. Proof. If ('i't)t is a homotopy from the zero map of A to the identity map of A, then (<Pt)T is a homotopy (by Theorem 7.5.1) from the zero map of An to the identity map of An. Thus, An is contractible. 0 If A is a C*-algebra, we define its 3u3pen3ion to be the C*-algebra SeA) = {f E A[O, 1] I f(O) = f(l) = O}. 
7.5. Bott Periodicity 247 Thus, S(A) is a closed ideal in C(A). If c.p: A -+ B is a *-homomorphism of C*-algebras, then r{; maps S(A) into S(B), so if we denote its restriction by S( c.p): S(A) -+ S(B), then S( c.p) is a *-homomorphism, and it is clear from Theorem 7.5.1 that if (c.pt)t is a homotopy of *-homomorphisms from A to B, then (S(c.pt))t is a homotopy of *-homomorphisms from S(A) to S(B). Hence, if A is contractible, so is its suspension S(A). If A is an arbitrary C*-algebra, set K 1 (A) = Ko(S(A)). If c.p: A -+ B is a *-homomorphism of C*-algebras, denote by cp. the homomorphism (S(<,o)).: K 1 (A) -+ K 1 (B). If there is a possibility of ambiguity, we shall write Ko(c.p) and K 1 (c.p) for the homomorphisms c.p.:Ko(A) -+ Ko(B) and c.p.: k 1 ( A) -+ k 1 ( B), respecti vely. It is straightforward to verify that A  K 1 (A), c.p ....-..+ c.p., gives a covariant functor from the category of C*-algebras to the category of abelian groups. 7.5.4. Lemma. Let A be a unital C*-algebra and a < (3 real numbers. If p: [a,,8] -+ A is a continuous path of projections, then there is a continuous path of unitaries u: [a,,8] -+ A such that p(t) = u(t)p(a)u.(t) for all t in [a,,8] and u( a) = 1. Proof. Suppose first that IIp(t) - p(a)1I < 1 for all t. Set v ( t) = 1 - p( a) - p( t) + 2p( t )p( a ) . It follows from Lemma 6.2.1 that vet) is invertible and u(t) = v(t)lv(t)I- 1 is a unitary such that p(t) = u(t)p(a)u.(t) for all t, and u(a) = 1. It is easily checked that the function t  u(t) is continuous. We reduce the general case to the preceding case by using the uniform continuity of p. There ex;ists a partition ao < al < ... < an of [a,,8] such that IIp(t) - p(s)1I < 1 for all t,s E [ai, ai+l]. Therefore, there is a contin- uous path of unitaries Ui: [ai, ai+l] -+ A such that p(t) = Ui(t)p(ai)ui(t) for t E [ai, ai+l], and Ui( ai) = 1. Set u = Uo on [ao, al], and if i > 0 and t E [ai,ai+l], set u(t) = Ui(t)ui-l(ai)... uO(al). This gives a well-defined continuous path t  u(t) ofunitaries such that pet) = u(t)p(a)u.(t) for all t E [a,,8] and u ( a) = 1. 0 7.5.1. Remark. Let f2 be a locally compact Hausdorff space and A a C*-algebra. If f E M n (C o (f2, A)), define 9 = c.p(f) E C o (f2, Mn(A)) by setting g( w) = (fij (w )) for all w E f2. It is a straightforward exercise to show that this gives a *-isomorphism f  c.p(f) from M n (C o (f2, A)) onto C o (S1, Mn(A)). We shall call this the canonical *-isomorphism. 
248 7. K-Theory of C*-Algebras 7.5.5. Lemma. Let A be a unital C*-algebra such that Un(A) = U(A) for all n > 1. Then [(I (A) = o. Proof. The inclusion SeA)  0([0, 1], A) has a unique extension to an injective unital *-homomorphism <p: S(A)- -+ C([O, 1], A). The inflation <p: Mn(S(A)-) -+ M n ( C([O, 1], A)) is therefore an injective *-homomorphism also. Composing this together with the canonical *-isomorphism 8 from Mn(C([O, 1], A)) to C([O, 1], Mn(A)), we get a *-isomorphism 'l/;n: Mn(S(At) -+ S1 n (A), where nn(A) = 8epM n (S(A)j = {j E C([O, 1], Mn(A)) I j(O) = 1(1) E Mn(C)}. Moreover, if r: S(A)-  C is the canonical map, and c denotes the *- homomorphism r!n(A)  Mn(C), j  j(O), then the diagram Mn(S(An  On(A) r'\. ! £ Mn(C) commutes. Now suppose that x E KI(A) = j{o(S(A)). Then we may write x = [P] - [In] for some projection p in M 2n (S(A)-). Since r(p) "-I r(ln EB On) in M 2n (C), there is a unitary u' in M 2n (C) such that u'r(p)u'* = In EB On. Hence, replacing p by u'pu'* if necessary, we may suppose that r(p) = In EB On. The element q = 'l/;2n(P) E r!2n(A) is a continuous path of projections in M 2n (A), so by Lemma 7.5.4 there is a continuous path u: [0,1] -+ M 2n (A) of unitaries such that q(t) = u(t)q(O)u*(t) for all t, and u(O) = 1 2n . Now q(l) = q(O) = In ED On, so u(l )(In Ef) On)u*(l) = In ED On. This implies that u(l) can be written in the form UI ED U2, where UI, U2 E Un(A). Since Un(A) = U(A) by hypothesis, there exist continuous paths VI, V2: [0,1] -+ Un(A) of unitaries such that Vi(O) = In and vi(l) = ui for i = 1,2. Set w(t) = U(t)(VI(t) ED V2(t)). Then w: [0, 1]  M 2n (A), t...... w(t), is a continuous path of unitaries such that w(O) = w(l) = 1 2n , and therefore w is a unitary in r!2n(A). Moreover, q(t) = w(t)q(O)w*(t) for all t, (1) since Vt(t) ED V2(t) commutes with q(O) = In ED On. There exists a unitary w' E M 2n (S(A)-) such that 'l/;2n( w') = w. Since "p2n (p) = "p2n ( w' )'l/;2n (1 n ED On )"p2n ( w')* by Eq. (1) we have p = w'(ln ED On)w'*. Hence, [P] = [In], and x - [P] - [In] = o. Thus, [<0 (S(A) ) = o. 0 
7.5. Bott Periodicity 249 7.5.6. Theorem. If A is a unital AF-algebra or a van Neumann algebra, then Kl(A) = o. Proof. It suffices to show that Un(A) = U(A) for all n, by Lemma 7.5.5. If A is a unital AF-algebra, so is Mn(A), and likewise if A is a von Neumann algebra, Mn(A) is one also. Thus, the theorem is proved if we show that U(A) = UO(A). The von Neumann algebra case is given by Remark 4.2.2. That U(A) = UO(A) if A is finite-dimensional is given by Theorem 2.1.12. If A is a unital AF-algebra and u E U(A), then there is a finite-dimensional C*-subalgebra B of A containing the unit of A, and a unitary v in B such that lIu - vII < 1. Hence, 111- uv* II < 1, so uv* E UO(A) by Theorem 2.1.12 again. Since v E U(B) = UO(B) C UO(A), this implies that u E UO(A). 0 It is sometimes the case that the only effective means of showing that a C*-algebra is not an AF-algebra is to show that its !(l-group is non-zero. It will be useful in a number of contexts to have an alternative way of looking at the suspension, and for this reason we introduce a new algebra: We denote by S the closed ideal of C(T) consisting of all functions 1 such that 1(1) = O. 7.5.7. Theorem. If A is a C*-algebra and, is the function [ 0 1 ]  T t t-+ e i2 11"t , , , then there is a unique *-isomorphism ,A from A Q9* S to SeA) such that ,A(a 0 I) = (I o,)a for all 1 E S and a E A. Proof. Uniqueness is obvious, so we show only existence. The map A x S ...... S(A), (a, f) 1-+ (/ 0 ,)a, is bilinear, so it induces a unique linear map ,A: A 0 S  S(A) such that ,A(a Q9 I) = (/ 0 ,)a for all f E S and a E A. It is easy to check that ,A is an injective *-homomorphism and therefore the function p: A Q9 S --+ R + , C 1-+ 111' A (c) II , is a C*-nonn. Since S is abelian it is nuclear (Theorem 6.4.15), so p is the unique C*-nonn on A 0 S. Hence"A is isometric, and can therefore be extended to an isometric *-homomorphism from A 0* S to S(A) which we shall also denote by ,A. To show that ,A is surjective, it suffices to show that S(A) is the closed linear span of the elements of the form fa where a E A and f E S(C). This follows from the easily verified fact that SeA) ...... A(O, 1), 9 1-+ gr, is a *-isomorphism (where gr denotes the restriction of 9 to (0,1)) and from an application of Lemma 6.4.16 to A(O, 1). 0 
250 7. K-Theory of C*-Algebras 7.5.2. Remark. The *-isomorphism in Theorem 7.5.7 is natural in the sense that if c.p: A -+ B is a *-homomorphism of C*-algebras, then the diagram c.p 0. id A 0. S --+ B 0. S !,A !,B S(A) S(c.p) --+ S(B) commutes. 7.5.8. Theorem. If J c.p O-+J-+A-+B-+O is a short exact sequence of C*-algebras, so is o -+ S(l) Sjj) SeA) ) S(B) -+ O. Proof. By Theorem 6.5.2, O J S j . id A S c.p . id B S 0 -+ 0. --+ . --+ 0.-+ is a short exact sequence. A straightforward diagram chase using this and Remark 7.5.2 then shows that o -+ S(l) Sjj) SeA)  S(B) -+ 0 is a short exact sequence. As with Ko, the functor K 1 is weak exact: o 7.5.9. Theorem. Suppose that O-+JLAB-+O is a short exact sequence of C*-algebras. Then the sequence K 1 (J)  K1(A) c.p.) K1(B) is exact. 
7.5. Bott Periodicity 251 Proof. This is immediate from Theorems 7.5.8 and 7.3.5. o We are going to derive a connection between [(1 to [(0. This requires some convoluted constructions. If <PI: Al -+ B and <P2: A 2 -+ Bare *-homomorphisms of C*-algebras, then C = {( a I , a2) E A I EI1 A 2 I <P 1 ( a I) = <P2 ( a2 ) } is a C*-subalgebra of Al Ef) A 2 , called the pullback of Al and A 2 along <PI and <P2. If <p: A -+ B is a *-homomorphism of C*-algebras, we denote by Zcp the pullback of A and B[O, 1] along the *-homomorphisms <P and B[O,l] -+ B, 1  J(O). The surjective *-homomorphism Zcp -+ A, (a, J)  a, . is called the projection of Zcp onto A, and the injective *-homomorphism A -+ Zcp, a  (a, <pea)) is the inclusion of A in Zcp. The *-homomorphism e:Zcp-+B, (a,/)/(l), is the canonical map from Zcp to B. The kernel of c is denoted by Ccp and is called the mapping cone of <p. Explicitly, Ccp = {(a, J) E A EI1 B[O, 1] 1/(0) = <p(a), 1(1) = OJ. The surjective *-homomorphism Ccp -+ A, (a, J)  a, is the projection from C t,p onto A. 7.5.10. Lemma. Let <p: A -+ B be a *-homomorphism of C*-algebras and suppose that if: Zt,p -+ A is the projection, and i: A -+ Zt,p is the inclusion, *-homomorphism. Then ifi = idA, and iif is homotopic to id zcp . Proof. It is obvious that ifi = ide For J E B[O,l] and t E [0,1], define Jt E B[O,l] by Jt(s) = I(ts). If (a,J) belongs to Zcp, so does (a, Jt), and the map <pt: Zt,p -+ Zt,p, (a, f)  (a, Jt), is a *-homomorphism. It is easily checked that (<pt)t is a homotopy from iif to idz . 0 cp 
252 7. K-Theory of C*-Algebras 7.5.11. Lemma. Let c.p: A -+ B be a *-homomorphism of C*-algebras and let 7r: Cr.p -+ A be the projection of Cr.p onto A. Then the sequence ,.., 7r *"" c.p*,.., Ko(CV') ---+ Ko(A) ---+ Ko(B) is exact. Proof. Let i: Zr.p -+ A be the projection, i: A -+ Zr.p the inclusion, and e: Zr.p -+ B the canonical map. If j: Cr.p -+ Zr.p is the inclusion, then ij = 7r and ei = c.p, so i*j* = 7r* and £*i* = c.p.. By Lemma 7.5.10, 1fi = id A and ii  id zlp , so 7r*i* = id and i.7r. = id by Theorem 7.3.7. Hence, £. = c.p*7r.. Since J £ o -+ Cr.p -+ Zr.p -+ B -+ 0 is a short exact sequence of C*-algebras, it follows from Theorem 7.3.5 that the sequence ,.., J*"" £,.., Ko( Cr.p) ---+ I<o(Zr.p) --:4 Ko(B) is exact. Hence, 7r*(im(j.)) = i*(ker(£*)). Since 7r.(im(j.)) = im(7r.) and 7r*(ker(£*)) = ker(c.p*), therefore im(7r*) = ker(c.p*). 0 If c.p: A -+ B is a *-homomorphism of C*-algebras, we call the injective *- homomorphism k: S(B) -+ Cr.p, f  (0, f), the inclusion of S(B) in Cr.p. It is easily checked that k 7r o -+ S(B) -+ Cr.p -+ A -+ 0 is a short exact sequence of C*-algebras, where 7r: Cr.p -+ A is the projection. Given a short exact sequence of C*-algebras J c.p o -+ J -+ A -+ B -+ 0, we define an injective *-homomorphism J: J -+ Cr.p, a  (j( a), 0), which we call the inclusion of J in Cr.p. Note that j = 7rJ. The surjective *- homomorphism cp: Cr.p -+ C(B), (a, f)  f, is the projection of Cr.p onto C(B). It is readily verified that '" '" J c.p o -+ J  Cr.p ---+ C(B) -+ 0 is a short exact sequence. 
7.5. Bott Periodicity 253 7.5.12. Lemma. Suppose that J c.p OJABO is a short exact sequence of C*-algebras, and let j: J  Ccp be the inclusion. Then Ko(C j ) = 0, and the map ].: Ko(J)  Ko(Ccp) is an isomorphism. Proof. Set D = {f E G(Gcp) I f(O) E j(J)}. Then D is a C*-subalgebra of G(Gcp), and the map t/J: G j  D, (a, f)  f, is a *-isomorphism. Hence, 'ljJ.: Ko(G j )  Ko(D) is an isomorphism. Let the map <p: Ccp  G(B) be the projection, and denote by <p' the *- homomorphism D  S(G(B)), f  cp 0 f. Suppose that g E S(C(B)). Then the map [0, 1]  B, t  g(t)(O), belongs to S(B), so by the surjectivity of S( c.p) (Theorem 7.5.8), there exists a map f' E S(A) such that c.p(f'(t)) = g(t)(O) for all t E [0,1]. The map f: [0, 1]  Gcp, t  (f'(t),g(t)), is continuous, and indeed fED and cp'(f) = cp 0 f = g. Therefore, cp' is surjective. The map j':G(J)-+D, fjof, is an injective *-homomorphism, and clearly cp' j' = 0, so im(J') C ker( <p'). Suppose f is an arbitrary element of ker( cp'). Then cp 0 f = 0, so for all t E [0,1] we have f(t) = (f'(t),O) E A EB G(B) for some f'(t) E A. But cp(f'(t)) = 0, so f'(t) = j(h(t)) for some element h(t) E J. The map h: [0, 1]  J, t  h(t), is continuous, because f is continuous, and moreover, h(l) = 0, so h E G(J). Also, j'(h)(t) = (] 0 h)(t) = (j(h(t)), 0) = f(t), so j'(h) = f. Hence, im(j') = ker( <.p'). It follows from what we have just shown that  A' o -+ G(J) L D  S(G(B))  0 
254 7. K-Theory of C*-Algebras is a short exact sequence. Therefore,  ", Ko(C(J))  Ko(D) J!..!. Ko(S(C(B))) is exact by Theorem 7.3.5. But Ko(C(J)) = 0 and Ko(S(C(B))) = 0, since C(J) and S(C(B)) are contractible. Hence, 0 = im(J) = ker(<p) = Ko(D), so Ko(C j ) = o. Using Theorem 7.3.5 again, since " " J 'P o --+ J  C", --+ C(B) -+ 0 is a short exact sequence, it follows that the sequence " " "" J. "" 'P.- Ko(J) --+ Ko(C",)  Ko(C(B)) = 0 is exact. Hence, im(J.) = ker( <P.) = Ko(C",), so J. is surjective. Finally, suppose that 7r: C j  J is the projection. Then " ,., 7r. - J.- o = Ko(C j ) --+ Ko(J) --+ Ko(C",) is exact by Lemma 7.5.11, so 0 = im(7r.) = ker(J.). Thus, J. is injective, and therefore we have shown that it is an isomorphism. 0 If A is a C*-algebra and f E S(A), then the map KA(f): [0, 1]  A, t  f(l - t), also belongs to S(A). It is easily checked that KA: S(A)  S(A), f  KA(f), is a *-isomorphism such that K = ide 7.5.13. Lemma. Suppose that O-+JLAB-+O is a short exact sequence of C*-algebras, so k 7r o  S(B)  C",  A -+ 0 is also a short exact sequence, where k and 7r are the inclusion and pro- jection maps. Let k: S(B)  C 1r and k': S(A)  C 1r be the inclusion maps associated to this second short exact sequence. Then kS( 'P) and k' KA are homotopic *-homomorphisms from SeA) to C 1r . 
7.5. Bott Periodicity 255 Proof. For f E S(A) and t E [0,1], define ft E C(A) by setting ft(s) = f(l - t + st). The map CPt: SeA) -+ C 1r , f.-. ((ft(O),cp 0 (KA(!))l-t),!t), is a *-homomorphism, cpo = kS( cP )KA, and CPI = k'. It is straightforward to verify that (cpt)t is a homotopy, and therefore kS( cp )KA  k', from which it follows that kS ( cp)  k' K A, using the fact that K = ide 0 Suppose that J c.p O-+J-+A-+B-+O (2) is a short exact sequence of C*-algebras and ]: J -+ Ccp and k: S(B) -+ Ccp are the inclusion maps. We denote the composition (].)-lk. by a. Thus, a is a homomorphism from j{l(B) to j{o(J), called the connecting homo- morphism (relative to the short exact sequence (2)). 7.5.14. Theorem. If J cP O-+J-+A-+B-+O is a short exact sequence of C*-algebras, then the sequence K1(A)  K1(B)  Ko(J)  Ko(A) is exact. Proof. Let k, 7r, k', and k be as in Lemma 7.5.13. Since 7r j = j, we have 7r .]. = j., and since k 7r o -+ S(B) -+ Ccp -+ A -+ 0 is a short exact sequence, the sequence K1(B)  Ko(Ccp)  Ko(A) is exact by Theorem 7.3.5. Hence, im(a) = j;l(ker(7r.)) = ker(j.). By Lemma 7.5.13 kS(cp)  k'KA, so k.S(cp). = k(KA).. If fl is the connecting homomorphism for the short exact sequence k 7r o -+ S(B) -+ Ccp -+ A -+ 0, then - a' - k.- [{leA) ---+ Kl(B) ---+ Ko(Ccp) is exact, by the first part of this proof. Since a' = k;lk = S(cp).(KA)., we have ker( a) = ker( k.) = im( a') = im( S( c.p).). 0 
256 7. K-Theory of C*-Algebras 7.5.3. Remark. If O-+JLAB-+O is a short exact sequence of C*-algebras, we say that it 3plits if there exists a *-homomorphism 'ljJ: B -+ A such that 'P'ljJ = idB. In this case the sequence - J. - 'P.- o -+ Ko(J) -+ Ko(A) -+ Ko(B) -+ 0 is a split short exact sequence also. That the equality im(j.) = ker( 'P.) holds is a consequence of Theorem 7.3.5, and that cp. is surjective follows from the fact that CP. 'ljJ. = ide Injectivity of j. follows from the exactness of the sequence - CP. - a - j.- I<l(A) -+ I{l(B) -+ /{o(J) -+ Ko(A) (Theorem 7.5.14) and the fact that j{l(cp) is surjective (/<1 (cp)K1 ('ljJ) = id). Recall that A denotes the Toeplitz algebra (see Section 3.5), and that this algebra is generated by the unilateral shift u. We shall make frequent use of the universal property of A: If v is an isometry in a unital C*-algebra B, then there is a unique *-homomorphism cp: A -+ B such that cp( u) = v (Theorem 3.5.18). We denote I{(H2) by K (H 2 is the Hardy space). The unique *-homomorphism r: A -+ C such that T( u) = 1 is the canonical map from A to C. 7.5.15. Theorem (Cuntz). Let D be a unital C*-algebra and T: A -t C be the canonical map. Then the map - (r 0. id). - /(o(A 0. D) -+ I{o(C 0. D) is an isomorphism. Proof. If B is a C*-algebra, we shall simplify the notation throughout this proof by writing B' for B 0. D. If cp: B -+ C is a *-homomorphism of C* -algebras, we shall write cp' for cp 0. idD. Let j: C -+ A be the unique unital *-homomorphism. Then rj = id, so r'j' = id, and therefore Tj = ide Thus, we have only to show that j T = ide Let e = 1 - uu. where u is the unilateral shift on the standard ortho- normal basis of H 2 , and denote by c the *-homomorphism A -+ K 0. A, a  e ° a. There is a *-isomorphism ,: (K 0. A) 0. D -+ K 0. (A 0. D) such that ,((b ° a) ° d) = b ° (a ° d) (cf. Exercise 6.9), and (,c').: Ko(A 0. D) -t Ko(K 0. (A 0. D)) is clearly the canonical isomorphism. Hence, c is an 
7.5. Bott Periodicity 257 isomorphism. Since Al = {all a E A} is a C*-subalgebra, and K*A is a closed ideal, of A * A, the set B = A  1 + K * A is a C*-subalgebra of A. A. Let 7r: B -+ B /(K * A) be the quotient *-homomorphism, and let (J denote the *- homomorphism A-+B, aal. Denote by C the pullback of B and A along the *-homomorphisms 7r and 7r(J. The maps i: K * A -+ C, b  (b,O), and p:C -+ A, (b,a)  a, are, respectively, injective and surjective *-homomorphisms, and i p o -+ K * A -+ C -+ A -+ 0 is a short exact sequence which splits (the *-homomorphism k:A-+C, a(a01,a), is a right inverse for p). Since A is nuclear (cf. Example 6.5.1), it follows from Theorem 6.5.2 that ., , o -+ (K * A)'  c' L A' -+ 0 is a short exact sequence, and this also splits, since p' k' = ide Hence, by Remark 7.5.3, i is injective. Set 1/J = ie. Then 1/J = ic is injective, since i and e are. To show that j T = id, it therefore suffices to show that Itl.' J ., T' = Itl.' 0/* * * 0/*. Let Zo = u 2 u*2 0 1 + eu*  u + ue  u* + e  e and Zl = u 2 u*2  1 + eu*  1 + ue 0 1. Then Zo and Zl are symmetries in B, and if for each t E [0,1] we set Ut = -i exp( i7r(l - t )zo /2) exp( i7rtz 1 /2), then the map [0, 1] -+ B, t  Ut, is a continuous path of unitaries such that Uo = Zo and Ul = Zl. Since 7r(Zt) = 1 for t = 0,1 we have 7r(Ut) = 1 for all t E [0,1]. 
258 7. K-Theory of C*-Algebras Let CPt: A -+ C be the unique *-homomorphism such that CPt( u) is the isometry (Ut( U ° 1), u) in C. Then (cpt)t is a homotopy, and therefore (cp)t is a homotopy. If v = u 2 u., then v is an isometry in the C*-algebra (1 - e)A(l - e), and (vOl, u) is an isometry in the unital C*-algebra 8((1 - e)A(l - e)) EB A. Hence, there is a unique *-homomorphism cP from A into this C*-algebra such that cp( u) = (v 01, u), and since (v 01, u) E C, we may suppose that cp is a *-homomorphism from A to C. Since 1/;( u) = (e ° u, 0), we have cp( u )1/;( u) = 1/;( u )cp( u) = cp( u )1/;( u.) = 1/;( u .)cp( u) = 0, from which it follows that cp and 1/; are orthogonal. There- fore, cp' and 1/;' are orthogonal, and cp and "p jT are orthogonal. Moreover, CPo = cp + 1/; and cp 1 = cp + "p j T, so cp = cp' +"p' and cp = cp' + 1/;' j' T'. Hence, by Theorem 7.3.7 and Remark 7.4.3, cp + 1/; = (cp' + 1/;'). = CP. = CP. = (cp' + "p'j'T'). = cp + 1/;jT, and therefore"p = "pjT. This proves the theorem. 0 Since B is *-isomorphic to B 0. C for any C*-algebra B, the preceding theorem implies in particular that Ko(A) = J{o(C), so i<o(A) = Z. If T: A -+ C is the canonical map, we denote its kernel by Ao. 7.5.16. Corollary. If D is an arbitrary C*-algebra, then I{i(A o 0.D) = 0 for i = 0, 1. Proof. First suppose that D is unital. If j: Ao -+ A is the inclusion map, then J T o -+ Ao -+ A -+ C -+ 0 is a short exact sequence that clearly splits, and therefore the corresponding sequence . , , o -+ A  A'  C' -+ 0 is a split short exact sequence (Theorem 6.5.2), where we retain the notation B' = B 0. D, cp' = cp 0. id, used in the proof of Theorem 7.5.15. It follows from Remark 7.5.3 that ., , o -+ ko(A)  i{o(A')  Ko(C') -+ 0 is a short exact sequence, and since T is an isomorphism by Theorem 7.5.15 therefore Ko(Ao 0. D) = o. Now suppose that D is not necessarily unital. The split canonical short exact sequence O-+D-+D-+C-+O gives rise to a split short exact sequence o -+ D 0. Ao -+ iJ 0. Ao -+ C 0. Ao -+ 0, 
7.5. Bott Periodicity 259 and therefore o  Ko(D 0. Ao)  Ko(D 0. Ao)  Ko(C 0. Ao)  0 is a short exact sequence. Since (by Exercise 6.10) B 0. Ao is *-isomorphic to Ao @. B for any C*-algebra B, and Ko(Ao @. D) = 0 by the first part of this proof, hence Ko(Ao 0* D) = O. Finally, by Theorem 7.5.7, S(Ao@*D) is *-isomorphic to (Ao0*D)*S and this algebra in turn is *-isomorphic to A o @* (D@. S) (by Exercise 6.9), so K 1 (A o O. D) is isomorphic to Ko(Ao O. (D O. S)) = o. 0 Let 7r be the unique *-homomorphism from A to C(T) such that 7r( u) = z, where u is the unilateral shift and z is the inclusion function of T in C. Clearly, 7r(Ao) C S. We use the same symbol 7r to denote the *-homomorphism from Ao to S got by restriction. It is easy to check that K C Ao. If j: K  Ao denotes the inclusion *-homomorphism, then J 7r o  K  Ao -+ S -+ 0 is a short exact sequence, by Theorem 3.5.11. Hence, by Theorem 6.5.2, for each C*-algebra D the sequence j * id 7r 0* id o -+ K @* D --+ Ao 0* D --+ S 0* D  0 is a short exact sequence. By Exercise 6.10, there is a unique *-isomorphism B: S . D -+ D @. S such that B(!  d) = d  f for all f E Sand d E D. Denote by 7rD the *-homomorphism from Ao0*D to S(D) got by composing 7r 0* id D , (} and ID (cf. Theorem 7.5.7), so 7rD = ID(}( 7r 0* id D ). Then j 0* id 7rD o -+ K @* D --+ Ao 0. D --+ S(D) -+ 0 is short exact, so we get a homomorphism 8:K 1 (S(D)) -+ Ko(K 0. D) (the connecting homomorphism). We denote by (3D the homomorphism from K 1 (S(D)) to Ko(D) got by composing the maps 8 -1 - - e. - [<}(S(D)) -+ [<o(K 0. D) --+ Ko(D), where e.: Ko(D) -+ Ko(K 0* D) is the canonical isomorphism. 7.5.17. Theorem (Bott Periodicity). For each C*-algebra D, the map (3D: K 1 (S(D)) -+ Ko(D) is an isomorphism. 
260 7. K-Theory of C*-Algebras Proof. Applying Theorem 7.5.14 to the short exact sequence j 0. id 7r D o -+ K 0. D --+ Ao 0. D --+ S(D) -+ 0 gives exactness of the sequence - 7r D. - a - (j 0. id). - K 1 (A o 0. D) --+ KI(S(D)) -+ Ko(K 0. D) --+ Ko(Ao 0. D), and Ki(A o 0. D) = 0 for i = 0,1, by Corollary 7.5.16. Hence, a and therefore {3D are isomorphisms. 0 Let 0 J J A  B 0 -+ --+ -+ !1' !o !{j . , cp' 0 J' J A' B' 0 -+ --+ --+ -+ be a commutative diagram of *-homomorphisms of C*-algebras and suppose the top and bottom rows are short exact sequences with corresponding connecting homomorphisms a and a'. Then the diagram KI(B) !a j{o( J)  K1(B') !a' 1'.) Ko(J') commutes. The proof is a simple diagram chase. If <p: D -+ D' is a *-homomorphism between C*-algebras, then the diagram o -+ K0.D j 0. id Ao0.D 7rD S(D) 0 --+ --+ -+ ! id 0.cp ! id 0.<p ! S( <p) j 0. id 7f' K 0. D' Ao 0. D' D S(D') 0 --+ --+ -+ o -+ commutes (this uses Remark 7.5.2), and the top and bottom rows are short exact sequences. If a and a' are the corresponding connecting homo- morphisms, then by the preceding observation the diagram KI(S(D)) KlCP)) KI(S(D')) !a !a' Ko(K 0. D) (id 0.cp). Ko(K 0. D') --+ 
7.5. Bott Periodicity 261 commutes. Also, Ko(D) !e cp.  Ko(D') ! e' - (id 0.cp). - Ko(K 0. D)  Ko(K 0. D') commutes, where e and e' are the canonical isomorphisms. Since (3D = e- 1 a and f3 D' = (e') -18', the diagram K1(S(D» Kl<P» K1(S(D'» ! (3 D ! (3 D' ( 1 ) i<o(D) cp.  Ko(D') commutes. 7.5.18. Theorem. If OJLABO is a short exact sequence of C*-algebras, the following sequence is exact: i<o( J) r8  I<o(A)  Ko(B) !8 Kl(B)  Kl(A)  K 1 (J). The map a: Ko( B)  [(1 (J) is defined to be the composition of the homomorphisms a-I a , - fJB'" - Ko(B) -+ I<I(S(B))  K 1 (J), where a' is the connecting homomorphism from the short exact sequence o  S(J) SJj) S(A)  S(B)  O. Proof. By Theorems 7.3.5, 7.5.9, and 7.5.14, we have only to show exact- ness at Ko(B) and K 1 (J). Now ker(a) = f3B(ker(8')) = f3B(im(K 1 (S(cp)))), by exactness of K1(S(A» Klcp» K1(S(B» L K1(J). But f3B(im(K 1 (S(cp)))) = im(Ko(cp)), by commutativity of Diagram (1) preceding this theorem. Thus, we have exactness at Ko(B). Since im( a) = im( a') = ker( i< 1 (j ) ), by exactness of K1(S(B» L k1(J) k) K1(A), the theorem is proved. o 
262 7. K-Theory of C*-Algebras 7.5.4. Remark. If O-+JLAB-+O is a split short exact sequence of C*-algebras, then it follows easily from Theorem 7.5.18 that o -+ KI(J)  KI(A) 'P.) KI(B) -+ 0 is a split short exact sequence of groups. We have already seen the corre- sponding result for Ko in Remark 7.5.3. 7.5.1. Eample. If c denotes the *-homomorphism C(T) -+ C, f  f(l), then o -+ S L C(T)  C -+ 0 is a split short exact sequence of C* -algebras, where j is the inclusion. Hence, if A is an arbitrary C*-algebra, o -+ S 0. A j id C(T) 0. A e id C 0. A -+ 0 is a split short exact sequence, by Theorem 6.5.2. It follows from Re- marks 7.5.3 and 7.5.4 that for i = 0, 1 the sequence - (j Q9. id). - (e 0. id). - o -+ Ki(S Q9. A) --+ I<i(C(T) 0. A) --+ Ki(C Q9. A) -+ 0 is split short exact, and therefore, if "" denotes the relation "is isomorphic to," we have Ki(C(T) 0. A)  Ki(S Q9. A) E9 i<i(C 0. A)  J(i(S(A)) E9 Ki(A) "" I(I-i(A) E9 Ki(A)  J{o(A) E9 J<I(A) (the third  is given by Theorem 7.5.17). In particular, Ki(C(T))  Ko(C) EB KI(C)  z. 7. Exercises 1. Let Al and A 2 be C*-algebras. Show that (a) KI(AI EB A 2 ) = Kl(A I ) EB I(I(A 2 ); (b) K I (A 1 ) = J{I(AI). Hence, extend Theorem 7.5.6 by showing that KI(A I ) = 0 if Al is an AF-algebra (unital or not). 
7. Exercises 263 2. Let C be the Calkin algebra on an infinite-dimensional separable Hilbert space H. Calculate Ki(K(H)) and Ki(C) for i = 0,1. 3. Show that if 'P,,,p: A -+ B are homotopic *-homomorphisms between C*-algebras, then [(1 ('P) = [(1 ("p). 4. Show that the functor K I is "continuous," that is, K I (An) - liIIl- KI(An). More precisely, suppose that A is the direct limit of a sequence of C*-algebras (An, CPn2 - 1' and that G is the direct limit of the cor- responding sequence (K I (An), CPn.)  1. Suppose that cpn: An -+ A and Tn: KI(A n ) -+ G are the natural maps. Show that there is a unique iso- morphism T: G -+ K}(A) such that the diagram [(I (An) Tn G ----+ c.p: !T KI(A) commutes for all n. 5. Let H be a separable infinite-dimensional Hilbert space and e a rank-one projection in K(H). If A is a C*-algebra, the *-homomorphism c.p: A -+ [{(H) Q9. A, a  e Q9 a, is independent of the choice of e up to homotopy. The map e. = c.p.: K I (A) -+ [(1 (K(H) Q9. A) is called the canonical map. Show that it is an isomorphism. 6. Let A be the Toeplitz algebra, and let T: A -+ C be the canonical *-homomorphism. Extend Theorem 7.5.15 by showing that for an arbitrary C*-algebra D, the map - (T Q9. id). - Ki(A Q9. D) ---+ Ki(C 0. D) is an isomorphism for i = 0, 1 (use Corollary 7.5.16). 7. Let T be a tracial state on a C*-algebra A. Define Tn: Mn(A) -+ C by setting Tn(a) = L:I T(aii) if a = (aij). Show that Tn/n is a tracial state on the C*-algebra Mn(A). 
264 7. K-Theory of C*-Algebras 8. Let T be a tracial state on a C*-algebra A, and denote by the same symbol the unique (tracial) state on A extending A, and the tracial posi- tive linear functionals Tn on the C*-algebras Mn(A) obtained from T as in Exercise 7.7. If p, q E P[A] and p  q, show that T(p) = T(q). Show there is a well-defined homomorphism T*: Ko(A) -+ C such that T*([P] - [q]) = T(p) - T(q) for all x = [P] - [q] E Ko(A). 9. Let s: N \ {OJ -+ N \ {OJ and let M$ be the corresponding UHF algebra as in Sections 6.2 and 7.3. Denote by T the unique tracial state of Ma. Calculate the range T*(Ko(M a )). 7. Addenda Let A be a unital C*-algebra and let 'Pn: Un(A) -+ U n + 1 (A) be the group homomorphism defined by setting 4?n(a)=( ). Since CPn(U(A)) is contained in U+l(A), we obtain an induced homo- morphism 1/ln: Un(A)/U(A) -+ Un+1(A)/U+1(A). Denote by KI(A) the direct limit of the direct sequence of groups (Un(A)/U(A), 1/ln)=l. There is an isomorphism from K 1 (A) onto K1(A). An order unit for a partially ordered group G is an element u E G+ such that for each x E G+ there exists n > 1 such that x < nu. A countable partially ordered group G is a dimen3ion group if (a) whenever nx E G+ and n > 1, then x E G+; . (b) whenever Xl, X2, Yl, Y2 E G satisfy Xi < Yj for all i and j, there exists z E G such that Xi < z < Yj for all i andj. Condition (b) is called the Rie3z interpolation property. If A is a unital AF-algebra, then Ko(A) is a dimension group with an order unit, and in the reverse direction, if G is a dimension group with an order unit, there is a unital AF-algebra A such that Ko(A) is order isomorphic to G (Effros-Handelman-Shen). We say that C*-algebras A and Bare 3tably i30morphic if there is a separable infinite-dimensional Hilbert space H such that K(H) 0* A and K(H) 0* Bare *-isomorphic. If A is an AF-algebra, one can make Ko(A) into a partially ordered group in all cases, whether A is unital or not. AF-algebras A, B are stably isomorphic if and only if 1<0 (A) and Ko(B) are order isomorphic as partially ordered groups (Elliott). There is an analogue of Theorem 7.2.10 (also due to Elliott) for non-unital AF-algebras, where Ko(A) has to be endowed with additional structure (a "scale"). References: [Eff], [Goo], [Bla]. 
7. Addenda 265 Let n > 1 and let On be the C*-algebra generated by n isometries VI,. . . , V n such that VI vi + . · . + V n V: = 1. This algebra is called the Cuntz algebra. It is simple, and independent of the choice of VI,. . . , V n up to *-isomorphism. The Ko-group has torsion in the case of these algebras: Ko(On) = Z/(n - 1). One also has KI(On) = O. Let (} be an irrational number in [0,1], and let As denote the ir- rational rotation algebra (Exercise 3.8). The K-theory is given by Ko (As) = KI(As) = Z2. There is a unique tracial state T on As, and T.CKo(As)) = Z + Z(}. References: [Bla], [Cun]. 
Appendix In this appendix all vector spaces are relative to the field K, which may be R or C. Let r be a non-empty family of semi norms on a vector space X. The smallest topology on X for which the operations of addition and scalar multiplication and all of the seminorms in r are continuous is called the topology generated by r. If £ > 0 and PI, . . . , pn E r, let U ( €; PI , . . . , Pn) = {x E X I P j ( x) < € (j = 1, . . . , n ) } . These sets form a basic system of neighbourhoods of 0 in X. If (XA);\EA is a net in X, then (XA)AEA converges to a point x of X if and only if limA p( x,\ - x) = 0 for all pEr. A pair (X, r) consisiting of a vector space X and a topology r generated by a family of semi norms on X is called a locally convex (topological vector) space. There are other more geometric characterisations of these spaces but these will be irrelevant to us. It is easy to check that if r generates the topology on a locally convex space X, then X is Hausdorff if and only if r is separating, that is, for each non-zero element x E X there is a seminorm pEr such that p( x) > o. A.l. Theorem. Let r be a family of seminorms generating the topology on a locally convex space X, and suppose that r is a linear functional on X. The following conditions are equivalent: (1) r is continuous. (2) There are a finite number of elements PI, . . . , Pn of r and there is a positive number M such that Ir(x)1 < M mx pj(x) 1 $J :5 n (x EX). Proof. The implication (2) => (1) is clear. To show the reverse impli- cation, suppose that Condition (1) holds; that is, r is continuous. Then 267 
268 Appendix the set S = {x E X I Ir(x)1 < I} is a O-neighbourhood in X, so there is a positive number £ and there are semi norms PI, . . . , Pn E r such that U(e;PI,... ,Pn) C S. If P = maxljnPj, then p(x) < c => Ir(x)1 < 1. Set M = 2/e. If p(x) > 0, we have p( M;(x» ) = £/2 < e, so Ir( M;(x» )1 < 1. Consequently, Ir(x)1 < Mp(x) for all x E X. 0 Let X be a normed vector space. The weak* topology on X* is gener- ated by the family of semi norms Px: r t-+ Ir(x)1 (x E X). For x E X, denote by x the linear functional on X* defined by setting x( r) = r( x ). A.2. Theorem. Let X be a normed vector space. Then a linear functional (}: X* -+ K is weak* continuous if and only if (} = x for some x EX. Proof. We show the forward implication only, because the reverse impli- cation is trivial. Suppose that (J is weak* continuous. By Theorem A.1, there exist vectors XI,..., X n E X and there is a number M E R+ such that 18( r)1 < M max{ Ir(xi )1, . . . , Ir( xn)l} (r E X*). Hence, (} equals 0 on ker( XI) n. . . n ker( x n ). It follows by elementary linear algebra that (} = Alxl +.. .+Anxn for some scalars AI'...' An. Thus, (} = X, where x = Al xI + . . . + Anxn. 0 A.3. Theorem. Let r be a linear functional on a locally convex space X. Then r is continuous if and only ifker(r) is closed in X. Proof. The forward implication is obvious. Suppose then that Y = ker( r) is closed. To show that r is continuous, we may suppose it is non-zero. If C = r- l (l), there is a vector Xo in C, so C = Xo + Y. Hence, C is closed, and X \ C is an open neighbourhood of o. Let r be a generating family of seminorms for the topology of X. Then there is a positive ntUl1ber £ and there are elements PI,... ,Pn E r such that U = U(£;PI,... ,Pn) C X \ c. For any point x E U, choose, E K such that 1,1 = 1 and ,r(x) = Ir(x)l. Then ,x E U and I" = r(,x) E R+ \ {I}. If I" > 1, then ,x/I" E U, so r( ,x / 1") =I 1; that is, 1 =11, a contradiction. This argument proves that I" < 1. Hence, x E U => Ir(x)1 < 1, and therefore, for all x EX, Ir(x)1 < (2/£)InPj(x). _1_ By Theorem A.1, r is therefore continuous. o If Y is a vector subspace of a locally convex space X of codimension 1 in X, then there is a linear functional r on X with kernel Y, so if Y is closed in X, then r is continuous. Let x,y E X. The set [x,y] = {tx + (1- t)y 10 < t < I} is connected in X, since it is the continuous image of the unit interval [0,1] in R. If x, y are linearly independent, then [x, y] C X \ {O}. 
Appendix 269 If X is not one-dimensional, X \ {OJ is connected. For if x, y E X \ {OJ are linearly independent, then they are connected by a continuous path in X \ {OJ, as we have just observed, and if they are linearly dependent, there is a point z E X \ {O} such that z, x are linearly independent, and likewise for z, y, so we have a continuous path in X \ {OJ from x to z and then from z to y. This observation is used in the following result. A.4. Lemma. Let X be a real locally convex space of dimension greater than one and uppose that C is a non-empty open convex set in X not containing o. Then there is a non-zero element x of X such that CnRx = 0. Proof. The set S = Ut>otC is open in X. If x and -x are both in S, then there are positive numbers t l , t2 and there are elements Xl, x2 E C such that x = tixi = -t2X2. Convexity of C implies that 0 = (tixi +t2 X 2)/(t l +t2) E C, which is impossible by hypothesis. Hence, S n (-S) = 0. Since X \ {OJ is connected, it cannot be the union of the disjoint non-empty open sets S and -S, so there is a non-zero vector x of X such that neither x nor -x belongs to S. Hence, en Rx = 0. 0 Let p be a seminorm on a vector space X, and let Y be a vector subspace of X. Define a seminorm p' on X/V by setting p'(x + Y) = inf p(x + y). yeY If X is a locally convex space, so is X/V when endowed with the largest topology making the quotient map 7r: X  X/V continuous. For we may take a generating family r of semi norms for the topology of X such that maxljnPj E r if PI,...,Pn E r, and then r' = {p' I pEr} is a generating family of seminorms for the topology of X/V. To see this, it suffices to show that 7r is continuous and open when X/V is endowed with the topology generated by r'. Continuity of 7r is clear, and to show that 7r is open it suffices to observe that the image of a basic set U (c; p) under 7r is the basic set U(c; p') in X/Yo A.5. Theorem. If T is a non-zero continuous linear functional on a locally convex space X, then it is open. Proof. If Y = ker( T), then X / Y is a one-dimensional locally convex space, and the linear functional T': X/V -+ K, x + Y .-...+ T(X), is a continu- ous linear isomorphism. Hence, X/V is Hausdorff, as K is. Thus, if r is a generating family of seminorms of X / Y, then one of them, P say, must be non-zero. Using the fact that X/V is one-dimensional, every element of r must therefore be of the form o:p for some 0: E R+. Hence, X/V is in fact a normed vector space with norm P inducing its topology. Since p 0 T,-l 
270 Appendix is a norm on K, it is equivalent to the usual norm, and therefore r' is a homeomorphism. Hence, r is open, since it is the composition of r' and the quotient map from X to X/Y, and both of these maps are open. 0 A.I. Remark. If X is a complex vector space, then for each real-linear functional p: X -+ R there is a unique complex-linear functional r: X -+ C such that Re( r) = p (set r(x) = p(x) - ip( ix)). We shall use this to reduce some arguments to the "real" case. A.6. Theorem. Let X be a locally convex space and C a non-empty open convex set of X not containing o. Then there is a continuous linear functional r on X such that Re( r( x)) > 0 (x E C). Proof. By Remark A.1 we may suppose that the ground field is R. Let A be the set of all closed vector subspaces Y of X disjoint from C, and make it into a poset using the partial ordering given by set inclusion. Observe that A is non-empty, since the closure of the zero space is an element. If S is any totally ordered set of elements of A, then (US)- is an element of A majorising all elements of S. By Zorn's lemma, A admits a maximal element, Y say. Suppose now that X/Y is not one-dimensional (and we shall deduce a contradiction). Let 7r: X -+ X / Y be the quotient map. Then 7r( C) is a non-empty open convex subset of X/Y not containing 0, so by Lemma A.4 there is an element x E X such that 7r(x) is non-zero and 7r(C)nR7r(x) = 0. Put Y I = Y + Rx. Then Y 1 - E A and contains Y, so Y 1 - = Y by maximality of Y. Hence, x E Y, so 7r( x) = 0, a contradiction. Therefore, X/Y must be one-dimensional; that is, Y is of codimension one in X. It follows that there is a linear functional r on X with kernel Y, and r is necessarily continuous by Theorem A.3. Since r( C) is convex, it is an interval of R, and because it does not contain 0, it is contained in (0, +(0) or (-00,0). By replacing r with -r if necessary, we can suppose that r( C) C (0, +00), and this proves the result. 0 There are a number of closely related results in the literature that are called separation theorems. The following is one of them. A. 7. Theorem. Let C be a non-empty closed convex set in a locally convex space X and x E X \ C. Then there is a continuous linear functional r on X and a real number t such that Re(r(y)) < t < Re(r(x)) for all y E c. Proof. We may suppose that K = R. Let r be a generating family of semi norms for X. Since -x + C is a closed set not containing 0, there is a positive number c and there are elements PI,... ,pn of r such that the set U = U(C;PI,... ,Pn) is disjoint from -x + C. Hence, W = x + U is open and disjoint from C. The set W - C = UyECW - Y is open, and it is convex because Wand C are convex. Moreover, it does not contain o. 
Appendix 271 By Theorem A.6, there is a continuous linear functional r on X such that r(y) > 0 for all yEW - C. Now U  ker(r), since X = U  ImU, so there exists z E U such that r(z) < o. Setting t = r(x + z) and observing that x + z E W, we get the inequality r(x + z - y) > 0 for all y E C, so r(x) > t > r(y). 0 A.8. Corollary. Let C be a convex set in a locally convex space X. Then for any point x E X, x E C if and only if there is a net (XA)AEA in C such that (r(x A )) converges to r(x) for all continuous linear functionals r on X. Proof. Observe that C is convex, and apply Theorem A.7. o A.9. Corollary. Let Y be a closed vector subspace of a locally convex space X and x E X \ Y. Then there is a continuous linear functional r on X such that r(y) = 0 (y E Y) and rex) = 1. Proof. By Theorem A.7, there is a continuous linear functional p on X and a real number t such that Re(p(y)) < t < Re(p(x)) (y E V). Since o E Y, therefore t > 0, and p(x) =I- O. For each n E Nand y E Y, we have In Re(p(y))1 < t. Hence, Re(p(y)) = 0 (y E V). It follows p = 0 on Y. Set r = p/p(x). 0 A.I0. Corollary. Let x, y be distinct points of a Hausdorff locally convex space X. Then there is a continuous linear functional r on X such that rex) =I- r(y). Proof. Observe that x - y rf- Y = {O}, and apply Corollary A.9. 0 The usual form of the Hahn-Banach theorem asserts that if r is a bounded linear functional on a vector subspace Y of a normed vector space X, then there is a bounded linear functional r' on X extending r and of the same norm. We prove an analogue of this for locally convex spaces. If Y is a vector subspace of a locally convex space X, then Y is also a locally convex space when endowed with the relative topology. For if r is a generating family of seminorms for the topology of X, it is readily verified that r' = {p' I pEr} is a generating family of seminorms for the topology of Y, where p' is the restriction to Y of the seminorm p on X. A.ll. Theorem. Let r be a continuous linear functional on a vector subspace Y of a locally convex space X. Then there is a continuous linear functional r' on X extending r. Proof. Let r be a generating family of semi norms for the topology of X. Since r is continuous, it follows from Theorem A.1 that there are semi norms PI,... ,Pn E r and a positive number M such that Ir(y)1 < p(y) for all y E Y, where p is the continuous seminorm on X defined 
272 Appendix by p(x) = MmaxljnPj(x). If Z = p-I{O}, then Z is a closed vector subspace of X, and X/Z is a normed vector space under the well-defined norm IIx + ZII = p(x). Let 7r:X -+ X/Z be the quotient map. Then p: 7r(Y)  K, 7r(Y)  r(y), is a well-defined norm-decreasing linear functional. By the Hahn-Banach theorem, there is a linear functional p' on X/Z extending p and also norm- decreasing. It follows that the function T':X -+ K, x  p'(x + Z), is linear and 'T'(x)' < p(x) (x E X). Using Theorem A.1 again, T' IS continuous, and it clearly extends T. 0 The intersection of a family of convex subsets of a locally convex space X is itself convex. Hence, if S is a subset of X, there is a smallest convex subset co(S) containing S. We call co(S) the convex hull of S. It is easily verified that n n co( S) = {L t i x i I n > 1, t I , . · . , t n E R +, L t i = 1, X I , . · . , X n E S}. i=l i=l We write co (S) for the closure of co(S), and we observe that co (S) is the smallest closed convex set of X containing S. We call co (S) the cl03ed convex hull of S. A.2. Remark. If C I , . . . , C n are non-empty convex sets of X, then the convex hull co(C I U.. .UC n ) is the set of all elements tlxl +.. .+tnx n , where t l , . . . , t n are non-negative numbers such that t l +. . .+t n = 1 and Xl, . . . , x n are in C I , . . . , C n, respectively (the proof of this is an easy exercise). A.12. Theorem. Let C t ,..., C n be convex compact sets in a locally convex space X. Then CO(CI U ... U Cn) is compact. Proof. We may suppose that C 1 , . . . , C n are all non-empty. Let  denote the set of all non-negative numbers tl,. . . , t n such that t l + . . . + t n = 1. Clearly,  is compact in R n and the map n  XCI X . . . x C n -+ co( C I U . . . U C n), (t I , . . . , tn, X I , . . · , X n) ....... L t j x j , j=l is a continuous surjection. Hence, co( C I U . . . U Cn) is compact, since the product  x C I . . . X C n is compact. 0 
Appendix 273 A point x of a convex set C in a vector space X is an extreme point of C, if the condition x = ty + (1 - t)z, where y, z E C and 0 < t < 1, implies that x = y = z. Equivalently, x is an extreme point of C if and only if, whenever y, z E C are such that x = (y + z)/2, we necessarily have x = y = z. A non-empty convex subset F of C is a face of C if, whenever x E F and y, z E C and x = ty + (1 - t)z for some t E (0,1), we must have y,z E F. If C is non-empty, it is a face of C, and a non-empty intersection of a family of faces of C is a face of C. A point x E C is an extreme point of C if and only if {x} is a face of C. An extreme point of a face of C is an extreme point of C itself. A.13. Lemma. Let C be a non-empty convex compact set in a locally convex space X and suppose that r is a continuous linear functional on X. Let M be the supremum of all Re( r( x )) where x ranges over C. Then the set F of all x E C such that Re( r( x)) = M is a compact face of C. Proof. The set F is non-empty, since compactness of C implies that there is a point Xo of C such that M = Re( r( xo)). It is clear that F is convex. It is also closed in C and therefore compact. Suppose that x E F, and y,z E C and x = ty + (1 - t)z for some t E (0,1). Then M = Re( r(x)) = t Re( r(y)) + (1 - t) Re( r(z)). If y or z is not in F, then Re(r(y)) or Re(r(z)) is less than M, so tRe(r(y)) + (1- t)Re(r(z)) < tM + (1- t)M, and therefore M < M, a contradiction. This argument shows that y, z are in F and consequently F is a face of C. 0 The equation C = co (E) in the following is the Krein-Milman theorem, one of the great results of functional analysis with a vast range of applica- tions. A.14. Theorem. Let C be a non-empty convex compact set in a Hausdorff locally convex space X. Then the set E of extreme points of C is non-empty and C = co (E). Moreover, if S is a closed set ofC such that co (S) = C, then S contains E. Proof. Let A denote the set of all compact faces of C. This is non-empty, because C E A. We make A into a poset by setting F < F ' in A if F ' C F. A totally ordered family of elements of A has the finite intersection property, so by compactness of C its intersection is non-empty, and is therefore a face 
274 Appendix of C. Hence, every totally ordered family of A is majorised by an element of A, so by Zorn's lemma there is a maximal element, F say, in A. We claim that F is a singleton set. For suppose otherwise and let x, y be distinct elements of F. Then by Corollary A.10 there is a continuous linear functional r on X such that Re( r( x)) i: Re( r(y)). Set M = sup{Re(r(z)) I z E F}. Then the set Fo = {z E F I Re(r(z)) = M} is a compact face of F by Lemma A.13. Hence, Fo is a face of C, so Fo E A, and therefore, by maximality of F, we have Fo = F. Consequently, M is equal to Re( r( x)) and Re( r(y)), a contradiction, since these two numbers are distinct. Thus, F has to be a singleton, {x} say. Hence, x E E, so E is non-empty. Suppose now that co ( E) =I c, so there is a point z E C \ co ( E). By Theorem A. 7, there is a continuous linear functional r on X and a number t E R such that Re(r(y)) < t < Re(r(z)) for all y E co (E). Set M' = sup{Re( r(y)) lyE C} and F' = {y E C I Re( r(y)) = M'}. By Lemma A.13 again, F' is a compact face of C. It follows from the earlier part of this proof that F' has an extreme point, y say. Hence, y is an extreme point of C, and so y E E. But Y E F' implies that M' = Re(r(y)), so M' < t < Re( r( z)) < M', a contradiction. This argument, therefore, proves that co (E) = C. To show that an extreme point z of C lies in the set S, it suffices to show that for each neighbourhood U of 0 in X the intersection (z + U) n S is non-empty. If r is a family of semi norms generating the topology of X, then it is clear that we may suppose that U = U( €; PI, . . . , Pm) for some positive number € and for some seminorms PI,... ,Pm E r. Set W = tu, and observe that W is open. Now S C UxES(X+ W), so by compactness of S there exist elements Yl, . . . , Yn E S such that S C (Yl + W) U. . . U (Yn + W). For j = 1, . . . , n, set Sj = (Yj + W-) n C. Since U is convex, so is Yj + W-, and therefore, by convexity of C, the set Sj is also convex. It is clear that Sj is also compact and non-empty. Observe also that S C SI U . . . uSn. The set K = CO(SI U . . . USn) is compact by Theorem A.12, and since each set Sj C I<, we have S C K. Hence, C C 1<, since co (S) = C. In particular, z E I<. By Remark A.2, we can write z in the form z = tlXl + ... +tnx n , where t l ,...,t n E R+ and 
Appendix 275 t 1 + ... +t n = 1, and Xj E Sj for j = 1,... ,n. Since Xl,...,X n E C, and since z is an extreme point of C, it follows that z = Xj for some index j (take any j such that tj > 0). Hence, z E Sj, and therefore, z - Yj E W-. Consequently, the set (z - Yj + W) n W is non-empty. If x is an element, then Pi ( x) < £ / 2 and Pi ( z - Y j - x) < c /2, so Pi ( z - Y j) < c for i = 1, . . . , m. Hence, Yj - z E U, and therefore, (z + U) n S is non-empty. This proves the theorem. 0 
Notes These notes are very incomplete, and concerned only with those aspects of the theory most relevant to the material covered in this book. FUnctional analysis began to evolve at the turn of the century out of the work of Fredholm, Frechet, Hilbert, F. Riesz, and Volterra on integral equations, eigenvalue problems, and orthogonal expansions. The spectral theorem is due to Hilbert. The first abstract treatment of normed vector spaces is due to Banach in his 1920 thesis. The theory of von Neumann algebras originated in a paper published in 1929 by von Neumann [vN], where he introduced these algebras under the name "rings of operators," and in which he proved his famous double commutant theorem. Applications to theoretical physics provided a moti- vation for von Neumann's interest. He treated the foundations of quantum mechanics from the point of view of operator algebras. Gelfand and Naimark introduced the class of C*-algebras in 1943 [GN]. They proved one of the most fundamental results of the theory by showing that C*-algebras can be faithfully represented as closed self-adjoint algebras of operators on Hilbert spaces. The principal results of the early theory of C*-algebras are due to Fell, Glimm, Kadison, Kaplansky, Mackey, and Segal, among others. The uniformly hyperfinite algebras (UHF algebras) form an important class of C*-algebras, and were first studied by Glimm in the early 1960s. The more general class of AF -algebras was introduced by Bratteli (early 1970s), who initiated their classification. This classification was completed by Elliott, whose formulation was not originally stated in K-theoretic terms (as it is in the present text). The study of tensor products of C*-algebras was initiated by Turumaru in 1952, but major progress in the theory did not commence until the mid 1960s. Some important contributors to this area are Effros, Guichardet, Lance, and Takesaki. 277 
278 Notes Homological algebraic methods made a very significant impact on oper- ator theory in the early 1970s with the work of Brown, Douglas, and Fillmore on the classification of essentially normal operators [BDF]. The K-theory of C*-algebras, and a powerful generalisation called KK-theory and due to Kasparov, have had profound applications to both C*-algebra theory, and to other areas, such as differential geometry. Two early and fundamental results are the Pimsner- Voiculescu six-term exact sequence for computing the K-theory of the crossed product of a C*-algebra with Z [PV], and Connes' theorem for computing the K-theory of crossed prod- ucts of C*-algebras with R [Con 1]. Single operator theory and operator algebra theory are vastly more extensive then an introductory volume such as this can indicate. For a deeper understanding of the subject the reader is referred to [Bla], [Dix 1], [Dix 2], [KR 1], [KR 2], [Ped], [Sak], and [Tak]. 
References [BDF] L. Brown, R. Douglas, and P. Fillmore, Unitary equivalence modulo the compact operators and extensions of C*-algebras, Proc. Conf. on Oper- ator Theory, Springer Lecture Notes in Math. 345 (1973), 58-128. [Bla] B. Blackadar, K- Theory for Operator Algebras. MSRI publications no. 5, Springer- Verlag, New York, 1986. [BMSW] B. A. Barnes, G. J. Murphy, M. R. Smyth, and T. T. West, Riesz and Fredholm Theory in Banach Algebras. Research Notes in Mathematics 67, Pitman, London, 1982. [Cnw 1] J. B. Conway, Subnormal Operators. Pitman, Boston, 1981. [Cnw 2] J. B. Conway, A Course in Functional Analysis. Graduate Texts in Mathematics 96, Springer-Verlag, New York, 1985. [Coh] D.L. Cohn, Measure Theory. Birkhauser, Boston, 1980. [Con 1] A. Connes, An analogue of the Thom isomorphism for crossed prod- ucts of a C*-algebra by an action of R, Advances in Math. 39 (1981), 31-55. [Con 2] A. Connes, Non Commutative Differential Geometry. Chapter I: The Chern Character in K Homology. Chapter II: De Rham Homology and Non Commutative Algebra. Publ. Math. I.H.E.B. 62 (1986),257-360. [Cun] J. Cuntz, K-Theory for certain C*-algebras, Ann of Math. 113 (1981), 181-197. [Dix 1] J. Dixmier, Von Neumann Algebras. North-Holland, Amsterdam, 1981. [Dix 2] J. Dixmier, C*-Algebras. North-Holland, Amsterdam, 1982. [Dou 1] R. G. Douglas, Banach Algebra Techniques in Operator Theory. Aca- demic Press, New York, 1972. [Dou 2] R. G. Douglas, On the C*-algebra of a one-parameter semigroup of isometries, Acta Math. 128 (1972), 143-152. [Eff] E. Effros, Dimensions and C*-Algebra3. CBMS Regional Conf. Sere in Math., no. 46, Amer. Math. Soc., Providence, 1981. 279 
280 References [Enf] P. Enflo, A counterexample to the approximation problem in Banach spaces, Acta Math. 130 (1973), 309-317. [GN] I. Gelfand and M. Naimark, On the embedding of normed rings into the ring of operators in Hilbert space, Mat. Sb. 12 (1943), 197-213. [Goo] K. Goodearl, Note3 on Real and Complex C*-Algebra3. Shiva Publish- ing Ltd., Nantwich, 1982. [Hal] P. R. Halmos, A Hilbert Space Problem Book. Springer-Verlag, New York, 1982. [Kel] J. L. Kelley, General Topology. Springer-Verlag, New York, 1975. [KR 1] R. V. Kadison and J. R. Ringrose, Fundamental3 of the Theory of Operator Algebra3 I. Academic Press, New York, 1983. [KR 2] R. V. Kadison and J. R. Ringrose, Fundamental3 of the Theory of Operator Algebra3 II. Academic Press, New York, 1986. [Lan] E. C. Lance, Tensor products and nuclear C*-algebras, Operator Alge- bra3 and Application3 (ed. R. V. Kadison), Proc. Symp. Pure Math. Pt 1 38 (1982), 379-399. [Mur] G. J. Murphy, Ordered Groups and Toeplitz Algebras, J. Operator Theory 18 (1987), 303-326. [Ped] G. K. Pedersen, C*-Algebra3 and their A utomorphi3m Group3. Aca- demic Press, London, 1979. [PV] M. Pimsner and D. Voiculecu, Exact sequences for K-groups and Ext-groups of certain cross-products of C*-algebras, J. Operator Theory 4 (1980), 93-118. [Rie] M. Rieffel, C*-algebras associated with irrational rotations, Pacific J. Math. 93 (1981),415-429. [Rud 1] W. Rudin, Real and Complex AnalY3i3. McGraw-Hill, New York, 1966. [Rud 2] W. Rudin, Functional AnalY3i3. Tata McGraw-Hill, New Delhi, 1977. [Sak] S. Sakai, C*-Algebra3 and W*-Algebra3. Springer-Verlag, New York, 1971. [Tak] M. Takesaki, Theory of Operator Algebra3 I. Springer-Verlag, New York, 1979. [TL] A. E. Taylor and D. C. Lay, Introduction to Functional AnalY3i3. Wiley, New York, 1980. [Top] D. M. Topping, Lecture3 on Yon Neumann Algebra3. Van Nostrand, London, 1971. [vN] J. von Neumann, Zur Algebra der Funktionaloperationen und Theorie der normalen Operatoren, Math. Ann. 102 (1929),370-427. [Wes] T. T. West, The decomposition of Riesz operators, Proc. London Math. Soc. (9) 16 (1966),737-752. 
Notation Index a , 15 , 268 |a| , 45 , 46 a1/2 , 45 a+ , a~ , 45 A , 102 A , 12 , 39 A+ , 45 A» , 128 A , 160 Ap , 117 Asa , 36 Adw , 36 , 56 AH , 245 A5 , [A5] , 120 A <g>* B , 190 . 4 <g>7 B , 190 A ®ma« B , 193 fo , 259 Boo(fi) , 3 , 37 , 66 B(X) , B(X , Y) , 3 , 20 IMU , 190 \\c\\max , 193 co(5) , 272 C , C" , 115 Cv , 251 C*{a) , 41 C(A) , 246 C»(ft) , 2 Co(fi) , 2 , 37 Co(n , A) , 37 9 , 255 def(u) , 23 D , 3 ~ , 122 , 218 ss , 219 , 233 ¥>„ , 220 , 247 <£ , 40 , 229 <p®il> , 190 ¥>®*^ , 210 fx (f e c0(Q) , x e X) , 207 F(H) , 55 1a , 249 hull(5) , 157 hull'(S) , 160 HT , (Hr , <pT) , 94 [H , <p] , 160 (fr , ^)B , if , 162 (ff , V>)u , 163 # ® if , 186 im(u) , 48 ind(u) , 23 , 50 Jx (J ideal) , 161 k , ka , 241 Kyi , 254 ker(iJ) , 157 K , 256 K(X) , K(X , Y) , 20 K0(A) , 220 K0(A)+ , 219 tf , (A) , 229 , 247 Ki(tp) , 247 limA„ , 176 l]mGn , 174 !*(#) , 65 L2(#) , 60 L°°(fi , /i) , 2 , 37 Mv , 67 M(^) , 38 M(ft) , 66 Mn(A) , 94 nul(u) , 23 NT , 93 Q(A) , 14 , 15 T/i , ""B , 192 [p] , \pU , 219 P[i4] , 218 Prim(A) , 156 , 159 PS(A) , 144 r(a) , 9 , 10 a(a) , (JA(a) , 6 , 13 S (algebra) , 249 S (sequences) , 180 5X (S C PS(A)) , 161 S(A) (states) , 89 S(A) (suspension) , 246 S(v>) , 247 ©AA (algebras) , 30 , 37 ©A"A (operators) , 105 ®x(Hx , V\) , 93 tr , 63 , 65 , 179 r ®7 p , 199 T , 4 T„ , 99 u* (adjoint) , 48 u* (transpose) , 21 Up , 117 u ® t ; (operator) , 187 *7(A) , 17°(A) , 230 Un(A) , U°n(A) , 230 x (8) y (as operator) , 55 Zv , 251 281
This page is intentionally left blank
Subject Index A Adjoint , 48 AF - algebra , 183 , 196 Algebra , 1 normed , 1 * - algebra , 35 Approximate unit , 77 Ascent of an operator , 22 , 23 Atkinson theorem , 28 Automorphism , 36 B Banach algebra , 2 abelian , 15 Banach * - algebra , 36 Beurling spectral radius formula , 10 Beurling space , 98 Borel functional calculus , 72 Bott periodicity , 259 C C* - algebra , 36 abelian , 41 , 205 irreducible , 58 , 120 non - degenerate , 120 primitive , 157 simple , 86 Calkin algebra , 30 , 123 Cauchy - Schwarz inequality , 52 C* - completion , 175 CCR algebra , 167 Character , 14 , 40 , 144 Character space , 14 , 15 C* - norm , 175 Coburn theorem , 106 Commutant , 115 double , 115 Commutator ideal , 102 Compact operator , 18 , 25 , 54 , 123 Cone of a C* - algebra , 246 Cone in a group , 222 Connecting homomorphism , 255 Continuity of K0 , 236 Continuity of K0 , 239 Contractible C* - algebra , 246 Convex hull , 272 closed , 272 C* - seminorm , 175 C* - subalgebra , 36 hereditary , 83 , 166 Cuntz theorem , 256 Cyclic representation , 140 Cyclic vector , 134 D Defect of an operator , 23 Density theorem , 131 Descent of an operator , 22 Diagonal(isable) operator , 26 , 54 , 55 Direct limit of C* - algebras , 176 283
284 Index of groups , 174 Direct sequence of C* - algebras , 175 of groups , 173 Direct sum of Banach algebras , 30 of C* - algebras , 37 of representations , 93 Disc algebra , 3 , 5 , 16 Double centraliser , 38 Double commutant theorem , 116 E Elliott theorem , 226 Equivalence of projections , 122 , 218 stable , 219 Essential ideal , 82 Essential spectrum , 30 Exact sequence , 231 , 232 Extension of C* - algebras , 211 Extreme point , 273 F Face , 273 Factor , 182 Faithful representation , 93 Finite - dimensional C* - algebra , 194 Fredholm alternative , 25 , 26 Fredholm operator , 23 , 103 Functional calculus , 43 Borel , 72 G GCR algebra , 169 Gelfand - Mazur theorem , 9 Gelfand - Naimark - Segal representation , 94 , 140 Gelfand - Naimark theorem , 94 Gelfand representation , 15 Gelfand theorem , 9 Gelfand transform , 15 Generating set for a(n) C* - algebra , 41 hereditary C* - algebra , 83 ideal , 4 von Neumann algebra , 117 GNS representation , 94 , 140 H Hardy space , 96 Hartman - Wintner theorem , 100 Hereditary C* - subalgebra , 83 , 166 Hermitian element , 36 , 37 , 40 Hilbert - Schmidt norm , 59 , 61 Hilbert - Schmidt operator , 60 Homomorphism , 5 * - homomorphism , 36 Homotopic , 233 Homotopy , 233 Hull - kernel topology , 159 Hyperfmite algebra , 182 I Ideal , 4 , 79 essential , 82 maximal , 4 modular , 4 , 13 Idempotent operator , 27 Index of an operator , 23 , 24 , 29 , 104 Inflation , 94 Integral operator , 19 , 26 , 61 Invariant subspace , 58 Involution , 35 Irreducibility , 58 , 120 , 143 algebraic , 152 Isometry , 36 ♦ - isomorphism , 36 J Jacobson topology , 159 Jordan decomposition , 92 K Kadison transitivity theorem , 150 Kaplansky density theorem , 131 Krein - Milman theorem , 273 L Left ideal , 4 maximal , 4 , 155 modular , 13 , 155 Liminal C* - algebra , 167
Index Locally convex (topological vector) space , 267 M Mapping cone , 251 Matrix algebra , 94 Maximal abelian von Neumann algebra , 134 Maximal C* - norm , 193 Maximal ideal , 4 Maximal tensor product , 193 Modular ideal , 4 , 13 Multiplication operator , 67 Multiplier algebra , 39 , 82 Murray - von Neumann equivalence , 122 N Natural map , 174 , 176 Non - degenerate C* - algebra , 120 Non - degenerate representation , 142 Normal element , 36 , 90 Normed algebra , 1 Nuclear C* - algebra , 193 , 205 , 212 Nullity of an operator , 23 O Operator , 3 bounded below , 21 compact , 18 , 25 , 54 , 123 multiplication , 67 normal , 36 , 72 , 136 Operator matrix , 95 Order isomorphism , 223 Ordered group , 110 Orthogonal * - homomorphisms , 245 P Partial isometry , 50 Partially ordered group , 222 Polar decomposition , 51 , 119 Polarisation identity , 49 Positive element , 45 Positive group homomorphism , 223 Positive linear functional , 87 Postliminal C* - algebra , 169 285 Prime ideal , 158 Primitive C* - algebra , 157 Primitive ideal , 156 Projection Banach space , 27 self - adjoint , 36 , 50 Pseudo - inverse , 27 Pullback , 251 Pure state , 144 Q Quasinilpotent element , 16 Quotient algebra , 4 Quotient C* - algebra , 80 R Range projection , 119 Reducing subspace , 50 Representation , 93 cyclic , 140 direct sum , 93 equivalent , 143 faithful , 93 GNS , 94 , 140 irreducible , 143 non - degenerate , 142 restriction , 163 universal , 94 Resolution of the identity , 72 Riesz theorem , 99 S Self - adjoint element , 36 , 37 , 40 Self - adjoint functional , 92 Self - adjoint set , 35 Separating family of seminorms , 267 Separating vector , 134 Separation theorem , 270 Sesquilinear form , 49 bounded , 52 , 53 hermitian , 52 norm of a , 52 positive , 52 Short exact sequence
286 Index of C* - algebras , 211 of groups , 232 Simple C* - algebra , 86 Spatial C* - norm , 190 , 208 Spatial tensor product , 190 Spectral mapping theorem , 43 Spectral measure , 66 Spectral radius , 9 , 13 Spectral theorem , 72 Spectrum (of an algebra) , 15 , 160 Spectrum (of an element) , 6 , 9 , 13 Split short exact sequence , 232 , 256 Square root in a C* - algebra , 45 Stability of K0 , 244 Stable equivalence , 219 Stably finite C* - algebra , 221 State , 89 pure , 144 tracial , 179 Strongly continuous function , 130 Strong (operator) topology , 113 Subalgebra , 1 * - subalgebra , 35 Suspension C* - algebra , 246 Symbol (of Toeplitz operator) , 99 Symmetry , 231 T Takesaki theorem , 205 Tensor product of C* - algebras , 190 of Hilbert spaces , 186 Toeplitz algebra , 102 Toeplitz operator , 99 Topology generated by seminorms , 267 strong (operator) , 113 cr - weak = ultraweak , 126 weak (operator) , 126 Trace - class norm , 63 , 64 Trace - class operator , 63 Trace (of an operator) , 63 Tracial positive functional , 179 Transitivity theorem , 150 Transpose , 21 Trigonometric polynomial , 96 Type I C* - algebra , 169 U UHF algebra , 180 Ultraweak topology , 126 Uniformly hyperfinite algebra , 180 Unilateral shift , 30 , 99 , 106 Unital homomorphism , 5 Unital normed algebra , 1 Unitarily invariant set , 161 Unitary element , 36 , 42 Unitary equivalence of elements , 36 of representations , 143 Unitary operator , 49 , 73 Unitisation , 12 , 35 , 39 Universal representation , 94 V Vigier theorem , 113 Volterra integral operator , 20 von Neumann algebra , 116 , 128 abelian , 136 von Neumann double commutant theorem , 116 W Weak exactness , 232 , 250 Weak (operator) topology , 126 Weak* topology , 268 Wiener space , 98 Wiener theorem , 18 Wold - von Neumann decomposition , 105
This book introduces the reader , graduate student , and non - specialist alike to a lively and important area of mathematics . By its careful and detailed presentation , the book enables the reader to approach the contemporary literature with confidence . A plentiful number of exercises and the choice of topics , which reflect current research interests , distinguish this book from other texts . In addition to the basic theorems of operator theory , including the spectral theorem , the Gelfand - Naimark theorem , the double commutant theorem , and the Kaplanski density theorem , some major topics covered by this text are : K - theory , tensor products , and representation theory of C* - algebras . ISBN 0 - 12 - 511360 - 9