LONDON MATHEMATICAL SOCIETY LECTURE NOTE SERIES Managing Editor: Professor N J. Hitchin. Mathematical Institute.
University of Oxford . 24-2'1 St Giles. Oxford OX I 3LB. United Kingdom
The titles below are availahle from booksellers. or from Cambridge Uni v ersity Press at www .cambridge.org 46
p-adic Analys i s: a shon course on recent work,
59
Applica bl e differential geometry.
86
Topolog ical topics.
N. KOBLITZ
M. CRAMPIN & F.A.E. PIRANI
I.M. JAMES (ed )
88
FPF ring theory.
90
Pol yt opes and symmetry.
C. FAITH & S. PAGE
96
Dio phantine equations over function fields.
'17
Varieties of constructive mathematics.
S.A. ROBERTSON R.C. MASON
D.S. BRIDGES & F. RICHMAN
'1'1 100
Methods of differential geometry in algebraic topology. M. KAROUBI & C. LERUSTE Stopping time techniques for analysts and prohahilists. L. EGGHE
105
A local spectral theory for closed operators.
107
Compactification of Siegel moduli schemes.
109
Dio phanti ne analysis.
1 13
I. ERDELYI & WAN G SHENGWANG C.-L. CHAI
J. LOXTON & A. VAN DER POORTEN ( ed s)
Lectures on the asymptotic theory of ideals.
D. REES
116
Representations o f algebras.
119
Triangulated catego ri e s in the representation theory of finite-dimensional algebras.
121 128
Proceedings of Gro/lPS
-
PJ. WEBB (ed)
Sf Andre ...s 1985.
D. HAPPEL
E. ROBERTSON & C. CAMPBELL (eds)
Descriptive set theory and the structure of sets of uniqueness.
Model theory and modules.
Algeb ra ic . extremal & metri c combinatorics. M.-M. DEZA. P. FRANKL & I.G. ROSENBERG (eds)
140
M. PREST
A.S. KECHRIS & A. LOUVEAU
130 131
Geometric aspects of Banach spaces.
E.M. PEINADOR & A. RODES (eds)
141
Sur v eys in comb ina tori cs 19N9.
J. SIEMONS (ed)
144
Introduction to unifor-m spaces,
I.M. JAMES
146
Cohen - Macaulay modules over Co hen - Macaulay rings. A.N. RUDAKOV ef ,,/
Y. YOSHINO
148
Helices and vector bundles.
149
Solitons. nonlinear- evolution equations and inver-se scattering, M. ABLOWITZ & P. CLARKSON
150
Geometry of low-dimensional man i folds I.
S. DONALDSON & C.B. THOMAS (eds)
151
Geometry of low-dimensional manifolds 2.
S. DONALDSON & C.B. THOMAS (e ds)
152
Oligomorphic permutation groups.
153
L- function s and arithmetic.
155
Classification theories of polarized varieties, TAKAO FUJITA Geometry of Banach spaces. P.FX. MOLLER & W. SCHACHERMA YER (eds)
158
P. CAMERON
J. COATES & M.J. TAYLOR (eds)
159
Groups St Andrews 1989 volume I.
C.M. CAMPBELL & E.F. ROBERTSON ( eds)
160
Groups St Andrews 1989 volume 2,
C.M. CAMPBELL & E.F. ROBERTSON ( ed s)
161
Lectures on block theory.
163
BURKHARD KOLSHAMMER
Topics in var-ieties of group representations,
Quasi-symmetric designs.
Surveys in com binatori cs. 1991.
168
S.M. VOVSI
M.S. SH�IKANDE & S.S. SANE
164 166
Representations of algebras.
A.D. KEEDWELL (ed)
H. TACHIKAWA & S. BRENNER (eds)
Boolean function complexity.
M.S. PATERSON (ed)
169 170
Manifolds with singulari t ies and the Adams-Novikov spectral sequence.
171
Squares.
172
Algebraic varieties.
173 174
Le ctures on mechanics.
175 176 177 178 179 180
B. BOTVINNIK
A.R. RAJWADE
GEORGE R. KEMPF
Discrete groups and geometry.
W.J. HARVEY & c. MACLACHLAN (eds)
J.E. MARSDEN Ad ams memorial symposium on algebraic topology I. A d ams memorial symposium on algebraic topology 2.
Applic ations of categories in computer science.
L ower K- and L-theory.
A. RANICKI
Complex projective geometry.
N. RAY & G. WALKER (eds)
N. RAY & G. WALKER (eds)
M. FOURMAN. P. J OHNST ONE & A. PITTS (eds )
G. ELLlNGSRUD ef 01
183
Lectures on ergodi c theory and Pes in the ory on compact manifolds, GA. NIBLO & M .A . ROLLER (eds) Geometric group theory II. G. A. NIBLO & M.A. ROLLER (eds) Shintani zeta functions. A. YUKIE
184 185 186 187
Ar ithmet ical functions. W. SCHWARZ & J. SPILKER Re pr es entations of solvable groups. O. MANZ & T.R. WOLF Complex ity: knots. colourings and counting. D.J.A. WELSH Surveys in combinatorics. 1993. K. WALKER (ed)
188 189 190 191 192 194 195 196 197 198
Local analysis for the odd order theorem. H. BENDER & G. GLAUBERMAN Locally presentable and accessible categories. J. ADAMEK & J. RO SICKY Polynomial invariants of finite groups, D.J. BENSON Finite geometry and combinatorics. F. DE CLERCK ef al Symplectic geometry, D. SALAMON (ed)
181 182
199 200 201 202 203 204 205 207
Geometric group theory I.
Independent random variahles and rearrangement invariant spaces. Arithmetic of blowup al gebras , WOLMER VASCONCELOS
M. POLLICOTT
M. BRAVERMAN
Microlocal an aly sis for differential operators. A. GRIGIS & J. SJO ST RAND Two-dimensional homotopy and combinatorial group theory, C. HOG-ANGELONI ef al The algebraic characterization of geometric 4-mani folds , JA. HILLMAN Invariant potential theory in the unit ball of cn, MANFRED STOLL The Grothendieck theory of dessins d'enfant. L. SCHNEPS (ed) Singularities, JEAN-PAUL BRASSELET (ed) The technique of pseudodifferential operators, H.O. CORDES Hochschild cohomology of von Neumann algebras, A. SINCLAIR & R. SMITH Combinatorial and geometric group theory, AJ. DUNCAN, N.D. GILBERT & J. HOWIE (eds) Ergodic theory and its connections with harmonic anaJysis, K. PETERSEN & I. SALAMA (eds) Groups of Lie type and their geometries, W M. KANTOR & L. DI MARTINO (eds)
208 209 210
N.J. HITCHIN. P. NEWSTEAD & W.M. OXBURY (eds)
Vector bundles in algebraic geometry.
Arithmetic of diagonal hypersurfaces over finite fields.
F.Q. GOUVEA & N. YUI
211
Hilben C··modules. E.C. LANCE Groups 93 Galway I St Andrews I. C.M. CAMPBELL et af(eds)
214
Generalised Euler·Jacobi inversion formula and asymptotics beyond all orders. V. KOWALENKO,I af
215 216
Stochastic panial differential equations. A. ETHERIDGE (ed)
218 220
Algebraic set theory.
212
217
Groups 93 Galway I St Andrews II.
Number theory 1992-93.
C.M. CAMPBELL 'I af (eds)
S. DAVID(ed)
Quadratic forms with applications to algebraic geometry and topology.
Surveys in combinatorics. 1995.
221
Harmonic approximation.
222
Advances in linear logic.
SJ. GARDINER
J.-Y. GIRARD. Y. LAFONT & L. REGNIER (eds)
223
Analytic semigroups and semilinear initial boundary value problems.
224
Computability. enumerability. unsolvability. A mathematical introduction to string theory.
225
226
S. FERRY. A. RANICKI & J. ROSENBERG (eds)
Novikov conjectures. index theorems and rigidity I.
S. FERRY. A. RANICKI & J. ROSENBERG (eds)
Novikov conjectures. index theorems and rigidity II.
228
Ergodic theory of Zd actions.
230
Prolegomena to a middlebrow arithmetic of curves of genus 2.
231 232
233 234 235
M. POLLICOTT & K. SCHMIDT (eds)
Ergodicity for infinite dimensional systems.
G. DA PRATO & J. ZABCZYK
H. BECKER & A.S. KECHRIS
S. COHEN & H. NIEDERREITER (eds)
Finite fields and applications.
Number theory 1993-94.
J.W.S. CASSELS & E.V. FLYNN
K.H. HOFMANN & M.w. MISLOVE (eds)
Semigroup theory and its applications.
The descriptive set theory of Polish group actions. Introduction to subfactors.
KAZUAKI TAIRA
S.B. COOPER. T.A. SLAM AN & S.S. WAINER (eds) S. ALBEVERIO ,I af
227 229
A. PFISTER
PETER ROWLINSON (ed)
A. JOYAL & I. MOERDIJK
V. JONES & V.S. SUNDER
S. DAVID(ed)
H. FETTER & B. GAMBOA DE BUEN
236
The James forest.
237
Sieve methods. exponential sums. and their applications in number theory.
238 240
241
Stable groups.
FRANK O. WAGNER
Surveys in combinatorics. 1997.
R.A. BAILEY (ed)
L. SCHNEPS & P. LOCHAK (eds)
242
Geometric Galois actions I.
243
Geometric Galois actions II.
245
Geometry. combinatorial designs and related structures.
244
L. SCHNEPS & P. LOCHAK (eds)
Model theory of groups and automorphism groups.
246
p-Automorphisms of finite p-groups.
248
Tame topology and o-minimal structures.
247 249 250
251
252 253
254
255
256
257
258
259
Analytic number theory.
Characters and blocks or finite groups.
Gmbner bases and applications.
Geometry and cohomology in group theory. S. DONK1N
AJ. SCHOLL & R.L. TAYLOR (eds)
P.A. CLARKSON & F.w. NIJHOFF(eds)
HELMUT VOLKLEIN .1 af
Aspects of Galois theory.
An introduction to noncommutative differential geometry and its physical applications 2ed.
Sets and proofs.
S.B. COOPER & J. TRUSS (eds)
Models and computability.
Groups St Andrews 1997 in Bath. II.
C.M. CAMPBELL ,I af
C.M. CAMPBELL ,I af
C.W. HENSON. J. IOVINO. A.S. KECHRIS & E. ODELL
BILL BRUCE & DAVID MOND (eds)
Singularity theory.
265
Elliptic curves in cryptography.
New trends in algebraic geometry.
Surveys in combinatorics. 1999.
K. HULEK. F. CATANESE. C. PETERS & M. REID (eds)
I. BLAKE. G. SEROUSSI & N. SMART J.D. LAMB & D.A. PREECE (eds)
Spectral asymptotics in the semi·classical limit. Ergodic theory and topological dynamics.
M. D1MASSI & J. SJOSTRAND
M.B. BEKKA & M. MAYER
N.T. VAROPOULOS & S. MUSTAPHA
270
Analysis on Lie groups.
272 273
Character theory for the odd order theorem. T. PETERFALVI
271
274
Singular penurbations of differential operators.
Spectral theory and geometry.
S. ALBEVERIO & P. KURASOV
E.B. DAVIES & Y. SAFAROV (eds)
The Mandlebrot set. theme and variations.
TAN LEI (ed)
275
Descriptive set theory and dynamical systems.
277
Computational and geometric aspects of modem algebra.
276
278 279
280 281
282 283
284 285 286
287 288 289 290 291 294
J. MADORE
S.B. COOPER & J. TRUSS (eds)
263
267
P. KROPHOLLER. G. NIBLO. R. STOHR (eds)
Symmetries and integrability of difference equations.
Analysis and logic.
268
G. NAVARRO
Galois representations in arithmetic algebraic geometry.
262
269
LOU VAN DEN DRIES
ROBERT CURTIS & ROBERT WILSON (eds)
B. BUCHBERGER & F. WINKLER(eds)
Groups St Andrews 1997 in Bath. I.
264
J.W .P. HIRSCHFELD ,I af
Y. MOTOHASHI (ed)
260 261
D. EVANS (ed)
E.I. KHUKHRO
The atlas of finite groups: ten years on.
The q-Schur algebra.
G.RH. GREAVES .1 af
A. MARTSINKOVSKY & G. TODOROV (eds)
Representation theory and algebraic geometry.
Singularities of plane curves.
M. FOREMAN 'Illf
E. CASAS-ALVERO
Global amactors in abstract parabolic problems.
Topics in symbolic dynamics and applications.
M. D. ATKINSON ,I af
J.W. CHOLEWA & T. DLOTKO
F. BLANCHARD. A. MAASS & A. NOGUEIRA (eds)
Characters and automorphism groups of compact Riemann surfaces. THOMAS BREUER Explicit birational geometry of 3-folds. ALESSIO CORTI & MILES REID (eds)
Auslander-Buchweitz approximations of equivariant modules. M. HASHIMOTO Nonlinear elasticity. Y. FU & R. W. OGDEN (eds)
Foundations of computational mathematics. R. DEVORE. A. ISERLES & E. SOLI (eds) Rational points on curves over finite fields. H. NIEDERREITER & C. XING Clifford algebras and spinors 2ed. P. LOUNESTO
Topics on Riemann surfaces and Fuchsian groups. E. BUJALANCE. A.F. COSTA & E. MARTINEZ (eds) Surveys in combinatorics. 2001. J. HIRSCHFELD (ed) Aspects of Sobolev-type inequalities. L. SALOFF-COSTE Tits buildings and the model theory of groups. K. TENT (ed) A quantum groups primer. S. MAJID Introduction to operator space theory. G. P1SIER
London Mathematical Society Lecture Note Series. 294
Introduction to Operator Space Theory
Gilles Pisier Texas A&M University & University of Paris 6
CAMBRIDGE UNIVERSITY PRESS
PUBLISHED BY THE PRESS SYNDICATE OF THE UNIVERSITY OF CAMBRIDGE
The Pitt Building. Trumpington Street. Cambridge. United Kingdom CAMBRIDGE UNIVERSITY PRESS
The Edinburgh Building. Cambridge. CB2 2RU. UK 40 West 20th Street. New York. NY 10011-4211. USA 477 Williamstown Road. Port Melbourne. VIC 3207. Australia Ruiz de Alarc6n 13.28014 Madrid. Spain Dock House. The Waterfront. Cape Town 800 I. South Africa http://www.cambridge.org
© Gilles Pisier 2003 This book is in copyright. Subject to statutory exception and to the provisions of relevant collective licensing agreements. no reproduction of any part may take place without the written permission of Cambridge University Press. First published 2003 Typeface CMR 10112 pt.
System ~TEX2e [TBj
A catalog record for this book is available from the
Briti.~h
Library.
Library of COllgress Catalogillg ill Publicatioll Data
Pisier. Gilles. 1950Introduction to operator space theory I Gilles Pisier. p. cm. - (London Mathematical Society lecture note series; 294) Includes bibliographical references and index. ISBN 0-521-81165-1 (pbk.) I. Operator spaces. I. TItle. II. Series. QA322.2 .P545 2003 515'.732-dc21 2002031358 ISBN 0 521 81165 I paperback
Transferred to digital printing 2004
CONTENTS
O. Introduction 1. 2.
3. 4. 5.
6. 7.
8. 9.
Part I. INTRODUCTION TO OPERATOR SPACES Completely bounded maps The minimal tensor product. Ruan's theorem. Basic operations 2.1. IVIinimal tensor product 2.2. Ruan's theorem 2.3. Dual space 2.4. Quotient space Quotient by a subspace Quotient by an ideal 2.5. Bidual. Von Neumann algebras 2.6. Direct sum 2.7. Intersection, sum, complex interpolation 2.8. Ultraproduct 2.9. Complex conjugate 2.10. Opposite 2.11. Ruan's theorem and quantization 2.12. Universal objects 2.13. Perturbation lemmas Minimal and maximal operator space structures Projective tensor product The Haagerup tensor product Basic properties l\'hlltilinear factorization Injectivity /projectivity Self-duality Free products Factorization through R or C Symmetrized Haagerup tensor product Complex interpolation Characterizations of operator algebras The operator Hilbert space Hilbertian operator spaces Existence and unicity of OH. Basic properties Finite-dimensional estimates Complex interpolation Vector-valued Lp-spaces, either commutative or noncommutative Group C*-algebras. Universal algebras and unitization for an operator space Examples and comments 9.1. A concrete quotient: Hankel matrices
1
17
28 28 34 40 42 42 43 47 51 52 59 63 64 65 67 68 71
81 86 86 92 93 93 98 101 102 106 109 122 122 122 130 135 138 148 165 165
Contents
vi 9.2. 9.3. 9.4. 9.5. 9.6. 9.7.
Homogeneous operator spaces Fermions. Antisymmetric Fock space. Spin systems The Cuntz algebra On The operator space structure of the classical Lp-sp~ces The C·-algebra of the free group with n generators Reduced C·-algebra of the free group with n generators 9.8. Operator space generated in the usual Lp-space by Gaussain random variables or by the Rademacher functions 9.9. Semi-circular systems in Voiculescu's sense 9.lD. Embeddings of von Neumann algebras into ultraproducts 9.11. Dvoretzky's theorem lD. Comparisons
172 173 175
178 182 183
191 200 210 215 217
Part II. OPERATOR SPACES AND C*-TENSOR PRODUCTS 11. 12. 13. 14. 15. 16. 17. 18.
19. 20. 21. 22. 23. 24.
25. 26. 27. 28.
C*-norms on tensor products. Decomposable maps. Nuclearity
Nuclearity and approximation properties C*(lFoo ) Q9 B(H) Kirchberg's theorem on decomposable maps The Weak Expectation Property (WEP) The Local Lifting Property (LLP) Exactness Local reflexivity Basic properties A conjecture on local reflexivity and OLLP Properties C, C f , and C". Exactness versus local reflexivity Grothendieck's theorem for operator spaces Estimating the norms of sums of unitaries: Ramanujan graphs, property T, random matrices Local theory of operator spaces. Nonseparability of OSn B(H) Q9 B(H) Completely isomorphic C* -algebras Injective and projective operator spaces Part III. OPERATOR SPACES AND NON-SELF-ADJOINT OPERATOR ALGEBRAS Maximal tensor products and free products of operator algebras The Blecher-Paulsen factorization. Infinite Haagerup tensor products Similarity problems The Sz.-Nagy-Halmos similarity problem
Solutions to the exercises
227 240 252 261 267 275 285 303 303 305 309 316 324 334 348 354 356
365 384 396 407 418
Contents References Subject index Notation index
vii
457 477
479
Chapter o. Introduction The theory of operator spaces is very recent. It was developed after Ruan's thesis (1988) by Effros and Ruan and Blecher and Paulsen. It can be described as a noncommutative Banach space theory. An operator space is simply a Banach space given together with an isometric linear embedding into the space B(H) of all bounded operators on a Hilbert space H. In this new category, the objects remain Banach spaces but the morphisms become the completely bounded maps (instead of the bounded linear ones). The latter appeared in the early 1980s following Stinespring's pioneering work (1955) and Arveson's fundamental results (1969) on completely positive maps. We study completely bounded (in short c.b.) maps in Chapter 1. This notion became important in the early 1980s through the independent work of Witt stock [Witl-2], Haagerup [H4], and Paulsen [Pa2]. These authors independently discovered, within a short time interval, the fundamental factorization and extension property of c.b. maps (see Theorem 1.6). For the reader who might wonder why c.b. maps are the "right" morphisms for the category of operator spaces, here are two arguments that come to mind: Consider EI C B(Ht} and E2 C B(H2) and let rr: B(Ht} -) B(H2) be a C*-morphism (Le. a *-homomorphism) such that rr(E 1 ) C E 2 • Then, quite convincingly, u = rrlEl: EI -) E2 should be an "admissible" morphism in the category of operator spaces. Let us call these morphisms of the "first kind." On the other hand, if a linear map u: EI -) E2 is of the form u(x) = VxW with V E B(H1 , H 2) and W E B(H2, HI), then again such an innocentlooking map should be "admissible" and we consider it to be of the "second kind." But precisely, the factorization theorem of c.b. maps says that any c.b. map u: EI -) E2 between operator spaces can be written as a composition EI ~E3~E2 with UI of the first kind and U2 of the second. This is one argument in support of c.b. maps. Another justification goes via the minimal tensor product: If EI C B(Hd and E2 C B(H2) are operator spaces, their minimal tensor product EI ®minE2 is defined as the completion of their algebraic tensor product (denoted by EI ®E2) with respect to the norm induced on EI ®E2 by the space B(HI ®2H2) of all bounded operators on the Hilbertian tensor product HI ®2 H2 (this norm coincides with the minimal C*-norm when EI and E2 are C*-algebras). Moreover, the isometric embedding EI ®min E2 C B(HI ®2 H 2)
turns EI ®min E2 into an operator space. The minimal tensor product is discussed in more detail in §2.1. It is but a natural extension of the "spatial" tensor product of C*-algebras. In some sense, the minimal tensor product is the most natural operation that is defined using the "operator space structures" of El and E2 (and not only their norms). This brings us to the second
2
Introduction to Operator Space Theory
argument supporting the assertion that c.b. maps are the "right" morphisms. Indeed, one can show that a linear map u: El --+ E2 is c.b. if and only if (iff) for any operator space F the mapping IF ® u: F ®min El --+ F ®min E2 is bounded in the usual sense. Moreover, the c. b. norm of u could be equivalently defined as lIulicb = sup IIIF ® ull, where the supremum runs over all possible operator spaces F. (Similarly, u is a complete isometry iff IF ® u is an isometry for aU F). AI:, an immediate consequence, if v: Fl --+ F2 is another c.b. map between operator spaces, then v ® u: Fl ®min El --+ F2 ®min E2 also is c.b. and (0.1) In conclusion, the c. b. maps are precisely the largest possible class of morphisms for which the minimal tensor product satisfies the "tensorial" property (0.1). So, if one agrees that the minimal tensor product is natural, then one should recognize c. b. maps as the right morphisms. While the notion of c.b. map (which dates back to the early 1980s, if not sooner) is fundamental to this theory, this new field really took off around 1987 with the thesis of Z. J. Ruan [RulJ, who gave an "abstract characterization" of operator spaces (described in §2.2). Roughly, his result provides a "quantized" counterpart to the norm of a Banach space. When E is an operator space, the norm has to be replaced by the sequence of norms (II lin) on the spaces Mn (E) ~ Mn ®min E of all n x n matrices with entries in E. (The usual norm corresponds to the case n = 1.) In this text, we prefer to replace this sequence of norms by a single one, namely, the norm on the space K®min E with K = K (£2) (= compact operators on £2). Since K = U.Mn, it is of course very easy to pass from one viewpoint to the other. The main advantage of Ruan's Theorem is that it allows one to manipulate operator spaces independently of the choice of a "concrete" embedding into B(H). In particular, Ruan's Theorem leads to natural definitions for the dual E* of an operator space E (independently introduced in [ER2, BPI]) and for the quotient Ed E2 of an operator space El by a subspace E2 c El (introduced in [Rul]). These notions are explained in §§2.3 and 2.4. It should be emphasized that they respect the underlying Banach spaces: The dual operator space E* is the dual Banach space equipped with an additional (specific) operator space structure (Le., for some 11. we have an isometric embedding E* C B(11.» and similarly for the quotient space. In addition, the general rules of the duality of Banach spaces (for example, the duality between subspaces and quotients) are preserved in this "new" duality. More operations can be defined following the same basic idea: complex interpolation (see §2.7) and ultraproducts (see §2.8). We will also use some
o.
Introduction
3
more elementary constructions, such as direct sums (§2.6), complex conjugates (§2.9), and opposites (§2.IO). Although we described Ruan's thesis &<; the starting point of the theory, there are many important "prenatal" contributions that shared this new area. Among them, Christensen and Sinclair's factorization of multilinear maps [CSI] stands out (with its extension to operator spaces by Paulsen and Smith [PaS]). Going back further, there is an important paper by Effros and Haagerup [EH], who discovered that operator spaces may fail the local reflexivity principle, a very interesting phenomenon that is in striking contrast with the Banach space case (their results were inspired by Archbold and Batty's results [AB] for C* -algebras). Closely connected to theirs, Kirchberg's work on "exact" C* -algebras (see [Kil, \Va2]) has also been very influential Given a normed space E, there are of course many different ways to embed it into B(H). Two embeddings ]1: E ----> B(Hl ) and ]2: E ----> B(H2 ) are considered equivalent if the associated norms on lIfn(E) (or on IC ® E) are the same for all n ~ 1. By an operator space structure on E (compatible with the norm) what is meant usually is the data of an equivalence class of such isometric embeddings]: E ----> B(H) for this equivalence relation. Blecher and Paulsen [BPI] observed that the set of all operator space structures admissible on a given normed space E has a minimal and a maximal element, which they denoted by min(E) and ma..x(E). We summarize their results and various related open questions in Chapter 3. The minimal tensor product naturally appears as the analog for operator spaces of the "injective" tensor product of Banach spaces. Thus, both Effros-Ruan [ER6-8] and Blecher and Paulsen [BPI] were led to study the operator space analog of the "projective" tensor product in Grothendieck's sense. Recall that a mapping u: E ----> F (between Banach spaces) is nuclear iff it admits a factorization of the form
where Q, (3, ~ are bounded mappings and ~ is diagonal with coefficients (~71) in fl. M~reover, the nuclear norm N(u) is defined as N(u) = inf{lIall2: 1~71II1P1I}, where the infimum rllns over all possible factorizations. As is well known, the space IC = K(£2) of all compact operators on £2 is the noncommutative analog of Co, while the space 8 1 of aU trace class operators on £2 (with the norm IIxlis t = tr(lxl) is the noncommutative analog of £1. Now, if E, Fare operator spaces, a mapping u: E ----> F is called "nuclear in the o.s. sense" (introduced in [ER6]) if it admits a factorization of the form '" Ll. {3 E ----> IC ----> 8 1 ----> F,
Introduction to Operator Space, Tlleory
4
where a, f3 are c.b. maps and 6: /C
-+
8 1 is of the form
6(x) = axb with a, b Hilbert-Schmidt. Then the os-nuclear norm is defined as
(Here, of course,
lIall2' IIbll 2denote the Hilbert-Schmidt norms.)
We describe some of the developments of these notions in Chapter 4. In Banach space theory, Grothendieck's approximation property has played an important role. Recall that Enflo [En] gave the first counterexample in 1972 and Szankowski [Sz] proved around 1980 that the space B(£2) of all bounded operators on £2 fails the approximation property. Quite naturally, this notion has an operator space counterpart. In the Banach space case, Grothendieck proved that a space E has the approximation property iff the natural morphism 1\
V
E* ® E -+ E* ® E
from the projective to the injective tensor product is one to one. We describe in Chapter 4 Effros and Ruan's operator space version of this result. In Chapter 5, we introduce the Haagerup tensor product E1 ®h E2 of two operator spaces E 1 , E 2 • This notion is of paramount importance in this young theory, and we present it from a somewhat new viewpoint. We prove that if E 1, E2 are subspaces of two unital C*-algebras A 1, A 2, respectively, then E1 ®h E2 is naturally embedded (completely isometrically) into the (C*algebraic) free product A1 * A2 ([CES]). We also prove the factorization of completely bounded multilinear maps due to Christensen and Sinclair [CS1] (and to Paulsen and Smith [PaS] for operator spaces) and mimy more impor- ' taut properties like the self-duality, the shuffle theorem (inspired by [ERlO, EKR]), or the embedding of E1 ®h E2 into the space of maps factoring through the row or column Hilbert space ([ER4]). We also include a brief study of the symmetrized Haagerup tensor product recently introduced in lOP], and we describe the "commutation" between complex interpolation and the Haagerup tensor product ([Ko,P1]). As an application, we prove in Chapter 6 a characterization of operator algebras due to Blecher-Ruan-Sinclair ([BRS]). The question they answer can be explained as follows: Consider a unital Banach algebra A with a normalized unit and admitting also an operator space structure (Le., we have A c B(H) as a closed linear subspace). When can A be embedded into B(H) as a closed unital subalgebra without changing the operator space structure? They prove that a necessary and sufficient condition is that the product mapping
o. Introduction
5
defines a completely contractive map from A Q9h A into A. The isomorphic (as opposed to isometric) version of this result was later given by Blecher ([B4]). We include new proofs for these results based on the fact (due to ColeLumer-Bernard) that the class of operator algebras (i.e. closed subalgebras of B(H» is stable under quotients by closed ideals. We also give an analogous characterization of operator mod ules, following [CES]. Curiously, the simplest of all Banach spaces, namely, the Hilbert space £2, can be realized in many different ways as an operator space. Theoretical physics provides numerous examples of the sort, several of which are described in Chapter 9. Nevertheless, there exists a particular operator space, which we denote by OH, that plays exactly the same central role for operator spaces as the space £2 among Banach spaces. This space OH is characterized by the property of being canonically completely isometric to its antidual; it also satisfies some remarkable properties with regard to complex interpolation. The space OH is the subject of Chapter 7 (mainly based on [PI]). Since this space gives a nice operator space analog of £2 or L 2 , it is natural to investigate the case of £p or Lp for p =I- 2, as well as the case of a noncommutative vector valued Lp- \Ve do this at the end of Chapter 7, and we return to this in several sections in Chapter 9. However, on that particular topic, we should warn the reader of a certain paradoxical bias: If we do not give to this subject the space it deserves, the sole reason is that we have written an extensive monograph [P2] entirely devoted to it, and we find it easier to refer the reader to the latter for further information. In Chapter 8 we introduce the group C* -algebras (full and reduced) and the universal C* -algebra C* (E) of an operator space E, as well as its universal operator algebra (resp. unital operator algebra) OA(E) (resp. OA,,(E». Every theory, even one as young as this, displays a collection of "classical" examples that are constantly in the back of the mind of researchers in the field. Our aim in Chapter 9 is to present a preliminary list of such examples for operator spaces. l\lost of the classical examples of C* -algebras possess a natural generating subset. Almost always the linear span of this subset gives rise to an interesting example of operator space. The discovery that this generating operator space (possibly finite-dimensional) carries a lot of information on the C* -algebra that it generates has been one of the arguments supporting operator space theory. In Chapter 9 we make a special effort (directed toward the uninit.iated reader) to illustrate the theory with numerous concrete "classical" examples of this type, appearing in various areas of analysis, such as Hankel operators, Fock spaces, and Clifford matrices. Moreover, we describe the linear span of the free unitary generators in the "full" C*-algehra of the free group (§9.6) as well as in the "reduced" one (§9.7). We emphasize throughout §9 the class of homogeneous Hilbertian operator spaces, and we describe the span of independent Gaussian random variables (or the Rademacher functions)
6
Introduction to Operator Space Theory
in L p , in the operator space framework (§9.8). Our treatment underlines the similarity between the latter space and its analog in Voiculescu's free probability theory (see §9.9). Indeed, it is rather curious that for each 1 ::; p < 00 the linear span in Lp of a sequence of independent standard Gaussian variables is completely isomorphic to the span in noncommutative Lp of a free semi-circular (or circular) sequence in Voiculescu's sense (see Theorem 9.9.7). Thus, if we work in Lp with 1 ::; p < 00, the operator space structure seems to be roughly the same in the "independent" case and in the "free" one, which is rather surprising. Our description in §9.8 of the operator space spanned in Lp by Gaussian variables (or the Rademacher functions) is merely a reinterpretation of the noncommutative Khintchine inequalities due to F. Lust-Piquard and the author (see [LuP, LPPJ). These inequalities also apply to "free unitaries" (see Theorem 9.8.7) or to "free circular" variables (see Theorem 9.9.7). In view of the usefulness and importance of the classical Khintchine inequalities in commutative harmonic analysis, it is natural to believe that their noncommutative (Le., operator space theoretic) analog will play an important role in noncommutative Lp-space theory. This is why we have devoted a significant amount of space to this topic in §§9.8 and 9.9. Moreover, in §9.10, we relate these topics to random matrices by showing that the von Neumann algebra of the free group embeds into a (von Neumann sense) ultraproduct of matrix algebras. One can do this by using either the residual finiteness of the free group (as in [Wa1J) or by using one of Voiculescu's matrix models involving independent Gaussian random matrices suitably normalized, and Paul Levy's concentration of measure phenomenon (see [MSJ). Finally, in §9.11, we discuss the possible analogs of Dvoretzky's Theorem for operator spaces (following [P9J). In Chapter 10 we compare the various examples reviewed in Chapter 9, and we show (by rather elementary arguments) that, except for the few isomorphisms encountered in Chapter 9, these operator spaces are all distinct. This new theory can already claim some applications to C* -algebras, many of which are described in the second part of this book. For instance, the existence of an "exotic" C*-algebra norm on B(H) Q9 B(H) was established in [JP] (see Chapter 22). Moreover, this new ideology allows us to "transfer" into the field of operator algebras several techniques from the "local" (Le. finite-dimensional) theory of Banach spaces (see Chapter 21). The main applications so far have been to C* -algebra tensor products. Chapters 11 to 22 are devoted to this topic. We review in Chapters 11 and 12 the basic facts on C* -norms and nuclear C* -algebras. Since we are interested in linear spaces (rather than cones) of mappings, we strongly emphasize the "decomposable maps" between two C*-algebras (Le., those that can be decomposed as a linear combination of (necesssarily at most four) completely positive maps) rather than the completely positive (in short c.p.) ones themselves. Our treatment
o.
I11troductio11
7
owes much to Haagerup's landmark paper [HI]. For a more traditional one emphasizing c.p. maps, see [Pal]. Recall that, if A, Bare C* -algebras, there is a smallest and a largest C*norm on A0B and the resulting tensor products are denoted by A0minB and A 0 max B. (This notation is coherent with the previous one for the minimal tensor product of operator spaces.) Moreover, a C* -algebra A is called nuclear if A 0 m in B = A 0 max B for any C* -algebra B. In analogy with the Banach space case explored by Grothendieck, the maximal tensor product is projective but not injective, and the minimal one is injective but not projective. Therefore we are naturally led to distinguish two classes: first, the class of C* -algebras A for which the "functor" B
->
A 0 max B is injective
(this means that B c C implies A 0 max B class of C* -algebras A for which the "functor" B
->
c A
0 max C), and, second, the
A 0 m in B is projective
(this means that B = C/I implies A 0 m in B = (A 0 m in C)/(A 0 m in I)). The first class is that of nuclear C* -algebras reviewed in Chapter 12 (see Exercise 15.2), and the second one is that of exact C* -algebras studied in Chapter 17. In Chapter 12 we first give a somewhat new treatment of the well-known equivalences between nuclearity and several forms of approximation properties. We also apply our approach to multilinear maps into a nuclear C* -algebra (see Theorem 12.11) in analogy with Sinclair's and Smith's recent work [SS3] on injective von Neumann algebras. Then, Chapter 13 is devoted to a proof of Kirchberg's Theorem, which says that there is a unique C*-norm on the tensor product of B(H) with the full C* -algebra of a free group. The next chapter, Chapter 14, is devoted to an unpublished result of Kirchberg showing that the decomposable maps (Le. linear combinations of completely positive maps) are the natural morphisms to use if one replaces the minimal tensor product of C* -algebras by the maximal one. The same proof actually gives a necessary and sufficient condition for a map defined only on a subspace of a C* -algebra, with range another C* -algebra, to admit a decomposable extension. The next two chapters are closely linked together. In Chapters 15 and 16 we present respectively the "weak expectation property" (WEP) and the "local lifting property" (LLP). We start with the C*-algebra case in connection with Kirchberg's results from Chapter 13, and then we go on to the generalizations to operator spaces. In particular, we study Ozawa's OLLP
8
Introduction to Operator Space Theory
from [Oz3J. At the end of Chapter 16 we discuss at length several. equivalent reformulations of Kirchberg's fundamental conjecture on the umqueness of the C*-norm on C*(lFoo ) ® C*(lFoo ). For instance, it is the same as asking whether LLP implies WEP. In Chapter 17 we concentrate on the notion of "exactness" for either operator spaces or C* -algebras. Assume that A embedded into B(H) as a C* -subalgebra. Then A is exact iff A®min B embeds isometrically into B(H) ®max B for any B. Equivalently, this means that the norm induced on A ® B by B(H) ®max B coincides with the min-norm. This is not the traditional definition of exactness, but it is equivalent to it (see Theorem 17.1). The traditional one is in terms of the exactness of the functor B -+ A ®min B in the C"-category (see (17.1», and for operator spaces there is also a more appealing reformulation in terms of ultraproducts: Exact operator spaces X are those for which the operation Y -+ Y ®rnin X essentially commutes with ultraproducts (see Theorem 17.7). The concept of "exactness" owes a lot to Kirchberg's fundamental contributions [Kit-3]. In particular, Kirchberg proved recently the remarkable definitive result that every separable exact C* -algebra embeds (as a C* -subalgebra) into a nuclear one. However, in the operator space framework, the situation is not as clear. When X, Yare operator spaces, we will say that X "locally embeds" into Y if there is a constant C such that, for any finite-dimensional subspace E eX, there is a subspace E c Y and an isomorphism u: E -+ E with Ilullcbllu-1llcb ::; C. We denote (0.2) dsy(X) = inf{C}, that is, dsy(X) is the smallest constant C for which this holds. With this terminology, an operator space X is exact iff it locally embeds into a nuclear C* -algebra B. Actually, for such a local embedding we can always take simply B = K (and E can be a subspace of Aln with n large enough). It is natural to introduce (see Chapter 17) the "constant of exactness" ex(E) of an operator space E (and we will prove in Chapter 17 that it coincides with the just defined constant dsx::.{E». This is of particular interest in the finite-dimensional case, and, while many of Kirchberg's results extend to the operator space case, many interesting questions arise concerning the asymptotic growth of these constants for specific E when the dimension of E tends to infinity. (See, e.g., Theorems 21.3 and 21.4.) In Chapter 18 we describe the main known facts concerning "local reflexivity." While every Banach space is "locally reflexive" (cf. [LiR]), it is not so in the operator space category, and this raises all sorts of interesting questions.
O. Introduction
9
For instance, Kirchberg [Kil] proved that, for C* -algebras, exactness implies local reflexivity, but the converse remains open. For operator spaces, we will see, following [EOR], that I-exact implies I-locally reflexive (see Theorem 18.21), but the converse is now obviously false since there are reflexive nonexact spaces. We will also show that any "noncommutative L 1-space" (Le., any predual of a von Neumann algebra) is locally reflexive ([EJR]); we follow the simpler approach of [JLM]. We will describe the properties C, C', and C", which are at the origin of the study of local reflexivity for C* -algebras ([AB, EH]). We also return to the OLLP from Chapter 16. We show that the latter for X** implies local reflexivity for X. Moreover, we discuss several interesting conjectures from [Oz3, Oz6] that seem closely related t.o the old quest.ion of whether an ideal in a separable C* -algebra is automatically the range of a bounded linear projection. For instance, it is a very interesting open question of whether the space B(£2) equipped with its maximal operator space stl'llCture (in the sense of Chapter 3) is locally reflexive (see also [03] for related problems). In another direction, the operator space version of the approximation property (called the OAP, see Definitions 17.11) seems, for C* -algebras at least, closely related to exactness via the so-called slice map properties (see Corollary 17.14 and the remark below Remark 17.17). In Chapter 19, we present a version of Grothendieck's factorizat.ion t.heorem adapted to operator spaces, following [JP]. See [PiS] for a different version obtained very recently. In the Banach space context, Grothendieck's theorem implies (see [P4]) that every bounded map u: LOO ---+ L1 factors through L2. Moreover (see [P4, Chapt.er 9]) the same is true for any bounded map u: A ---+ B* when A and B are arbitrary C* -algebras. \Ve prove (see Corollary 19.2) that if E c A and FeB are exact operator spaces, then any c.b. map u: E ---+ F* factors through a Hilbert space and the corresponding bilinear form on E x F extends to a bounded bilinear form on A x B. In Chapter 21, we prove that for n > 2 the metric space OSn of all ndimensional operator spaces is not separable for its natural metric, in sharp contrast to the Banach space analog. We give two approaches to this key result, one based on the factorization from Chapter 19 and one based on a specific constant C(n) studied in Chapter 20. This constant quantifies a certain asymptotic phenomenon for n-tuples of unitary N x N matrices when the size N tends to infinity. The proof that C(n) < n involves surprisingly deep ingredients (Property T, expanders, random matrices), which are described in Chapter 20. In answer to a question of Kirchberg, it was proved in [JP] that B(H) 0rnin B(H) i- B(H) 0 max B(H). This is described in Chapter 22. An important role is played behind the scene by the full C* -algebra of the free group IF 00 on infinitely many generators. This C*-algebra is denoted by C*(IFoo). The
10
Introduction to Operator Space Theory
ideas involved emphasize the importance of a subclass among operator spaces, namely, those spaces E such that any of their finite-dimensional subspaces embed almost completely isometrically into CoO (F00). These are the spaces satisfying dSCO(Foo)(E) = 1 in notation (0.2). The constant dsc*Wx)(E) is abbreviated to dJ(E) in Chapter 22. In the finite-dimensional case, these spaces form a separable subclass, for a natural metric, of the class (itself nonseparable) of all finite-dimensional operator spaces. Moreover, this subclass is stable under duality and various tensor products. It turns out that many of the questions that have been examined for the CoO -algebra K (of all compact operators on £2) in connection with exactness have interesting analogs for the CoO -algebra CoO (F00) (see Chapter 22). Given the interplay between CoO -algebras and operator spaces, it is natural to ask: If two CoO-algebras At and A2 are isomorphic as operator spaces (Le., completely isomorphic), are they isomorphic as CoO -algebras? The answer is negative. However, it turns out that At and A2 must share numerous CoOproperties such as nuclearity, exactness, WEP, and injectivity (for von Neumann algebras). These questions are discussed in Chapter 23. In Chapter 24 (mainly a survey) we study injective and projective operator spaces. In the third part of the book, we concentrate on non-self-adjoint operator algebras. The typical examples are algebras of bounded analytic functions on some domain, the sub algebra of B(C2) formed of all triangular matrices (and more generally the so-called nest algebras [Dal]), or the unital algebra generated by a single operator T in B(C 2 ). Their. behavior is usually quite different from that of CoO -algebras. In Chapter 25 we return to the study of the maximal tensor product, already considered mainly for CoO-algebras in Chapters 11 and 12. Here we study more generally the maximal tensor product At ®max A 2 , when At and A2 are two unital operator algebras. This was first investigated in [PaP]. We discuss the analogue of nuclearity for unital operator algebras and various related questions. Our results are closely related to Haagerup's results [HI] on the relation between the decomposability properties of c.b. maps into a CoO-algebra A and the nuclearity of A, but our approach seems new. A bit surprisingly, it turns out that several basic facts remain valid for non-selfadjoint operator algebras. For instance, a unital operator algebra A satisfies B ®min A = B ®max A (isomorphically) for any unital operator algebra B iff the identity on A is approximable by finite-rank decomposable maps in a suitable way (see Theorem 25.9). Since we allow isomorphism (and not only isometry) in B®minA = B ®max A, there are clearly non-self-adjoint examples satisfying this (for instance, finite-dimensional quotients of the disc algebra, as discussed in Example 25.6); however, in the isometric case, we prove that only self-adjoint algebras can satisfy this (see Theorem 25.11). Further results are given in [LeM4].
O. Introduction
11
In Chapter 26 we present a striking factorization theorem due to Blecher and Paulsen [BP2]. We enlarge their framework slightly: First we apply their idea to study the operator algebras generated by an operator space, and then we reformulate their results in terms of infinite Haagerup tensor products. These results put the spotlight on a subclass of operator algebras that we call "full" (see Definition 26.3); these are the algebras A such that any unital contractive homomorphism from A to any B(H) is automatically completely contractive. The term "maximal" would perhaps be more appropriate but would create a serious risk of confusion with the maximal operator spaces of Chapter 3. We show, for instance (see Corollary 26.13). that the disc algebra and the bidisc algebra (which are both full) are completely isometric to a quotient of an infinite Haagerup tensor product of copies of the two-dimensional It -space (equipped with its natural operator space structure, which is the maximal one). Actually, the same result holds for any C* -algebra or for any full unital operator algebra. l\:Ioreover, the free product of two unital operator algebras Al * A2 is completely isometric to a quotient of an infinit.e Haagerup tensor product of the form Al ®h A2 ®h Al ®h A 2 ··· (see Corollary 26.16). These statements are closely related to those of the next chapter, where we give general conditions under which analogous results can hold for finite Haagerup tensor products instead of infinite ones. In Chapter 27 we study the Kadison similarity problem and the equivalent derivation problem. This leads us to consider more generally the class of operator algebras satisfying the following property, which we call (SP): Any bounded unital homomorphism u from A to B(H) (H arbitrary) is automatically completely bounded. We describe our recent work [P17] showing that an operator algebra A satisfies this iff all the matrices over A admit a specific factorization into a product of terms with a bounded length. This leads to a new invariant for an operator algebra A, its "length," denoted by leA), which also happens to be equal to its "similarity degree," denoted by d(A). There are (non-self-adjoint) examples showing that any integer can occur as the length of some A. In Chapter 28, we discuss the Sz.-Nagy-Halmos similarity problem: When is an operator in B(H) similar to a contraction'? We describe the recently found example of a polynomially bounded operator that is not similar to a contraction. This uses Hankel operators and the completely bounded ideology. Related interpretations for operator spaces are given. About the exercises. The exercises are really meant as complements to the text, although some of them can serve as ordinary exercises. The motivation behind this choice of format is mostly to avoid interrupting the flow of the presentation by technical details, which we prefer to leave as "exercises for
12
Introduction to Operator Space Theory
the reader." In this way, we hope to make the contours of '~the big picture" more visible. Note, however, that since we provide full solutions for all of the exercises at the end of this book, this practice does not reduce in any way the "self-containedness" of the material! We should apologize for sometimes omitting precise references either when the source for an exercise is not clear to us or when it can be easily guessed from the main text. Occasionally, an exercise may contain seemingly new results (for instance Exercises 5.8,5.9 and 5.10). A _bibliographical note. There are many books on (self-adjoint) operator algebras, mainly on C* and von Neumann algebras, starting from the great alltime classics [Dil, Di2]; going on to [Sa], [Ar2], [Ta3], [Ped], [StZ], [Sun], and [KaR]; and continuing with the more recent texts [Ky], [Wa2], [Da2], [Fi], and also [Co2] (on noncommutative geometry), [Bla2, WeO, RLL] (on J(-theory), [VDN, HiP] (on free probability) [GHJ, JS, Jon, P03, EvK] (on subfactors), and, recently, [RS] (on the classification program and entropy). For non-self-adjoint operator algebras, we only see [Ar3, DaI]' and also to some extent [Ni] and [SNF]. D. Blecher and C. Le IVlerdy are currently preparing a new book on that subject. However, there are still very few books on completely bounded maps or on operator spaces. Not long ago, the only ones on c.b. maps were [Pal] (a new edition of this one is to appear soon) and [PlO] (see also the more specialized texts [PI, P2]), but we can now refer the reader to the very recently published book [ERIl], which presents the fundamentals of operator space theory. While [ERll] has some overlap with the present volume, the two books are actually quite complementary in both style and content.
Basic notation. Let X, Y be Banach spaces. We denote by Bx the closed unit ball of X and by B(X, Y) the space of all bounded maps from X to Y equipped with its usual norm. When X = Y, we set B(X) = B(X, X). For any set I, we denote by £2(I) the space of square summable complex scalars with the usual norm x 1-+ (2: IXiI2)1/2. When I = N (resp. I = [1, ... , n]), we denote this space by £2 (resp. £~). The space B(£~) can be identified in the usual way with the space of all n x n complex matrices. We will denote the latter space by Mn (or Mn(C» and we equip it with the norm of B(e~). Depending on the context, we will denote by eij the element of Aln , or of B(£2), corresponding to the matrix with all entries equal to zero except the one at the i-th row and the j-th column which is equal to 1. Acknowledgments. This book grew out of lecture notes written first for the French Mathematical Society in 1994 and gradually expanded after that. The
O. Introduction
13
existing manuscript served as notes for advanced graduate courses given at Texas A&lVI and Paris VI at several occasions during the period 1996-2001. I am very grateful to all those who made critical comments and pointed out numerous errors and misprints, particularly, A. Arias, G. Aubrun, P. Biaue, D. Blecher, 1\'1. Junge, C. Le Merdy, A. Nou, N. Ozawa, G. Radler, E. Ricard, and Z.J. Ruan. I am especially grateful to Sophie Grivaux and Xiang Fang for their careful reading of a nearly final version, \vhich allowed many correetions. The support of NSF and Texas Advanced Research Program 010366-163 is gratefully acknowledged. Finally, one more time, my warmest thanks go to Robin Campbell, whose incredible efficiency in typing this book has been extremely helpful.
PART I INTRODUCTION TO OPERATOR SPACES
Chapter 1. Completely Bounded Maps Let us start by recalling the definition and a few facts on C* -algebras: Definition 1.1. A C* -algebra is a Banach *-algebra, satisfying the identity
Ilx*xll = IIxl1 2 for any element x in tIle algebra. The simplest example is the space
B(H) of all bounded operators on a Hilbert space H, equipped with the operator norm. ]\!ore 'generally, any closed subspace
A
c B(H)
stable under product and involution is a C* -algebra. By classical results (Gelfand and Naimark) we know that every C*-algebra can be realized as a closed self-adjoint subalgebra of B(H). Moreover, we also know that every commutative unital C* -algebra can be identified with the space C(T) of all complex-valued continuous functions f: T ----> C on some compact space T. If A has no unit, A can be identified with the space Co(T) of aU complex-valued continuous functions, vanishing at infinity, on some locally compact space T. Of course the object of C* -algebra theory (as developed in the last 50 years; cf. [KaR, Ta3]) is the classification of C* -algebras. Similarly, the object of Banach space theory is the classification of Banach spaces. In the last 25 years, it is their classification up to isomorphism (and NOT up to isometry) that has largely predominated (cf., e.g., [LTI-3, P4]). This already indicates one major difference between these two fields since, if Al and A2 are two C* -algebras,
Al isomorphic to A2
=}
Al isometric to A 2.
In particular, a C* -algebra admits a unique C* -norm. So there is no "isomorphic theory" of C* -algebras. However, in recent years, operator algebraists have found the need to relax the structure of C* -algebras and consider more general objects called operator systems. These are subspaces of B(H) containing the unit that are stable under the involution but not under the product. The theory of operator systems was developed using the order structure repeatedly, and it is still mostly an isometric theory. The natural morphisllls here are the "completely positive" maps (cf. [St, Arl]). We refer the reader to a survey by Effros [EI] and a series of papers by Choi and Effros (especially [CE3]). Even more recently, operator algebraists have done a radical simplification and considered just "operator spaces":
Introduction to Operator Space Theory
18
Definition 1.2. An operator space is a closed subspace of B(H). Equivalently, since we can think of C* -algebras as closed self-adjoint subalgebras of B(H), we can think of operator spaces as closed subspaces of C· -algebras. Operator space theory can be considered as a merger of C* -algebra theory and Banach space theory. It is important to immediately observe that any Banach space can appear as a closed subspace of a C*-algebra. Indeed, for any Banach space X (with the dual unit ball denoted by B x ')' if we let T
= (Bx.,a(X*,X)),
then T is compact and we have an isometric embedding Xc C(T).
Hence, since C(T) is a C*-algebra (and C(T) c B(H) with H = e2 (T)), X also appears among operator spaces. So operator spaces are just ordinary Banach spaces X but equipped with an extra structure in the form of an embedding Xc B(H). The main difference between the category of Banach spaces and that of operator spaces lies not in the spaces but in the morphisms. We need morphisms that somehow keep track of the extra information contained in the data of the embedding X C B(H); the maps that do just this are the completely bounded. maps.
Definition 1.3. Let E C B(H) and F C B(l() be operator spaces and consider a map B(l() B(H)
For any n
~
U
U
E
F
1, let
be the space of n x n matrices with entries in E. In particular, we lJave a natural identification Mn(B(H)) ~ B(e 2(H)), where
e2(H)
means
!f EB H
a fortiori its subspace
EB··· EB H,. Thus, we may equip Mn(B(H)) and
'" n times
1. Completely Bounded Afaps
19
with the norm induced by B(€'2(H)).
Then, for any n map
~
1, the linear map u: E
->
F allows us to define a linear
defined by
A map
'11:
E
->
F is called completely bounded (in s}lOrt c. b.) if sup n~l
Ilun II Al" (E)--+Al" (F) < 00.
We define
and we denote by
CB(E,F) the Banach space of all c. b. maps from E into F equipped wit}l tlle c. b. norm. This space will replace the space B(E, F) of all bounded operators from E into F. (\Ve will see later on that it can be equipped with an operator space structure. ) If G c B(L) is another operator space and if v: F -> G is c.b., then the compositon vu: E -> G clearly remains c.b. and we have
Of comse, when n = 1, 1 x 1 matrices are just elements of E, so M 1 (F) is nothing but u itself. In particular we have
Ilull
~
U1:
All (E)
->
IIuli c b
and
CB(E, F) c B(E, F). When Ilulicb ~ 1, we say that u is "completely contractive" (or "a complete contraction" ). The notion of isometry is replaced by that of "complete isometry": A map u: E -> F is called a complete isometry (= u is completely isometric) if
Un: Mn(E)
->
Mn(F)
is an isometry for all n ~ l. Similarly, a map 'u: E -> F is called completely positive (in short c.p.) if Un: Afn{E) -> Afn(F) is positive for all n (in the order structure induced by the C*-algebras Mn{B{H» and Mn(B(K)). Moreover, we should emphasize
Introduction to Operator Space Theory
20
Definition 1.4. Two operator spaces E, F are called completely isomorphic if there is a linear isomorphism u: E ---+ F such that u and u -1 are c. b. We will say that E, F are completely isometric if there is a linear isomorphism u: E ---+ F that is a complete isometry (or, equivalently, that satisfies Ilulicb = lIu- 11lcb = I). In that case, we will often identify these spaces, although this might sometimes be abusive. Proposition 1.5. Let A1 C B(H1), A2 c B(H2) be two C*-algebras; let E1 C AI, E2 C A2 be two operator spaces; let 1f: A1 ---+ A2 be a representation such that 1f(Ed C E 2; and let u: E1 ---+ E2 be the restriction of 1f. Then u is completely bounded and lIuli cb ::; 1. Moreover, if 1f is injective, u is completely isometric. Proof. It is well known that a C* -algebra representation 1f automatically has norm at most 1 and a closed range (cf. [Ta3, p. 21-22]). Therefore, 111f11 ::; 1, but since 1fn : Mn(Ad ---+ Afn (A 2) also is a C* -algebra representation, we again have II1fnll ::; 1 for all n, and hence Ilulicb ::; 1. Moreover, if a representation 1f is injective, it is necessarily isometric (since its inverse must also have norm at most 1), and hence 1fn itself is isometric for all n. • We can measure the "c. b. distance" between E and F by setting
dcb(E,F) = inf{llullcbllu-11Icb I u: E
---+
F complete isomorphism}.
If E, F are not completely isomorphic, we will set
dcb(E, F) =
00.
Examples. When E, F are Banach spaces we can view them as operator spaces via the embeddings E
c
C(BE*),
Fe C(BF*).
This is of course not a very interesting operator space structure, but it shows that - to some extent - Banach space theory can be viewed as embedded into operator space theory, since for a map C(BE*)
u E
we have necessarily
u bounded {:} u
c.b.
~Maps
1. Completely Bounded
21
and
Ilull = Ilullcb· Actually (see Proposition 1.10), this remains true when E is an arbitrary operator space, assuming only that F is equipped with its "commutative structure" as above. Moreover, it is easy to check that Ilull = Ilulicb for any rank one mapping u between operator spaces. This implies of course that if dim(E) = 1, then it.s commutative operator space structure is the only possible one on E. Here are more interesting examples: In B(£2) consider the column Hilbert space C
= span{eil
R
= span{elj I j
liE foi}
(1.1 )
and the row Hilbert space E
foi}.
(1.2)
vVe will also need their finite-dimensional versions:
Cn = span{eil 11 :S i :S n} Rn = span{elj 11 :S j :S n}. Then, as Banach spaces, Rand C are indistinguishable, since they are both isometric to £2, that is, we have (1.3) However, as operator spaces, they are not isomorphic. Actually they are extremely far apart, since we have (see [Matl-2]) (1.4) which is the maximal distance possible between any two n-dimensional operator spaces. Actually, it can be shown (cf., e.g., [P5, p. 270], [ER4]) that for any u: R -4 C (or u: C -4 R) we have (HS stands for Hilbert-Schmidt)
Ilulicb = lIullHs.
(1.5)
For the proof, see the solution to Exercise 1.1. It follows that, for any isomorphism u: Rn - 4 Cn, we have
Introduction to Operator Space Theory
22
which implies dcb(Rn, en) ~ n. For the converse it suffices to observe that the map u: Rn -+ On taking elj to ejl (=transposition) satisfies Ilulleb = liullHS and lIu-llicb = lIu-lliHs Letting n -+ 00, this gives us a simple example of an isometric map from R to 0 that is not c.b. A fortiori, the transposition x - tx is isometric but is not c.b. either on B(f2) or on K. More precisely, let Tn: Afn - Mn denote the transposition of matrices. Then one can prove (see Exercise 1.2)
=.;n
=.;n.
(1.6) These examples Rand 0 are fundamental. Indeed, using the Haagerup tensor product (denoted by ®h) presented in Chapter 5, one can reconstruct the whole of B(f2) or B(H) using Rand 0 as the basic "building blocks" more precisely, we have A1n = On ®h R n , K(f2) = 0 ®h R, and of course B(f2) = K(f2)**' More generally, let HI. H2 be two Hilbert spaces and let 1-l = HI EB H 2. The mapping
x-
(x0 °0)
is an isometric embedding of B(HI' H 2) into B(1-l). Using this, we can view B(H1 ,H2) as an operator space. Note that the norm induced on l\{n(B(H1 , H 2)) by Mn(B(1-l)) coincides with the norm of the space B(f'2(Hd, f'2(H2)). In particular, we will often use the following:
Notation. Let H be an arbitrary Hilbert space. For any h E H, we denote by he E B(C, H) and hr E B(H*, q the isometric embeddings defined by V..\ E C
Vf. E H*
h e (..\) =..\h
hr(f.) = (f., h).
We will denote by He and Hr the resulting operator space structures on H. Recall that the dual H* can be canonically identified with the complex conjugate Hilbert space H. In particular, we have
Let a: HI - HI and b: H2 -+ H2 be bounded operators and let Uab: B(HI' H 2) - B(Ht. H 2) be defined by uab(T) = bTa. Clearly, Uab is c.b. and lIu ablleb :5 lIallllbli. Taking either HI or H2 one-dimensional, this implies immediately for any Hilbert space H
VU: He - He
lIulieb =
lIuli and Vv: Hr - Hr Ilvlleb = IIvll.
1. Completely Bounded Afaps Indeed, we have u(h c)
\lu: C
-4
C
23
= [u(h)Jc and analogously for r. In particular,
Ilulicb = Ilull
\Iv: R
and
-4
R
Ilvllcb = Ilvll.
(1.7)
The theory of c.b. maps clearly is the basis for operator space theory. It emerged in the early 1980s through the works of Wittstock [Witl-2], Haagerup [H3J, and Paulsen [Pa3], who proved (independently) a fundamental factorization and extension theorem for c. b. maps. This factorization is a generalization of earlier important work by Stinespring and Arveson ([St, Arl]) who proved a factorization/extension theorem for completely positive maps. Theorem 1.6. (Fundamental Factorization/Extension Theorem.) Consider a c.b. map B(H) B(K)
u E
u
--+
u F
Then there is a Hilbert space ii, a representation n: B(H) --+ B(ii),
and operators VI: K
ii, V2: ii
-4
-4
K such that
\lxEE
IIVIIII1V211
=
Ilulicb
and (1.8)
Conversely, if (1.8) holds tllen u is c.b. and Ilulicb :::; IIVIII 11V211 (in addition, if = V2*' then u is completely positive). Aloreover, u admits a c.b. extension u: B(H) -4 B(K)
VI
B(H)
~
B(K)
~
F
u E SUell
that
U
Ilulicb = Ilulicb.
For a proof, see either [Pal]' [PlOJ, or [P5J; the latter extends to the case when Hand K are Banach spaces. This theorem explains the claim that c.b. maps keep track of the operator space structure. Indeed, it shows that (as explained in the Introdl\ction) every c. b. map is the restriction of the composition of a representation and a two-sided multiplication. For emphasis and for later reference, we state as separate corollaries parts of Theorem 1.6 that will be used frequently in the sequel. The first is the extension property of B(K), which can be viewed as an operator-valued version of the Hahn-Banach Theorem:
24
Introduction to Operator Space Theory
Corollary 1.7. Let E, E be operator spaces so that E c 1£ c B(H). Then --+ B(K) admits a c.b. extension u: E --+ B(I() with lI"IIcb = IIullcb.
any c.b. map u: E
Proof. We simply let
u be the restriction of x
1-+
V27r(X)Vl to
E.
•
The second is the dilation property of unital complete contractions:
Corollary 1.8. Let E c B(H) be an operator space containing I. Consider a map u: E --+ B(K). Ifu(l) = I and IIulicb = 1, t11en there is a Hilbert space il with K c il and a representation 7r: B(H) --+ B(il) such that 'VxEE
In particular, u is completely positive. Proof. By Theorem 1.6, we have u(·) = V27r(·)Vl . By homogeneity, we may assume IIVI II = 11V2 II = 1. Since I = u(I) = V27r(I) VI = V2Vi, VI must be an isometric embedding of I( into H. Identifying K with VI (K), u(·) = V27r(.)Vl becomes u(·) = PK7r(·)IK. • Finally, the third corollary is the decomposability of c. b. maps as linear combinations of c.p. maps:
B(I() can be decomposed as U = U2 + i( U3 - U4), where Ul, U2, U3, U4 are c.p. maps wit11 IIuj IIcb ~ IIulicb.
Corollary 1.9. Any c.b. map u: E Ul -
--+
Proof. By Theorem 1.6, we have u(·) = V27r(·)Vl. Let us denote V = VI and = W*7r(·)V. Then the result simply follows from the polarization formula: We define Ul, U2, U3, U4 by
V2 = W*, so that u(·)
= 4- l (V + W)*7r(·)(V + W), U2(-) = 4- l (V - W)*7r(-)(V - W), U3(·) = 4- l (V + iW)*7r(.)(V + iW), U4(·) = 4- l (V - ilV)*7r(·)(V - iW). Ul(-)
Then IIujllcb ~ 1 for j = 1,2,3,4 and U = Ul - U2 actually IIUl + u211cb ~ 1 and IIU3 + u411cb ~ 1).
+ i(U3 -
U4). (Note that •
Proposition 1.10. Let F C B(H) be an operator space. Let AF be the C* -algebra generated by F. (i) For any n ~ 1 and any x in Mn(F) we have
IIxllM,,(F)
~sup {ilL Ai/LjXijIlF I Ai E C,I-"j L/l-"j/2 ~
I}.
E C,
L
IAil2 ~ 1,
1. Completely Bounded AJaps
25
(ii) Assume either AF commutative or dim(F) = 1. Tllen we have equa.lity in (i). Aforeover, in either case, if E is an arbitrary operator space, any bounded map u: E ~ F is c.b. and satisfies 1I11llcb = Iluli. (iii) For any E, F, every finite-rank map u: E ~ F is c.b. Proof. (i) is an easy exercise. When AF is commutative, we can assume AF = Co(f2) and also Mn(AF) = C o(f2; Mn) for some locally compact space f2. Then equality in (i) is very simple to check. When dim(F) = 1, the verification is again an easy exercise. The second assertion in (ii) then follows by applying (i) in E and the equality case in F. Thus any map of rank one is c.b., which implies the same for any finite-rank map. • Note that (ii) implies that (not too surprisingly!) there is only one abstract operator space structure on C. Remark 1.11. Let E l , E2 be two Banach spaces. Consider an element x = E ai 0 bi in the algebraic tensor product El 0 E 2 . The "injective" tensor norm (in Grothendieck's sense) is defined as
IIxliv = sUp{I(~l 0~2,X)11 ~l E BE~,6 E BE;} = sup {IL~1(ai)6(b;)116 E BE;,6 E BE:Z}' Note that we can write alternatively
J
Ilxliv = ~l~l~.:; {IIL6(ai )bi II E = 6~l~.:2 {IILai6 (bi )IIEJ· v We denote by El 0 E2 the completion of El 0 E2 for this norm, and we call it the injective tensor product of E1, E 2 . In particular, for any Banach space E, we have for any x E ]'lIn 0 E
Ilxll A/.. ®E v =suP{IIL>'iJ-tjXijlll(>'i)'(llj)Ecn'LI>'iI2~1'Ll/ljI2~1} E
=suP{IILeij~(Xij)IIA/" I~EBE'}'
(1.9)
Note that for any locally compact space f2 and any Banach space B (in particular for B = lIJn ) we have an isometric isomorphism v Co(f2, B) = C o(f2) 0 B. Remark. Let ct(n) be the best constant C such that, for any E, F, any map u: E ~ F of rank n satisfies
Ilulicb
~
Cllull·
We will see in Theorem 3.8 later that n/2 ~ ct(n) ~ n and in Chapter 7 that ct(n) ~ n/2 1 / 4 (due to Eric Ricard), but the exact value of ct(n) does not seem to be known. The following result due to R. Smith [Sm2] is often useful.
Introduction to Operator Space Theory
26
Proposition 1.12. Consider E C B(H) and u: E Then we have
---->
MN
B(e!f,e!f).
Proof. This can be proved using the fact that, if Xl, ... , Xn is a finite subset of e!f with L:~ IIxii/2 ~ 1, then (we leave this as an exercise for the reader) there are an n x N scalar matrix b = (bjk) with II(bjk)11 < 1 and vectors Xl, ... ,XN in e!f such that L:~ IIxil12 ~ 1 and N
Vj ~ n
Xj = LbjkXk. k=l
Similarly, for any Yl, ... , Yn in £!f there are a scalar matrix c II(Cil)11 ~ 1 and fiI,·.·, fiN in e!f such that L:~ IIfiil12 ~ 1 and
= (Cil)
with
N
Vi ~ n
Yi = LCi/fil. l=1
Hence for any n x n marix (aij) in Mn(E) we have n
L i,j=l
where (O:lk) Therefore:
E
N
(u(aij)Xj, Yi) = L
(U(O:lk)Xk, fil),
k,l=l
MN(E) is defined by (O:lk) = c*.(aij).b (matrix product).
• Remark 1.13. Consider a!, ... ,an and bl, ... ,bn in B(H). Let a E A1n (B(H)) (resp. bE Mn(B(H))) be the n x n matrix that has al,"" an (resp. bl , ... , bn ) on its first column (resp. row) and zero elsewhere; that is, we have
Then (1.10)
1. Completely Bounded lHaps Indeed, we have Iiall = Ila*aII 1/ 2 and Ilbll Ilball :S Ilbllllall, and hence
27
= Ilbb*11 1 / 2 . Moreover, we have
l\Jore generally, for any x = (Xij) in lIIn (B(H)) we have Ilbxall :S Ilbllll:rllllall, and hence
Note that it is easy to extend this remark to n
= 00.
Exercises Exercise 1.1. Prove (1.5). Exercise 1.2. Prove (1.6). Exercise 1.3. Let u: E -> F be a mapping between operator spaces. Show that for any aI, ... ,an in E we have II
L u(aj)*u(aj)lIl/2 :s Ilullcbli L ajajll1/2 and II L u(aj)u(aj)*lll/2 :s Ilullcbli L ajajII1/2.
Exercise 1.4. Let u: E -> F be a mapping between operator spaces. Show that Ilulicb = sup{llvullcb I v: F -> Mn Ilvllcb:S I}. n~l
Exercise 1.5. (Schur Multipliers) (i) Let {Xi I i :S n} and {Yj I j :S n} be elements in the unit ball of a Hilbert space I<. Then the mapping u: lIIn -> Mn defined by u([aij]) = [aij(xi, Yj)] is a complete contraction. In addition, if Xi = Yi for all i, then u is completely positive. (ii) More generally, let S, T be arbitrary sets. We will identify an element of B(C 2(T),C 2(S)) with a matrix {a(s, t) I (8, t) E S x T} in the usual way. Let {xs I s E S} and {Yt I t E T} be elements in the unit ball of a Hilbert space I<. Then the mapping u: B(C 2(T),C 2(S)) -> B(C 2(T),C2(S)) that takes (a(s,t))(s,tlESXT to (a(s,t)(xs,Yt))(s,tlESXT is a complete contraction. In addition, if S = T and Xt = Yt for all t, then u is completely positive.
Chapter 2. The Minimal Tensor Product. Ruan's Theorem. Basic Operations Preliminary. At the present stage, it is worthwhile to revise the definition of an operator space. We will distinguish between a concrete and an abstract one. To be more precise, let V be a (complex) vector space. Bya concrete operator space structure on V we mean the data of a linear embedding J: V ~ B(H) with which we can obtain from V (after completion) an operator space E in the sense of Definition 1.2. Now, let J 1 : V ~ B(Ht} and h: V ~ B(H2 ) be two such linear embeddings. We will say that the corresponding concrete operator space structures are congruent if the associated sequences of norms on Jlfn (V) (obtained using either J 1 or J2 ) are identical. Equivalently, this means that, for any operator space F and any map u: V ~ F, the c.b. norm of u is the same whether we use one operator space structure on V or the other. Congruence is obviously an equivalence relation. Then, by an abstmct operator space structure on V we will mean an equivalence class of concrete operator space structures on V for this relation. With this terminology, an abstract operator space is just a vector space V equipped with a sequence of norms coming from some concrete o.s. structure, just like a normed space is simply a vector space equipped with a norm. However, it is not true that any sequence of norms on the spaces 1Ifn(V) can occur. The object of Ruan's Fundamental Theorem is to identify precisely the sequences of norms that are relevant, that is, that correspond to a concrete o.s. structure on V.
2.1. Minimal tensor product Consider two operator spaces E c B(H)
and
Fe B(K).
Then their minimal (or spatial) tensor product is defined as the completion of the algebraic tensor product E Q9 F with respect to the norm induced by B(H Q92 K) via the embedding E Q9 Fe B(H Q92 K).
We denote by E Q9min F the resulting space and by II Ilmin its norm. Note that we have lIeQ9!1I = lIellll!1I for all e in E and all! in F. Clearly, EQ9millF is an operator space.. In particular, we note the completely isometric identity (2.1.1)
2. The Minimal Tensor Product. Ruan's Tlleorem. Basic Operations Note also the obvious identification H
B(H)
Q9min
Q92
K c:= K
B(K) c:= B(K)
Q92
Q9min
29
H, whence
B(H),
and after passing to subspaces, again completely isometrically E
Q9ll1in
F c:= F
Q9ll1in
E.
A tensor product with the latter property is usually called commutative. On the other hand, if G c B(L) is another operator space (on the Hilbert space L), we have (H Q92 K) Q92 L c:= H Q92 (K Q92 L) c:= H Q92 K Q92 L, which yields the associativity:
(E
Q9 m in
F)
Q9min
G c:= E
Q9ll1in
(F
Q9min
G).
More generally, if E}, ... , En is any number of operator spaces with E; c B(Hi) we define El Q9ll1in ••• Q9ll1in En as the completion of El Q9 •.. Q9 En for the norm induced by B(HI Q92 ••• Q92 Hn). We then have (E
Q9min
F)
Q9min
G c:= E
Q9min
F
Q9 m in
G,
which will allow us from now on to drop all parentheses. Let Hn C H be an n-dimensional subspace and let PH" : H ~ Hn be the orthogonal projection. Using an orthonormal basis, we may identify Hn with the canonical n-dimensional Hilbert space £2' so that we may also identify B(Hn) with AIn . Let v: B(H) ~ B(Hn) = Mn be the mapping taking x to PH" XIH n • Let en be the collection of all such mappings with Hn arbitrary n-dimensional. Note that B(Hn) Q9min F c:= AIn(F). Then it is easy to show that for any x = L: ai Q9 bi E E Q9 F we have
For the proof see the solution to Exercise 2.1.1. This shows that the norm of E Q9min F does not depend on the particular embedding of F but only on the "abstract" operator space structure of F in the above sense: Indeed, the latter determines the norms of Afn(F). Exchanging the roles of E anf F, we obtain the same for E. Moreover, using (2.1.1) and the associativity described above, one easily checks that the resulting norms on the spaces AIn(E Q9min F) also depend only on the sequences of norms on A{n(E) and AIn(F); equivalently, the equivalence class (modulo
Introduction to Operator Space T11eory
30
congruence) of E®minF depends only on those of E and F. (This also follows easily from (2.1.6) or (2.1.7).) Notation: Throughout these notes, we denote by IE the identity map on E. The above tensor product is a "good" tensor product in the sense that, if E 1 , F 1 , E 2, F2 are operator spaces and if
are completely bounded, then
is c.b. and we have (2.1.3)
As we show later, this follows immediately from the next result, using (the commutativity and) the decomposition
Proposition 2.1.1. ([DCH]) Let u: E ---+ F be a c.b. map. Then for any o.s. G the mapping IG®u: G®E ---+ G®F is bounded when G®E and G®F are equipped with the norm II Ilmin' Moreover, if we denote by UG: G ®min E ---+ G ®min F the extension of IG ® u, we actually have Ilulicb = sup lIuGIl = sup lIuGllcb, G
(2.1.4)
G
where the supremum TUns over all possible operator spaces G. Finally, for any G, if U is completely isometric, then UG is also. Proof. Using (2.1.2) applied to elements of both G ®min E and G ®min F we easily derive that, for any x in G ® E, we have
Therefore we have sup IIIG ® u: G ®min E
---+
G ®min FII ~ Ilulicb.
(2.1.5)
G
Then, taking G = Mn (n ~ 1) and recalling (2.1.1), we see that we actually have equality in (2.1.5). Finally, since G is arbitrary, we can replace G by Afn (G) and use the obvious identification
2. The Minimal Tensor Product. Ruan's Tlleorem. Basic Operations
31
It is then easy to check that sup IIIG G
0 '1111 = sup IIIG 0 Ullcb, G
whence the second equality in (2.1.4). The last assertion should now be easy to check. • Proof of (2.1.3). We have by (2.1.4)
Ilu101L21lcb
Ilul 0 u211cb
~
~
Ilu10IF21IcbIIIE[ 0U211cb, and hence
II Ulllcbll u21lcb.
For the converse, recall that for any Xl E E], X2 E E2 we have IlxlIIE,llx21IE2' From this it is very easy to deduce that
If we then replace n, In, we obtain
Ul
by hI" 0
'111
and
U2
by
Il\l",
0
'112
Ilxl 0X211min =
and take the sup over
n,m Since lIIn 0 min 111m
~
1I1nm, by commutativity and associativity we have
and this last norm is clearly (e.g., by the first part of the proof)
~
Ilu10u2l1cb .
•
Remark 2.1.2. Now that we have checked (2.1.3) it is easy to verify two more elegant variants of (2.1.2). Again, consider X = La; 0 bi E E 0 F. We have (2.1.6) Ilxllmin = sup v(a;) 0 bill. ' n,vEB"
where Bn obtain
=
{v: E
~
I IIvllcb
lIfn
Ilxllmin =
sup
IlL ~
l\l" (F)
I}. Moreover, applying this twice we
{IlL v(a;) 0 w(b;) II
AInm
},
(2.1. 7)
where the supremum runs over n, In ~ 1 and all pairs v: E ~ lIIn , w: F ~ lIIm with Ilvllcb ~ 1 and Ilwllcb ~ 1. (We can of course restrict to n = In if we wish.) It is useful to observe that this tensor product is analogous to the injective tensor product of Grothendieck in Banach space theory; that is, for any (closed) subspace SeE, the embedding
S
0 min F
-----+
E
@min F
Introduction to Operator Space Theory
32
is an isometric embedding. More generally, if are completely isometric, then
Ul: El -
Fl
and
U2: E2 -
F2
is also completely isometric. Let us denote briefly by
the space of all compact operators on £2. It is worthwhile to observe that a map U: E-F is c.b. iff l,e
®
u:
K ®min E - K ®min F
is bounded and we have (2.1.8) Similarly: U
is a complete isomorphism
U
is a complete isometry
u is completely positive
¢:}
¢:}
¢:}
l,e ® U is an isomorphism.
h:. ® U is an isometry.
l,c ® u is positive.
Actually, K can be replaced by any large enough C* -algebra in these statements; more precisely:
Definition 2.1.3. We will say that an operator space X is sufficiently large if for any c > 0 and any n there is a subspace Xn C X and an isomorphism Un: Mn - Xn such that Ilunllcbllu;;lllcb < 1 + c. We have then obviously
Proposition 2.1.4. Let X be a sufficiently large operator space and let F be a linear map between operator spaces. Then
u: E -
(2.1.9)
Moreover, u is a complete isomorphism (resp. complete isometry) iff Ix ® u defines an isomorphism (resp. an isometry) from X ®min E to X ®min F. Remark 2.1.5. When E C B(H) and F C B(K) are C*-subalgebras, E®mlnF C B(H®2K) is also a C*-subalgebra. Thus the minimal tensor product makes sense in the category of C* -algebras. We will study tensor products of C*-algebras in detail in Chapters 11-22. For the moment, note that
2. Tlle
~l\finimal
Tensor Product. Ruan's Theorem. Basic Operations
33
(see Proposition 1.5) when E, Fare C* -algebras, the resulting C* -algebra E ®min F does not depend on the concrete embeddings E C B(H} and Fe B(K), but only on the abstract C*-algebra structures on E, F. Remark 2.1.6. Let E C B(H} and F C B(/{} be operator spaces and let A be an auxiliary C* -algebra. Then the bilinear mapping PA: A®min E x A®min F
----+ A®lllin E®lllin
F
defined by has norm:::; 1. Indeed, it suffices to prove this when E = B(H) and F = B(K}. Then PA appears as a restriction of the ordinary product map of the C* -algebra A ®min B(H} ®Illin B(K); more precisely, we can write
•
whence the announced result.
Exercises Exercise 2.1.1. Prove (2.1.2). More generally, let HOI C H be a directed net of subs paces such that UHOI is dense in H. Then let VOl: E ----+ B(H,,} be defined by vOI(e} = PH"eIH". We then have IIv(e)1I = lim
"
for any e in E, and for any x IIxlimin
= lim 01
=
i IIvOI(e}1I
L Qi ® b; E E ® F
jilL
VOl (ad
® bill
we have
B(H")0m,,,F
.
Exercise 2.1.2. With the notation of the preceding exercise, consider a linear map u: F ----+ E c B(H). Show that
1171ll cb =
Slip
IIvOIulicb.
01
Exercise 2.1.3. Let HI, H2 be Hilbert spaces. Consider Ti E B(H;} (i = 1,2). Let T = TI EB T2 E B(HI EB H 2}. (This means that T(h l • h2) = (TlhJ, T2h2) V'h; E Hi'} Note that IITII = max{IITtll,IIT211}. Let E be an operator space. Consider Ui E CB(E, B(Hi)) (i = 1,2), and let U E CB(E, B(HI EB H 2)) be defined by u(x) = UI(X) EB U2(X). Show that Ilulicb = max{lI ulll cb,lI u 2I1cb}.
34
Introduction to Operator Space Theory
Exercise 2.1.4. Consider
Xi E
B(H),
Yi E
B(K) (1 :::; i :::; n). Show that
and also
Moreover, the first inequality becomes an equality either if Xi or if Yi = eli for all i.
= eil for all
i
2.2. Ruan's Theorem We will now state Ruan's Fundamental Theorem ([Rull). We consider a complex vector space E given together with a sequence of norms on the spaces Mn(E). By this we mean that, for each n, we are given a norm an on Mn(E)j moreover, we assume them compatible in the following sense: If we view Mn(E) as embedded into Mn+I(E) via the embedding that completes a matrix by adding zeroes, then an coincides with the restriction of an+l to Mn(E). Thus, rather than work with a sequence of norms, we may as well work with a single norm on the "union" of all the spaces Mn(E). More precisely, let us denote by /(0 the union of the increasing sequence All C ... C !V[n C M n+1 .... We equip /(0 with the norm induced by the spaces Aln or, equivalently, by B(f 2 ), so that the completion of /(0 for this norm can be identified with /(. Let
We define a norm a on /(o[E] in the obvious way, that is, "Ix E /(o[E]
a(x)
= n->oo lim an(x).
Note that the preceding limit is actually stationary: when X E Afn(E) then am(x) = an(x) for all m ;::: n. We will denote by /( ®a E the completion of /(0 ® E with respect to the norm a. For any x in /(o[E] and any a, b in /Co, we denote by a.x.b the (left. and right, respectively), matricial product of the scalar matrices a and b by the matrix x E /Co [E]. It is easy to verify that all the sequences of norms that come from an operator space structure on E satisfy the following two axioms (R I ) and (R 2 ). AXIOM (RI): For all n we have
2. The Minimal Tensor Product. Ruan's Theorem. Basic Operations
35
AXIOM (R2): For all integers n, m we have
where we denote by x EB y the (n
+ m)
x (n
xEBy= ( X0
+ m)
matrix defined by
0)y .
Note that (R 2 ) implies in particulaT that an coincides with the norm induced on Mn(E) byan+l. We will also consider the following
AXIOM (R): For any finite sequences (ai) and (b i ) in /(0 and for any finite sequence (Xi) in /(o[E] we have
Ruan's Theorem. ({Rul}) Let E be a complex vector space. Let (an) be as above, a sequence of norms on tll€ spaces l\{n(E), and let a be the corresponding norm on /(00 E. The following assertions are equivalent.
(i) Tlw axioms (Rt) and (R 2 ) are satisfied. (ii) Tlle axiom (R) is satisfied. (iii) For a suitable Hilbert space H, there is a linear embedding J: E --t B(H) such that, for anyn, hI" 0J is an isometry between (l\fn(E), an) and Mn(.J(E)) C l\fn(B(H)) (equipped with the induced norm). (iv) For a suitable Hilbert space H, there is a linear embedding J: E --t B(H) SUell that iKl) 0 J is an isometry between (/(00 E, a) and (/(00 J(E),II·llmin). (v) For a suitable Hilbert space H, tllere is a linear embedding J: E --t B(H) such that h:: 0 J extends to an isometric isomorphism from between /( 0 0 E and /( 0 min J(E).
In other words, Axiom (R) (resp. (R 1 ) and (R 2 )) characterizes the norms a (resp. the sequences of norms (an)) that come from a (necessarily uniquely determined) concrete operator space structure on E. Ruan's Theorem establishes a one-to-one correspondence between the set of abstract operator space structures on a given vector space E and the set of norms on /(00 E that satisfy Axiom (R) (or equivalently the set of sequences of norms on l\fn(E) satisfying (R I , R2)). It is very useful to define new and nontrivial operations on operator spaces (such as duality, quotient, interpolation), the most important one probably being the duality.
36
Introduction to Operator Space Theory
Remark. The following simple preliminary observation will be worthwhile. Assume (R). Then it is easy to show that, for any pin /(0 with IlpllK = 1, we have (2.2.1 ) a(p ® e) = a(ell ® e). 'tie E E Thus if we define the norm on E by Ilell = a(ell ® e), we obtain a(p ® e) = IIpllliell for all p in /(0 and all e in E. Indeed, we may assume p E AfN with norm equal to 1. Then there is a factorization (consequence of the polar decomposition) p = uDv, where u, v are unitary matrices in MN (viewed as sitting in /(0) and D is diagonal in JlvlN with norm 1; moreover, composing u and v with suitable permutations of the basis, we may assume that D = E~ t5ieii with max It5i l = 11511 = 1. Clearly we have, by (R), a(uDv ® e) ::; a(D ® e) and, since D = u*pv*, we also have a(D ® e) ::; a(p ® e), whence a(D ® e) = a(p ® e). Note that t51ell ® e = en . (D ® e) . en and hence by (R) we have a(en ® e) = a(t51en ® e) ::; a(D ® e).
On the other hand, we also have N
D ®e
=
Lt5ieil' (en ® e)· eli, i=l
which implies by (R) again a(D ® e) ::; a(en ® e).
Thus we conclude a(p ® e)
= a( en ® e)
for any p in
/(0
with IIpllK
= 1.
•
The following simpler proof of Ruan's Theorem appears in [ER3]. Proof of Ruan's Theorem. The main step is to show that Axiom (R) implies (iv) above. So assume that (R) holds. We first make the following Claim. For any x in /(0 ® E with a(x) = 1 there is a Hilbert space Hx and a linear mapping J x : E --+ B(Hx) such that 'tIyE/(o®E
and
1I(h: ® Jx)(x) II = a(x). Indeed, if we take this claim for granted, we may set H = tBxHx, and for all e in E J(e) = Jx(e) E B(H),
EB x
2. T1le Minimal Tensor Product. Ruan's Theorem. Basic Operations
37
where the direct sum runs over the unit sphere of (Ko 0 E, a), then we clearly obtain (iv). We now turn to the proof of the claim. By the Hahn-Banach Theorem, there is a linear functional e E (Ko 0 E)* with a*(e) = 1 such that e(x)
= a(x).
By Exercise 2.2.3, we know that there are states ft and 12 on K such that, for any Y in Ko 0 E and any a, b in Ko, we have
le(a* . y. b)1 ::; (ft(a*a)12(b*b))1/2a(y).
(2.2.2)
This implies a fortiori that for any a, b in Ko and any e in E we have (2.2.3)
Lf
Indeed, we may assume that a, bE lIfN C K o, so that if we let PN = eii, we have a*b0e = a*· (PN 0e)·b and hence (2.2.3) follows from (2.2.2) applied to y = PN 0 e, once we recall (2.2.1). Let R be the linear span of the vectors {elj I j = 1,2, ... }. Let HI (resp. H 2 ) be the Hilbert spaces obtained after passing to the quotient and completion from R equipped with the scalar product (a, bh = it (b*a) (resp. (a, b)2 = 12(b*a)). Note Ko = span[a*b I a, b E R]. Then (2.2.3) can be rewritten as follows for all a, b in Ko: (2.2.4) where a and b denote the equivalence classes (modulo the kernels of (,,·h and (" 'h, respectively) corresponding to a, b. This allows us to (unambiguously) introduce the linear map v: E ---+ B(H2' HI) characterized by the equality VeE E
(v(e)(b), a)
= e(a*b 0 e).
We know by (2.2.4) that Ilv(e)1I ::; a(eu0e), but actually much more is true. We will now show that for any y in Ko 0 E we have (2.2.5) We may asstime y = Li,j:5N eij0Yij (Yij E E). Consider elements al,··· ,aN, bl , ... , bN in R such that (2.2.6) We then have (easy verification)
i,j
i,j
38
Introduction to Operator Space Theory
where
Now, by (2.2.2), we have
le(s* . y. t)1 ::; (h(s*s)h(t*t»1/20:(y), and (2.2.6) implies h(s*s) ::; 1 and h(t*t) ::; 1. Hence we obtain
LV(Yij)(bj),ai) ::; o:(y), i,j
which implies the right side of (2.2.5). To verify the other side, just observe that
i,j
i,j
= L (V(Yij )(elj), eli), i,j
from which the left side of (2.2.5) follows immediately because
We can now complete the proof of our original claim: Taking Y = x in (2.2.5) we find o:(X) = le(x)1 ::; II(ho @v)(x)11 ::; o:(x) and hence
o:(x) Moreover, (2.2.5) shows that
= II(hu @v)(x)ll.
1,(0
.
@
v is a contraction from (K o @ E,o:) to
B(f2(H2), f2(Ht). Therefore, If we now let Hx = HI EBH2 and J x =
(0 v) 0
0 '
we obtain the announced claim. This shows that (ii) ::::} (iv). To complete the proof, let us verify that (i) ::::} (ii). Assume (R 1 ) and (R 2 ). Then, if ai, Xi, bi are as in Axiom (R), we will use the following matricial rewriting: ai . Xi . bi = aXb,
L
where a, X, b are the matrices defined by
o1
XN'
b-
(b~
bl ) :
.
2. Tlle Minimal Tensor Product. Ruan's Tlleorem. Basic Opera.tions
39
Then (R I) implies (2.2.7) and (R 2 ) (iterated (N - I)-times) implies a(X) = max{a(xt}, ... ,a(xN)}.
But we know (Remark 1.13) that
hence (2.2.7) implies (R). Thus we have verified (i) =} (ii). Since (iv) and (iii) =} (i) are obvious, this completes the proof.
=}
(iii) •
Exercises Exercise 2.2.1. Let 8 be a set and let F real-valued functions on 8 such that 'VfEF
c £00(8) be a convex cone of
supf(s)
~
o.
sES
Then there is a net (Ao,) of finitely supported probability measures on 8 such that 'VfEF lim fdAo ~ o.
J
Exercise 2.2.2. Let B I , B2 be C*-algebras, let FI C B2 and F2 C B2 be two linear subspaces, and let c.p: FI X F2 -+ C be a bilinear form such that, for any finite sets (x{) and (x~) in FI and F 2, respectively, we have
Then there are states
fI and 12 on
BI and B 2, respectively, such that
Exercise 2.2.3. In the same situation as in Exercise 2.2.2, assume given a vector space G equipped with a norm a. Let l/J: FI x G X F2 -+ C be a trilinear form such that, for any finite sets (x{) in F I , (x~) in F2 and (Xj) in G, we have
Introduction to Operator Space Theory
40
Then there are states
It and h on Bl and B 2 , respectively, such that
2.3. Dual space
Let E be an operator space. Let E* be the Banach space dual of E, and then let a be the norm induced on K, Q9 E* by the space CB(E, K,). Then this satisfies (R) or, equivalently, (R 1 ) and (R 2 ). Consequently, E* can be equipped with an operator space structure for which we have an isometric embedding K, Q9min E* c CB(E, K,), (2.3.1) or; equivalently, for which we have isometrically (2.3.1 )' E* equipped with this structure will be called the operator space dual (in short the o.s. dual) of E.
This notion, which is the key to numerous developments, was introduced independently in [BPI] and [ER2]. It is important to notice that the embedding E* c B(H) is not constructed in any explicit way. The isometric inclusion (2.3.1) remains valid when K, is replaced by an arbitrary operator space F. More precisely, we have isometric inclusions (note that F Q9min E* ~ E* Q9min F by transposition) F
Q9min
E* C CB(E,F)
and
E*
Q9min
Fe CB(E,F).
(2.3.2)
By Remark 2.1.2 and Exercise 1.4, this can be reduced to the case when F = Mn; hence it is a consequence of (2.3.1)'. These inclusions illustrate one more time the analogy between, [Minimal tensor product/Completely bounded maps] on one hand, and [Injective tensor product/Bounded maps] on the other. Moreover, it is easy to check that, for a linear operator u: E --+ F, we have: u E CB(E, F) iff u* E CB(F*, E*)
and lIu* Ilcb
More precisely, for any n
~
= lIullcb.
(2.3.3)
1 we have (2.3.4)
2. Tile Minimal Tensor Product. Ruan's Theorem. Basic Operations
41
See the solutions to the exercises for detailed proofs. For example (see Exercise 2.3 ..5), we have completely isometrically R*
~
C
and
C*
~
R.
Similarly, if E, F are operator spaces, CB(E, F) can be given an operator space structure by equipping lI1n(CB(E, F)) ('11 2': 1) with the norm of the space CB(E, .l\fn(F)). By Ruan's Theorem again, this defines an operator space structure on CB(E, F). We then have an isometric identity
Aln(CB(E, F)) = CB(E, lI1n(F)). Equivalently, this corresponds to the norm 0: for which we have an isometric inclusion K Q9(} CB(E, F) c CB(E, K Q9min F). \\'ith this structure on CB(E, F), (2.3.1) and (2.3.2) become complete isonwtries. T\Iore generally, for any operator space G, we have a complf'tel~T isometric embedding GQ9 m in
CB(E,F)
C
CB(E,FQ9min
G)
(= CB(E,GQ9rnin F)),
(2.3.5)
which becomes an equality when G is finite-dimensional. From this viewpoint, we have a completely isometric identity
E* = CB(E,q. Exercises Exercise 2.3.1. Let E be an operator space and let E** = (E*)*. Prove that the inclusion E c E** is completely isometric ([BPI, ER2]). Exercise 2.3.2. Prove (2.3.3). Exercise 2.3.3. Prove (2.3.4). Exercise 2.3.4. Let E, F be operator spaces and let u: E ---> F be a bounded map. Show that u is a complete isomorphism (resp. a completely isometric isomorphism) iff u*: F* ---> E* is also one. Let ~i E R* be the coordinate functional defined by Show that the linear isomorphism u: R* ---> C taking ~i is a complete isometry. Similarly we have C* ~ R, and also R~ ~ C n , Rn completely isometrically.
Exercise 2.3.5. ~i
to
(2:: Xkelk) = eil
C~ ~
Xi.
42
Introduction to Operator Space Theory
2.4. Quotient space Quotient by a subspace. Let E2 C El c B(H) be operator spaces. Consider the tensor norm 0: on /C ® (Ed E 2) defined by the isometric identity (2.4.1) or, equivalently, for which we have isometrically (2.4.1)'
\:In? I
Then 0: satisfies Axioms (R 1 ) and (R 2 ), and hence there is an operator space structure on Ed E2 such that /C ®a (Ed E 2) is the minimal tensor product. This was introduced by Ruan. Here again the embedding Ed E2 c B(H) is not given by any natural explicit construction; nevertheless, by Ruan's Theorem, it exists. For example, with the preceding definition, the quotient space Loo/ Hoo is given an operator space structure with which it is completely isometric to the subspace of B(f2) formed of all the Hankelian matrices. This observation is a reformulation of the "vectorial" Nehari Theorem (see §9.1 later). Proposition 2.4.1. Let q: El ~ Ed E2 be tIle canonical surjection. A linear map u: Ed E2 ~ F into an arbitrary operator space F is c.b. iff uq is c. b. and we have (2.4.2) lIullcb = lIuqllcb. The proof is obvious. These notions are entirely compatible with the rules of the Banach space duality, for instance ([BPI, ER2]): Proposition 2.4.2. For any (closed) subspace SeE of an operator space E, we have tIle following completely isometric identities: S* ~ E* /S.L (E/S)*
= S.L.
(2.4.3) (2.4.4)
Proof. By Corollary 1.7, any u E CB(S, Mn) admits an extension CB(E, Mn) with lIullcb = Ilulicb. Equivalently, this means that CB(S, Mn)
uE
= CB(E, Afn)/N (isometrically),
where N = {u E CB(E, Mn) lUIs = OJ. Going back to the definitions, this means that Mn(S*) = Mn(E*)/Mn(S.L) (isometrically), whence (2.4.3). Similarly, by (2.4.2) we have CB(E/S,Mn ) = N
(isometrically),
and hence Mn((E/S)*) = Mn(S.L) (isometrically), from which (2.4.4) follows .
•
2. The Minimal Tensor Product. Ruan's Tlleorem. Basic Operations
43
In Banach space theory, a mapping u: E --> F (between Banach spaces) is called a metric surjection if it is onto and if the associated mapping from E / ker( u) to F is an isometric isomorphism. Moreover, by the classical open mapping theorem, u is a surjection iff the associated mapping from E / ker( u) to F is an isomorphism. The complete analog is as follows.
Definition 2.4.3. A mapping u: E --> F (between operator spaces) is called a complete metric surjection (resp. a complete surjection) if it is onto and if the associated mapping from E / ker( u) to F is a completely isometric isomorphism (resp. a complete isomorphism). Remarks. (i) A mapping u: E --> F is a complete metric surjection (resp. a completely isometric embedding) iff its adjoint u*: F* --> E* is a completely isometric embedding (resp. a complete metric surjection). Indeed, this follows immediately from Proposition 2.4.2 and Exercise 2.3.4. (ii) Similarly, u is a complete surjection (resp. a completely isomorphic embedding) iff u* is a completely isomorphic embedding (resp. a complete surjection) .
(iii) Obviously, u: E --> F is a complete metric surjection (resp. a complete surjection) iff h:Ji!Ju: K®minE --> K®minF is a metric surjection (resp. a surjection). When q is as in Proposition 2.4.1 we have Vy E EdE2
IIyliEtlE2 = inf{llxlll x E El q(x) = y},
(2.4.5)
and q maps the open unit ball of El onto that of Ed E 2.
Quotient by an ideal. This concludes the basic facts on the notion of quotient for an operator space. However, in the special case when El is a C* -algebra and E2 is an ideal in E 1, the preceding infimum is attained and q maps the closed unit ball of El onto the closed unit ball of Ed E 2. Although we will need this fact only much later in this book (and we advise the reader to skip the rest of this section on first reading), we prefer to include the corresponding background in this section, as follows. (I am indebted to N. Ozawa for explaining all this to me.) The key notion is that of a "quasicentral approximate unit" (cf. [Ar4]): Let I c A be an ideal in a C*-algebra. Then there is a nondecreasing net (O"ct) in the unit ball of A with O"ct ~ 0 such that, for any a in A and any b in I,
Such a net is called a quasi-central approximate unit. See [Ar4] or [Da2] for a proof of their existence. The main properties that we will use are summarized in the following.
Introduction to Operator Space Theory
44
Lemma 2.4.4. Let (O'oJ be as above. Then,for any a in A, 1I0':j2a-aO'~/211 O'ay/2a - a(l - 0'0)1/211 ----> o. Moreover, we have
---->
o and 11(1 -
Va E A Va,b E A
IIq(a)11 = lim Iia - O'oall lim 110'~2aa~/2 + (1- 0'0)1/2b(1 - 0'0)1/211
(2.4.6)
~ max{llall, Ilq(b)11}
(2.4.7) lim 110'~/2aO'~2 + (1 - 0'0)1/2a(1 - 0'0)1/2 - all = O. (2.4.8) Proof. The first assertion is immediate using a polynomial approximation of t ----> 0, so we turn to (2.4.6). Fix e > O. Let x E A be such that q(x) = q(a), Ilxll < Ilq(a)1I + e. Then a - x E I implies 11(1 - O'o)(a - x)11 ----> O. Hence II(l-O'o)all ~ 11(1-O'o)xll+ II(l-O'o)(a-x)lI· Therefore, since 111-0'011 ~ 1, we find lim ! 11(1 - O'o)all ~ II xII < IIq(a)11 + e. This proves (2.4.6). To verify (2.4.7), note that (see Remark 1.13) 110'~2aO'~/2 + (1 - 0'0)1/2b(1 - 0'0)1/211 ~ max{llall, IIbll}.
Now, if we replace b by b' such that q(b' ) = q(b), we have (by the first assertion) lim 11(1 - 0'0)1/2(b - b' )(l - 0'0)1/211
= lim 11(1 - O'o)(b - b')11 = 0,
and hence the left side of (2.4.7) is ~ max{llall, 11b'11}. Taking the infimum over b' we obtain (2.4.7). Finally, by the first assertion above we have a fortiori 1I000a - 0'~/2aO'~/211 ----> 0 for any a in A and 11(1 - O'o)a - (1 - 0'0)1/2a(1 0'0)1/211 ----> 0, from which (2.4.8) is immediate. •
Lemma 2.4.5. Let I c A be a (closed two-sided) ideal in a C* -algebra. Let q: A ----> A/I be the quotient map. Then Vx E A, Ve > 0, 3X1 E A witl] q(xJ) = q(x) such that IIxll < Ilq(x)1I
+e
and
Ilxl - xii ~ IIxll-lIq(x)ll·
Proof. Let (0'0) be as before. We set Xl
= 0'0(xllxll-11Iq(x)ll) + (1 - O'o)x.
We will show that, when 0: is large enough, this choice of Xl satisfies the announced properties. First we have clearly q(xJ) = q(x) (since 0'0 E I). We introduce x~
= 0'~2(xllxll-lllq(x)II)0'~/2 +
(1 - 0'0)1/2x(1- 0'0)1/2.
Choosing 0: large enough, by the preceding lemma we may assume IIXI - x~ II < e/2. Moreover, by (2.4.7) we may also assume IIx~1I ~ Ilq(x)11 + e/2, whence IIxIiI < IIq(x)1I + e. On the other hand IIx - XIII = 1I0'0xllxll- 1(ll x ll-lIq(x)11)11 ~ IIxll -lIq(x)lI· Thus we obtain the following well-known fact:
•
2. The Minimal Tensor Product. Ruan's Theorem. Basic Operations
45
Lemma 2.4.6. For any x in A, there is x in A sudl that q(x) = q(x) and IIxliA = IIq(x)IIA/I. Proof. Using Lemma 2.4.5 we can select by induction a sequence x, Xl, X2, ••• in A such that q(xn) = q(x), IIxnll :s; IIq(x)1I + 2- n and IIx n - xn-lll :s; IIXn-lll-lIq(xn-dll :s; 2- n +l. Since it is Cauchy, this sequence converges and its limit has the announced property. •
x
The following generalization was kindly pointed out to me by N. Ozawa.
Lemma 2.4.7. Let I c A be an ideal in a C*-algebra A. Consider an operator space E, let Q( E) = (E @min A) / (E Q9min 1), and let qE: E Q9min A --> Q(E) denote the quotient map. Then, for any fJ in Q(E), tllere is an element yin E Q9min A that lifts fJ (i.e., qE(Y) = fj) sucll tllat lIyllmin = IIfJIIQ(E). Proof. Choose Yo in E Q9min A such that qE(YO) = fJ. It is easy to check that qE has the properties appearing in Lemmas 2.4.4 and 2.4.5 suitably modified. Therefore, repeating the argument appearing before Lemma 2.4.6, we obtain a Cauchy sequence Yo, Yl, ... , Yn, ... in E Q9min A such that qE(Yn) = fJ for all n and llYn IImin --> IIfJIIQ(E) when 11, --> 00. Thus Y = lim Yn is a lifting with the same norm as y. •
Remark. In approximation theory, subspaces for which the infimum is attained in (2.4.5) are called "proximinal" See [H\VW, p. 50] for generalizations of the preceding facts to AI-ideals. In the second part of this book, we will need the following useful lemma, for which we need to first introduce a specific notation. Let I c B he a (dosed two-sided) ideal in a C* -algebra B. Let E be an operator space. As before, we denote for simplicity
Q(E) = B Q9min E. IQ9min E Then, if F is another operator space and if u: E --> F is a c.h. map, we clearly have a c. b. map UQ: Q(E) --> Q(F) naturally associated to I B
@ U
such that
Lemma 2.4.8. If U is a complete isometry, tlJen
uQ
is also one.
Proof. As above, the proof uses the classical fact that the ideal I has a two-sided approximate unit formed of elements aa: with 0 :s; aa: and lIaa:1I :-:; 1
46
Introduction to Operator Space Theory
(see, e.g., [Ta3, p. 27]). For simplicity we assume that E c F and U is the inclusion map. Replacing E c F by Mn(E) c Mn(F), we see that it clearly suffices to prove that uQ is isometric. Let To;: BQ9minF -+ IQ9minF be the operator defined by To; (bQ9Y) = ao;bQ9y. Note that IITo;lI ~ 1 and III - To; II ~ 1. Moreover, To;(cp) -+ cP for any cP in I Q9min F. Let us denote by d(·,·) the distance in the norm of B Q9min F. To show that uQ is isometric it suffices to show that for any x in B Q9 E we have
d(x,IQ9 min F) = d(x,IQ9min E). Let e = d(x,I Q9min E) and f = d(x,I Q9min F). Since E c F, it actually suffices to check that e ~ f. Assume f < 1. Then there is cp in I Q9min F such that IIx + cpllmin < 1. We have then, since 111 - To; II ~ 1,
Finally, since x E B Q9 E, we have To;x E I Q9 E and 11(1 - To;){cp)llmin therefore we obtain
and we conclude (by homogeneity) that e
~
f.
-+
0;
•
Exercises Exercise 2.4.1. Let u: E -+ F be a c.b. map between operator spaces. Show that the following are equivalent: (i) For any operator space G, the map IG Q9 u: G Q9min E -+ G Q9min F is surjective. (ii) There is a constant C such that, for any finite-dimensional subspace Fi c F, there is a linear map v: Fi -+ E with Ilvllcb ~ C such that uv(x) = x for all x in Fl. Moreover, when E = B(H), the above are also equivalent to the following kind of "local lift ability" : (iii) There is a net of of cb maps Vi: F -+ E with SUPi Ilvilicb ~ C such that UVi -+ IF pointwise. Exercise 2.4.2. Let E be an operator space. Let I C A be an ideal in a C* -algebra and let q: A -+ AI I be the quotient map. Consider a bounded linear map u: E -+ AI I. Assume that there is a net of complete contractions Vo;: E -+ A such that qvo; -+ U pointwise on E. Show that, if E is separable, then U admits a completely contractive lifting v: E -+ A, that is, we have IIvllcb ~ 1 and qv = u. (This is essentially due to Arveson [Ar4].) Hint: Let {Xi} be a dense sequence in the unit ball of E. Assume given a complete contraction (c.c. in short) Vn such that
Vi = 1, ... ,n.
2. The l\finimal Tensor Product. Ruan's Theorem. Basic Operations V n + 1:
Show that there is a c.c.
E
->
47
A such that 'Vi
= 1, ... , n + 1
and
I/(V n+1 Then v(x)
= lim vn(x)
-
Vn)Xil/ < 2.Tn
'Vi = 1, ... , n.
is the desired lifting.
2.5. Bidual. Von Neumann algebras For the convenience of the reader, we recall now a few facts concerning von Neumann algebras. A von Neumann algebra on a Hilbert space H is a selfadjoint subalgebra of B(H) that is equal to its bicommutant. For M c B(H) we denote by ]JJ' (resp. M") its commutant (resp. bicommlltant). By a well-known result due to Sakai (see [Sa]), a C* -algebra A is C* -isomorphic to a von Neumann algebra iff it is isometric to a dual Banach space, that is, iff there is a closed subspace X c A** SlIch that X* = A isometrically. If it exists, the predual X is unique and is denoted by A*. In particular, B(H)* can be identified with the space of all trace class operators on H, equipped with the trace class norm, and it is easy to check that a von Neumann algebra 111 c B(H) is automatically a(B(H), B(H)*)-closed. Thus, when it is a dual space, A can be realized as a von Neumann algebra on a Hilbert space Hand its predual can be identified with the quotient of B(H)* by the preannihilator of A. Note that any C*-isomorphism u: Aft -> Ah from a von Neumann algebra onto another is automatically bicontinuous for the weak-* topologies (a(Aft, ]JJh) and a(Ah Ah*)). Therefore the preduals are automatically isometric (see, e.g., [Ta3, p. 135]). A map u: lIf1 -> AI2 between von Neumann algebras is called normal if it is continuous for the a(Al1 , Alh ) and a(Ah, AI2 *) topologies or, equivalently, if there is a map v: 1112 * -> Alh of which u is the adjoint. Let A be a C* -algebra. The bidual of A can be equipped with a C* -algebra structure as follows: Let 7ru: A -> B(H) be the universal representation of A (Le., the direct sum of all cyclic representations of A). Then the bicommutant 7r u (A)", which is a von Neumann algebra on H, is isometrically isomorphic (as a Banach space) to the bidual A**. Using this isomorphism as an identification, we can now view A** as a von Neumann algebra, so that the canonical inclusion A -> A** is a *-homomorphism. The reader who prefers a more "concrete" description of the product operation on the bidual of A can use the following alternate viewpoint: For any pair a,b in A** let (ai) and (b j ) be nets in A with lIaili :::; lIall and Ilbill :::; Ilbll, converging respectively to a and b in the Mackey topology T(A**, A*). Then
48
Introduction to Operator Space Tl1eory
the products (aibj) converge in the same topology to an element of A** denoted by a· b. Similarly, (ai) converges, for r(A**, A*), to an element of A** denoted by a*. It can be shown that A ** is equipped with this product (called the Arens product), and this involution becomes a C*-algebra, which can be identified with the one just defined. We refer the reader to, for example, [BD, p. 213J for more information. We now return to operator spaces. Of course the bidual of an operator space E can be equipped with an o.s.s. simply by reiteration of the preceding definition in the preceding section, that is, by letting E** = (E*)*. This of course immediately raises the natural question of the relationship between the Banach bidual and its operator space counterpart. Fortunately, at this point the situation is very nice. Indeed ([BPI, Theorem 2.1J or [ER2, Theorem 2.2]) the inclusion E c E** is a complete isometry (for a proof, see the solution to Exercise 2.3.1). Therefore, if the underlying Banach space is reflexive, E and E** are identical operator spaces. In short, the Banach space reflexivity alone guarantees the operator space. More generally, we have
Theorem 2.5.1. ({B2, Th. 2.5]) For any operator space E, we have an isometric identity (2.5.1) Note that the two sides of (2.5.1) can clearly be identified as vector spaces. But we also claim in (2.5.1) that their norms are the same. For a proof, see Corollary 5.11 below. A moment of thought shows that, since (2.5.1) holds for any n, actually it must also be a complete isometry. In particular, let A be a C*-algebra. We equip A with its natural o.s.s. (coming from its Gelfand embedding into B(H». Then, by the preceding definition, the successive duals A*, A**, A***, and so on, can now be viewed as operator spaces. We will refer to these operator space structures on A * , A **, A ***, and so on, as the natural ones. Now assume that A is isometric to a dual Banach space, that is, we have A = (A*)* and, as explained above, A can be realized as a von Neumann algebra. Then, the inclusion A* C (A*)** = A* allows us to equip the predual A. with the o.s.s. induced by the one just defined on the dual A*, so we obtain an operator space, denoted by A~s, having A* as its underlying Banach space. Here again, a natural question arises: If we now consider the dual operator space to the one just defined, namely, (A~S)., do we recover the same operator space structure on A? Fortunately, the answer is still affirmative ([B2, Theorem 2.9]): We have (A~S)* = A completely isometrically.
2. Tlle
~Minil1lal
Tensor Product. Ruan's Tlleorem. Basic Operatiolls
49
In particular ([B2, Cor. 2.6]), this implies that, if A is an arbitrary C*algebra, the natural o.s.s. on A **, associated to its von Neumann algebra (or C*-algebra) structure (as described above), coincides with its o.s.s. as the bidual (in the o.S. sense) of A (itself equipped with its natural o.s.s.). However, if we leave the realm of C* -algebras, the situation is not as nice: Indeed, Le l'vlerdy [LeM3] has shown that, if A is merely an operator space that is at the same time a dual Banach space, then this can fail. Actually (see [LeM3]), it can happen that A is not completely isometric to any dual in the operator space sense. This is closely related to the lack of local reflexivity of B(H); see Chapter 18 for more on this topic. Let M c B(H), N c B(K) be von Neumann algebras with preduals A[*, N*. We will denote by M0N c B(H 02 K) the von Neumann algebra generated by Af 0 m in N. Thus Af 0N coincides with the bicommutant of M 0 N or, equivalently, with the closure of .M 0 N with respect to the weak-* topology of B(H 02 K). It is known (see, e.g., [Ta3]) that M0N is - up to isomorphism - independent of the particular "realizations" Al c B(H), N c B(K). It is the analog of the minimal tensor product in the von Neumann algebra framework. Moreover, M* 0 N* is dense in the predual (A/0N)*. Let us denote B = B(H 02 K) for simplicity. For any x in JIl*, the linear mapping x 0 I: Al 0 N ----> N is continuous with respect to the topologies a(B, B*) (induced on (M0N) and a(N, N*)). Hence this mapping unambiguously extends to one from M0N to N, which we will still denote (abusively) by x 0 I. Thus any tin M0N defines a linear map J(t): M*
---->
N by setting
J(t)(x) = (x 0I)(t). We can then state (cf. [ER8]):
Theorem 2.5.2. Tlle correspondence t ----> J(t) is a completely isometric isomorphism from lvf0N to CB(Af*, N), so that we have M0N
~
CB(M*, N)
(completely isometrically).
Proof. Since 111* 0 N* is dense in (M0N)*, it is clear that J is injective. Moreover, it clearly suffices to prove that t ----> J(t) is isometric, since "completely isometric" will follow automatically by replacing N by Mn(N).
Step 1: J is an isometric isomorphism when N is finite-dimensional. This is clear by (2.3.2). M
Step 2: We will now prove that J is an isometric isomorphism in the case Let {KoJ be a net of finite-dimensional subspaces of
= B(H), N = B(K).
Introduction to Operator Space Theory
50
K, directed by inclusion and such that UKo be the operation of compression defined by
We will denote by JB: B(H). this case.
-f
= K. Let Po: B(K)
-f
B(Ko)
B(K) the mapping corresponding to J in
Consider an element t in B(H)®B(K) = B(H ®2 K), and let u = JB(t): B(H)* - f B(K). For any Q let to = PH®2 KotIH®2 Ko E B(H ®2 Ko). Clearly to corresponds to Pou for the map J associated to the pair B(H), B(Ko); hence, since (as already observed) the finite-dimensional case is clear, we have by (2.3.2) Observing that UH ®2 Ko
=H
®2 K, we can write
o
Iltll = sup Iitoll
and also
0
(see Exercise 2.1.2) lIulicb = sup IlPoullcb, and hence we obtain o
V t E B(H)®B(K)
To complete this step, it remains to show that JB is onto CB(B(H)*, B(K)). For that purpose, let u E CB(B(H)*, B(K)). Let to E B(H ®2 Ko) be the tensor corresponding to Pou (using Step 1). Note that if Ko C K(3, we have necessarily PK"t(3IK", = to (since J is 1-1); hence we can "piece the to together," that is, there is a single operator t E B(H ®2 K) such that t = PH®2KotIH®2 Ko and Iltll = sup IItoli. Obviously we must have JB(t) = U. o
This completes the proof of Step 2.
Step 3: We will now prove that J is isometric and surjective in the general case. Let i: N - f B(K) denote the inclusion. Note that the inclusion M C B(H) is the adjoint of a complete metric surjection q: B(H)* - f M*. Consider a c.b. map u: M* - f N. We need to show that there is a t in AI®N with Iltll = Ilulicb such that J(t) = u. By Step 2, there is t in B(H®2K) with IItll = Iliuqllcb such that JB(t) = iuq. By Proposition 2.5.4, we have Iltll = Ilulicb. We claim that t E ]I.{ ®N. The proof of this is less elementary than may seem at first glance; it is based on a rather deep fact, the Tomika-Takesaki Theorem, which asserts that (M®N)' = M'®N' (cf. [Ta3, p. 226]). Equivalently, we have (M®N) = (M®N)" = (M' ® N')'. Thus, to show that t E AI®N, it suffices to prove that t commutes with AI' ® N' or, equivalently, that t commutes both with I ® N' and M' ® I. To verify this, let us assume for simplicity that H = £2(1), so that an element t of B(H ®2 K) can be identified with a matrix {tii I i,j E I} with
2. The Minimal Tensor Product. Ruan's Theorem. Basic Operations
51
entries t ij in B(K). Recall that for any x in B(H)* we have (x ® J)(t) = JB(t)(X) = iuq(x), and hence (x ® I)(t) E N. Equivalently, this means tij EN for any i,j. Thus for any y in N' we have
(J ® y)t = [ytij ) = [tijY) = t(J ® y), which shows that t E (J ® N')'. Exchanging the roles of M and N, we find that t E (M' ® I)'. Thus we conclude t E (Af' ® N')' = M"0N by the Tomita-Takesaki Theorem. This completes the proof of Step 3. •
2.6. Direct sum This notion is defined in the obvious way. Let (Ei)iEI be a family of operator spaces. Assume Ei C B(Hi). Let H = tBiEIHi be the Hilbertian direct sum. We will denote by tBiEIEi the operator space included in B(H) that is formed of all operators on H of the form x = tBiElxi with Xi E Ei and SUPiEI IIxi II < 00. It is easy to check that IIxll = sUPi.EI IIxi II. t-.'lore generally, for any X in Mn(tBiEIEi) we have (2.6.1) where Xi E Afn(Ei ) denotes the i-th coordinate of X. To generalize further, let F be any operator space ( e.g. F = K), let X E F ® (tBiEIEi) and let (Xi)iEI be the family naturally associated to X, with Xi E F ® E i ; then it is easy to check that
(2.6.2) In particular, for each i in J, both the inclusion of Ei into tBiE I Ei and the projection fr~m tBiEI Ei onto Ei are complete contractions. The latter is a complete metric surjection. We will denote by Co ( {Ei liE I}) the subspace of tBiE I Ei formed of aU families x = (Xi) for which Ilxi.IIEi -- 0 when i tends to infinity (along the net of complements of finite subsets of I). From the Banach space viewpoint, it would be more natural to denote tBiEI Ei instead by .eoo ({ Ei liE J}), but nevertheless we will use the notation tBiEIEi' which is widely used by operator algebraists. This notion of direct sum is the natural one when all the spaces Ei are C*algebras, and in that case, tBiEIEi also is a C*-algebra. However, in Banach space theory, there are many other possible direct sums. For instance, given
52
Introduction to Operator Space Theory
two Banach spaces Eo. EI and 1 :::; p :::; 00, one defines Eo EElp EI as Eo EEl EI equipped with the norm II (xo, xdll = (1lxollko + IIxlllk)I/P. When Eo, EI are given with an o.s.s. it is possible to also equip Eo EEl p EI with a natural o.s.s. (see §2.7). For the moment, we will describe this only for p = 1. Let P be the family of all possible pairs U = (uo, UI) of completely contractive mappings Uo: Eo ---4 B(Hu ), UI: EI ---4 B(Hu) (Hu = Hilbert). We define an embedding
by setting J(xo EEl Xl) = ffiuEP[UO(Xo) +UI(XI)]' It can be checked that J is an isometric embedding, and since ffiuEP B(Hu) is equipped with a natural o.s.s. (as a C*-direct sum), we obtain a natural o.s.s. on Eo EElI Er. It is easy to verify that this o.S.S. is characterized by the following universal property: For any operator space E, for any complete contractions Uo: EO---4 E and UI: EI ---4 E, the mapping (XO,XI) ---4 uo(xo) + UI(XI) is a complete contraction from Eo EElI EI to E. It is rather easy to check that we have completely isometric identities
We have restricted ourselves to the sum of two spaces, but everything we said extends to t'l-direct sums of an arbitrary family (Ei)iEI of operator spaces. We will denote by t'l ( {Ei liE I}) the resulting space. Again, the coordinatewise projection from t'l ( {Ei liE I}) onto Ei is clearly a complete metric surjection. Exercises Exercise 2.6.1. Show that t'1({Ei liE I})* = (fJiEIEi. Exercise 2.6.2. Show that CO({Ei liE I})* Exercise 2.6.3. Show that CO({Ei liE I})**
= t'1({Ei liE I}). = EEliEIEi*.
2.7. Intersection, Sum, Complex Interpolation Let (Eo, Ed be a compatible couple of Banach spaces in the sense of interpolation. This means that (Eo, E I ) comes equipped with continuous injections jo: Eo ---4 X and jl: EI ---4 X into a topological vector space X. For example, when Eo = Loo(JR.), EI = LI(JR.), one usually takes X = Lo(R.). Actually (see below), we can always reduce to the case when X is a Banach space. In 1960, A. Calderon (and J. L. Lions independently) introduced the complex interpolation spaces Eo
= (Eo, Edo and EO
= (Eo, Ed o
2. T11e Minimal Tensor Product. Ruan's Theorem. Basic Operations
53
associated to the parameter O<()<1. Roughly, the intermediate space Eo is a mixture of Eo and E 1; it varies continuously from Eo to E1 when () goes from 0 to 1 (cf. [C, BeL]); when () = 1/2, its unit ball behaves as some sort of "geometric mean" of the unit balls of Eo and E 1 • Let us briefly recall the main points of this theory (see [BeL, KPS]). One of the equivalent definitions of Eo is as follows. Let S = {z Eel 0 < Re( z) < I} be the usual vert.ical strip. Its boundary decomposes as oS = 00 U 01 with 00 = {z I Re( z) = O} and 8 1 = {z I Re(z) = I}. We denote by F(Eo,Ed the set of all bounded continuous functions f: S ~ Eo + E1 that are holomorphic on S and such that flu" and fla l are bounded continuous functions wit.h values, respectively, in Eo and E 1 • We define
oS
IIfIIF(E".Etl = max{sup Ilf(z)IIEo , sup Ilf(z)IIE1}' zEao
ZEal
Then the space Eo is defined as the subspace of Eo + E1 formed of all the values f(()) when f runs over F(Eo, E1)' Equipped with the norm IlxllEo = inf{llfIIF(Et),Etl I f E F(Eo, Et},j(()) = x},
(2.7.1)
it becomes a Banach space. Actually, it is often convenient to work with a smaller dass of functions than F(Eo, Ed, namely, the functions f for which there is a finite-dimensional subspace AI c Eo n E1 such that f(z) E lIf. Note that, viewed as an AI-valued function, f is then (automatically) bounded continuous on S and analytic on S. We will denote by Fo(Eo, Ed this subspace, which is useful because of the following (cf. e.g. [Sta]). Lemma 2.1.1. T11e space EonE1 is dense in Eo for any 0 < () < 1. Moreover, for any x in Eo n E1 we have
IlxllEe = inf{llfIIF(En,Etl
If E Fo(Eo, E 1), f(())
= x}.
(2.7.2)
Lemma 2.1.2. Let Eo and E1 denote, respectively, the closure of Eo n E 1 in Eo and E 1 • Then we 118ve an isometric identity
Proof. This is an immediate consequence of the previous lemma.
•
54
Introduction to Operator Space Theory
We will not define the space Ee precisely; it is a larger space (containing Ee isometrically; see [Be]) obtained as the values f«() when f is allowed to admit noncontinuous and even non-Bochner measurable boundary values, which can be described in terms of vector measures. This space is essential if one wants a precise description of the dual of Ee; however, it is possible, for our purposes here, to mostly avoid using Ee, which explains why we choose not to give more details. Now if Eo, El are equipped with operator space structures, we define the norm 0: by the identity (2.7.3) and similarly for Ee. More generally, for any n ;::: 1 we have (2.7.4) and also for rectangular matrices for any n, m ;::: 1 (2.7.5) See Corollary 2.7.7 for an extension of (2.7.5). Again, it is then easy to check the axioms (Rl' R2 ) so that we now have extended the complex interpolation functor to the category of operator spaces. Remark 2.7.3. For instance, given two arbitrary operator spaces E, F, we may consider Eo = E E9 F and El = E E91 F in the sense of §2.6. Since these have the same underlying vector space, they clearly form a compatible couple. By §2.6, both Eo and El have natural o.s. structures. Thus we can define the o.s.s. on the space E E9p F (1 < p < (0) as the one obtained by interpolation, that is, we set E E9 p F
= (Eo, Ede with () =
lip.
A similar argument applies for t'p ({Ei liE I}). The resulting operator space has many of the nice properties one would expect, for instance, we have a completely isometric identity
(Eo E9p Ed* = E~ E9p ' Ei In interpolation theory, the intersection Eo n El and the sum Eo + El play an important role. We will first review the analogous definitions for operator
2. Tlle lHinimal Tensor Product.
RWUl'S
Theorem. Basic Operations
55
spaces. Recall that our assumption that (Eo, E l ) form a com.patible couple means that there are continuous injections
of Eo and El into a topological vector space X. For simplicity, one usually identifies Eo and E I , respectively, with jo(Eo) C X and jl(El ) C X. Then Eo n El is the setwise intersection, equipped with the norm IlxllE"nEl = ma..x{llxIIE", IIxIIEJ. Similarly, Eo
+ El = {xo + Xl I Xo
E Eo, Xl E E 1 } with
With this norm, Eo + El becomes a Banach space, which we may substitute, without loss of generality, to the space X. Thus X can always be assumed to be a Banach space. Now suppose that Eo and El are equipped with operator space structures. We define an operator space structure on EonEI via the isometric embedding
defined by j(x) = (x, x), where, of course, Eo EEl EI is equipped with the o.s.s. defined in §2.6. Similarly, we can equip the sum Eo + EI with the operator space structure induced by the isometric identity
where A = {(x, -x) I x E Eo n Ed. Indeed, we have already defined the natural o.s.s. of Eo EElI El (§2.6) and of quotient spaces (§2.4), so this identity endows Eo + EI with a natural o.s.s. Equivalently, if we let I' be the set of all maps u: Eo + El -+ B(Htt ) such that IlulE"llcb ::; 1 and lIulEllicb ::; 1, then the mapping j: Eo + El -+ EElUEI,B(Hu) defined by j(x) = EElUEI'U(X) is a complete isometry. Therefore, the resulting o.s. Eo + El is characterized by the property that a linear map u : Eo + EI -+ B(H) is a complete contraction iff lIulE"llcb ::; 1 and lIulE11lcb::; l. It is easy to check that the Banach space K Q9 m in (Eo n E 1) can be identified isometrically with (K Q9min Eo) n (K Q9min El). The analogous property for the sum Eo + El is valid but only with an extra numerical factor equal to 2; namely, we have a natural isomorphism
Introduction to Operator Space Theory
56
that satisfies isometrically
IItPlicb :::;
2 and IItP-lllcb :::; 1.
(Note however that we have
Now assume that EonEl is dense both in Eo and in El (for their respective norms). Consequently, we have natural continuous injections
that allow us to consider (Eo, Ei) as a compatible couple. It is then easy to verify that
(Eo n Et)*
~
Eo
+ E;
and (Eo
+ Et)*
~
Eo n E;
completely isometrically. Then, by Calderon's fundamental duality theorem (cf. [C, BeL]), we have an isometric identity More explicitly, the duality can be described as follows. There is a natural inclusion (2.7.6) Indeed, consider ~ in (Eo, Ei)o and x in (Eo, El)o. By density, we may assume in addition that x E Eo n E l . We will show that
(Note that this makes sense since ~ E Eo + E; = (Eo n Ed*.) Indeed, for anye > 0, there are 1 E Fo(Eo, Ed and g E F(Eo , Ei) such that I«() = x, g«() = ~, 1I/11.F(Eo,Etl < Ilxll
+ e,
and
IIgll.F(E(~,Ei)
<
II~II
+ e.
Note that since the range 1(8) is bounded in Eo n E l , the function F: z
-+
(g(z), I(z»)
is bounded and continuous on S and analytic on S. By the maximum principle we have 1(~,x)1 = IF«()I :::; sup IF(z)1 zE8S
:::; (lIxll + e)(II~1I + e), which shows, since e > 0 is arbitrary, that (2.7.6) makes sense and is contractive.
2. The lI1inimal Tensor Product. Ruan's Theorem. Basic Operations
57
Theorem 2.7.4. The inclusion (2.7.6) is 8n isometric embedding, which is surjective if either Eo or El is reflexive. IHoreover, when Eo and El are operator spaces, (2.7.6) is completely isometric. The isometric case is due to Calderon [C] and Bergh [Be]. The analog for operator spaces is proved in [PI]. This extension, however, is much less trivial than may seem at first glance. At the bottom it uses some classical results due to Masani and Wiener and Helson and Lowdenslager on the factorization of matrix-valued analytic functions. The preceding statement is closely related to the paper [Ko]. The interpolation "functor," which associates to a pair (Eo, Ed the interpolated space (Eo, Ede, has many nice properties. The fundamental one is the interpolation property: Let (Eo, Ed and (Fo, Ft) be two compatible couples. If an operator T: Eo n El ---+ Fo n Fl extends boundedly both from Eo to Fo and from El to F l , then T extends boundedly from, (Eo, Ede to (Fo, Ft}e, and moreover
When (Eo, Et) and (Fo, Ft) are all operator spaces, it is easy to verify that the preceding statement remains valid with c.b. in the place of "bounded"; moreover we have
More generally, one may interpolate multilinear maps (and not only linear ones). In Chapter 7, we will make much use of a particular case, as follows.
Lemma 2.7.5. Let (Eo, Et), (Fo, Ft) be two compatible pairs and let u: (Eon Ed x (Fo n Ft) ---+ C be a sesqui1inear form that is of norm 1 simultaneollsly on Eo x Fo and El x Fl' Then u extends to a contractive sesqllilinear form on (Eo, Elh/2 x (Fo, Ft}1/2' Proof. This is routine from the definition, once one observes that, if f E F(Eo,Et) and 9 E F(Fo,Ft}, then the function z ---+ u(J(z),g(z)) is analytic .
•
Many other properties of the complex interpolation functor (e.g., reiteration) have immediate extensions to the operator space framework and are easy to prove using (2.7.3). However, this functor has a very important defect, which is well known to specialists, but is often overlooked by nonspecialists: It is neither injective nor projective. By this we mean that if we interpolate between two closed
58
Introduction to Operator. Space Tlleory
subspaces (resp. quotient spaces) of Eo and E 1 , we do not get a closed subspace (resp. quotient space) of the interpolation space (Eo, Et}o. l\Iore precisely, given a compatible Banach couple (Eo, Ed, if we give ourselves a linear subspace S c Eo n E1 and if we define So = SEll, Sl = SEt, it is in general not true that SEo coincides with (So, Sdo; that is, in general we have -Eo
S
=1=
(So, Sdo,
and the corresponding norms are not equivalent. Just the same, if we give ourselves a closed subspace R c Eo + E1 and if we define Ro = R n Eo CEo, R1 = R n E1 c E 1, then again in general we have
A simple and classical counterexample is provided by the Rademacher functions (see §9.8) that span £1 inside Loo and span £2 inside L 1, but nevertheless still span £2 (=1= (£11£2)0) in the intermediate spaces Lp = (Loo,Lt}o for 1 < p < 00, () = lip. There is however a classical instance where this difficulty disappears, when we have a linear projection that is simultaneously bounded from Eo to So and from E1 to Sl (or from Eo to Ro and from E1 to R 1). In that case, it is easy to check that we do have (So,St}o
~
-Eo
Sand
(Ro, Rt}o
= Rn Eo
with equivalent norms. The extension to operator spaces is immediate, but we state it for further reference. Proposition 2.7.6. Let (Eo, E 1 ) be a compatible couple of operator spaces. Let So, Sl be as above. Assume that there is a c.b. linear projection P: Eo -4 So that also extends completely boundedly to a projection from E1 to Sl' Then we have a completely isomorphic identification
and P defines a c.b. projection from Eo to SEo. Moreover, a similar statement is valid for the couple (Ro, Rt). In particular we have Corollary 2.7.7. Let (Eo,E1 ) be as before. Let Xc IC be a subspace for which there is a completely contractive projection Q: IC -4 X. Then we have a completely isometric identity (X
®min
Eo, X
®min
Et}o ~ X
®min
Eo.
2. The Minimal Tensor Product. Ruan's Theorem. Basic Operations
59
Proof. We recall (2.7.3) and apply Proposition 2.7.6 with (So,St) replaced by the couple (X ®min Eo, X ®min Et). • Remark 2.1.8. In general (Eo, Et}o depends very much on the choice of "compatibility" underlying the couple (Eo, E i ). Nevertheless, there is a simple "invariance principle" as follows. Let (Eo, Ed and (Fo, Ft) be two compatible pairs of Banach spaces. We assume the "compatibility" given by continuous injections io V E 0----+<1.,
D
r
°
jo v ----+ <1. ,
into (say) a Banach space X. Moreover, we assume that there are isometric isomorphisms Uo: Eo ---t Fo and Ui: Ei ---t Fi such that ik = jkuk (k = 0,1). This implies that (uo, Ui) unambiguously defines a map u: Eo +Ei ---t Fo +Fl (and also one from Eo n Ei to Fo n Ft). Then, this mapping U induces an isometric isomorphism from (Eo, Ei)o to (Fo, Ft}o for any 0 < () < 1. Clearly, in the operator space case, if we assume furthermore that Uo and Ui are actually completely isometric, then we obtain
(Eo, Et}o
~
(Fo, Fi)o
(completely isometrically).
Remark. In analogy with the complex case presented in [PI], the "real interpolation theory" (it la Lions-Peetre; cf. [BeL]) has been developed for operator spaces recently by Quanhua Xu (cf. [Xu]).
2.8. Ultraproduct We first recall several definitions. Let U be a nontrivial ultrafilter on a set I. Let (Ei)iEI be a family of Banach spaces. We denote by £ the space of all families x = (Xi)iEI with Xi E Xi such that SUPiEI II Xi. II < 00. We equip this space with the norm Ilxll = sUPiElllxi.ll. Let nu c e be the subspace formed of all families such that limu Ilxi II = o.
Definition. The quotient £jnu is called the ultraproduct oftlle family (Ei)iEI with respect to U. We denote it by IIiEIE;./U. It is important to observe that, for every element ± in £jnu admitting x E £ as its representative modulo nu, we have
Hence, the ultraproduct IIiEI E;./U appears as "the limit" ofthe spaces (Ei )iE I along U.
60
Introduction to Operator Space Theory
Remark. It is easy to check that, if (Si)iEI is a family of closed subspaces (Le., Si C Ei for each i), then the ultraproduct of the subspaces ITiE1SdU can be identified isometrically with a dosed subspace of the original ultraproduct ITiE! EdU. Similarly, the quotient of the ultraproducts can be identified with the ultraproduct of the quotients. Thus Banach space ultraproducts are both "injective" and "projective." Now assume that each space Ei is given equipped with an operator space structure. It is very easy to extend the notion of ultraproduct to the operator space setting. We simply define (2.8.1) This identity endows lIfn®(ITiEIXdU) with a norm satisfying Ruan's axioms, whence after completion a norm on K ® (ITiEIXdU). Alternatively, we can view an operator space as a subspace of a C* -algebra, observe that C* -algebras are stable by ultraproducts, and apply the preceding remark to realize any ultraproduct of operator spaces as a. subspace of an . ultraproduct of C* -algebras. This alternate route leads to the same operator space structure as in (2.8.1). Yet another equivalent description of I1iEI Ei as an operator space is the following: Consider the direct sum £ = E9iEI Ei as defined in the previous section. Then I1iEI EdU can be identified with the quotient operator space £/nu. It is easy to verify that the injectivity and projectivity of ultraproducts is preserved in the operator space setting: Let Si C Ei be a family of subspaces of operator spaces E i • Then we have a completely isometric embedding
II SdU c II EdU; iEI
(2.8.2)
iEI
moreover, the quotient (I1iEI EdU) / (I1 iE I SdU) can be identified completely isometrically with the ultraproduct I1iEI (Ed Si) /U. These properties follow easily from the definition (2.8.1) and the corresponding isometric identities mentioned in the above remark. Let us denote Eu = I1iEI EdU. Now let (Fi)iEI be another family of operator spaces and, again, let Fu = I1iEI FdU. For any i in I, consider a map Ui in CB(Ei,Fi ) with SUPiEllluillcb < 00. Then, (Ui)iEI defines a linear map from .Eu to Fu taking the equivalence class of (Xi)iEI to that of (Ui(Xi))iEI. Let
u: Eu -+ Fu be the resulting mapping. In the Banach space setting, it is well known that (2.8.3)
2. The lIIinimal Tensor Product. Ruan's Theorem. Basic Operations We claim that
u is c. b.
61
and
Ilulicb ::; limu Iludlcb.
(2.8.4)
l\Iore precisely, for any integer N and any map u: E ---> F between operator spaces, let lIullN = IIIJIlN ® uIIJllN(E)-->AlN(F)' As an immediate consequence of (2.8.1) and (2.8.3), we have (2.8.5) Whence
Ilulicb = sup lIullN = suplimu Ilu;IIN ::; limu Ilu;llcb, N
N
so we obtain (2.8.4) as announced. Note that if F is a finite-dimensional operator space and if F; = F for all i in I, then Fu = F canonically and completely isometrically (the identification Fu ---> F is the map that takes (X;);EI to limu .7:;). We now fix N :2: 1 and consider the particular case when F; = IIIN for all i in I. Then, by Proposition 1.12. we have
lIu;llcb = IluillN. Therefore, if we identify
Fu
with III N , we can rewrite (2.8.5) as follows: (2.8.6)
vVe claim that (2.8.6) implies that we have a completely isometric embedding
II E; jU (II E;/U) * '->
;EI
(2.8.7)
;EI
Indeed, the case N = 1 in (2.8.6) yields that the correspondence CUi );EI ---> U defines an isometric embedding and the general case of (2.8.6) (together with (2.3.2)') shows that it is completely isometric. In particular: Lemma 2.8.1. Fix d :2: 1. Let (E;);EI be a seqllence of d-dimellsiollal operator spaces. Then we have a completely isometric identification
(II. E;/U) * ~ II E; jU. iEI
(2.8.8)
;.EI
In other words, the ultraproducts "commute" with the duality.
Proof. Elementary (linear algebraic) considerations show that DiEI E;/U is d-dimensional and the same for D;E! E; jU. Hence (2.8.8) follows from .. (2.8.7). Analogous results can be formulated for the minimal tensor product.
62
Introduction to Operator Space Theory
Proposition 2.8.2. Let (Ei)iEI be an arbitrary family of operator spaces with I and U as above. Then, for any finite-dimensional operator space E, we have a completely contractive inclusion
~
II(Ei ®min E)/U iEI
(II
EdU )
®min
(2.8.9)
E.
iEI
Proof. Since E is finite-dimensional, the two spaces appearing in (2.8.9) are clearly identifiable as vector spaces. Observe that, by (2.6.2), we have an isometric identity
Ei ®min E). ($iEI Ei) ®min E = $( iEI Changing E to MN(E) (N = 1,2, ... ) shows that this is even a completely isometric equality. The quotient mapping q: f ---- f/nu is completely contractive, and hence it defines a complete contraction
q®
Ie:
f
®min
E ---- (f/nu)
®min
E
that clearly vanishes on nU®minE. Passing to the quotient modulo ker(q®IE), we obtain a completely contractive map (f®min
E)/(nu
®min
E)
~
(II
EdU )
®min
E.
iEI
By the preceding observation, f®minE ~ E9 iE I(Ei ®min E) and nU®minE can clearly be identified with the subspace of E9iEI(Ei ®min E) formed of all families that tend to zero along U. In other words, (f ®min E) / (nu ®min E) can be identified with niEI(Ei ®min E)/U. Thus we obtain the proposition. • Remark 2.8.3. Note that by inverting the mapping (2.8.9) we have, for any operator space X, a natural inclusion of the algebraic tensor product:
(II iEI
EdU)
®X C
I1(Ei iEI
®min
X)/U.
We will study in Chapter 17 the conditions on X that ensure that this inclusion extends to an embedding of the minimal tensor product (niEI EdU ) ®min X. Convention. Whenever we are discussing an ultraproduct niEl EdU of a family of Banach spaces or operator spaces, it will be convenient to identify
2. The Minimal Tensor Product. Ruan's Theorem. Basic Operatio1ls
63
abusively a bounded family (Xi)iE! wit.h X; E E; for all i in I wit.h the corresponding equivalence class modulo U, which it determines in D;E! E;/U. Thus, when we speak of (Xi)iE! as an element of DiE! E;/U, we really are referring to the equivalence class that it det.ermines. This abuse is consistent with the one rout.inely done in standard measure t.heory. Remark 2.8.4. Let (Eo:)o:E! and (Fo:)nE! be families with En = E and Fo: = F for all a in the index set I equipped with an ultrafilter U. Let E and F be the associated ultraproducts (or ultrapowers). Let
----+
F**.
Note in passing that iFT = T**iE' Let (EO:)O:E! be a net of subspaces of E directed by inclusion and such that Uo: E! Eo: = E. We can then still define
The Banach space analog of this is well known.
2.9. Complex conjugate Let E be a Banach space. We will denote by E the complex conjugate of E, that is, the vector space E with the same norm but with the conjugate multiplication by a complex scalar. We will denote by x ----+ x the identity map from E to E. Thus, the space E is anti-isometric to E. To be more "concrete," if an operator space E is given as a collection of bi-infinite matrices {(aij)} (representing operators acting on (2), then E can be thought of as the space of all operators with complex conjugate matrices
{(aij)}' It is easy to see that B(H) can be canonically identified with B(H), so that the embedding E
c
B(H) = B(H)
64
Introduction to Operator Space Theory
allows us to equip E with an operator space structure. Moreover, if E is a C* (resp. von Neumann) subalgebra, then so is E c B(H).
c
B(H)
It is easy to see that the corresponding norm a on K 0 E is characterized by the following identity: n
'r:/ai E K
'r:/Xi E E
a(" ~ ai 0 1
n
Xi) =
II" ~ ai 0
Xi·Ii. E' IC "" 'CIllllll
(2.9.1)
1
In other words, the operator space structure of E is precisely defined so that K 0 m in E is naturally anti-isometric to K 0 m in E. For each matrix (aij) in Mn(E), we simply have
Note, however, that there are examples (see [Bou2]) showing that E and E may fail to be isomorphic Banach spaces. Moreover ([C03]), there are examples of von Neumann algebras E that fail to be C*-isomorphic to E (actually, the example in [C03] has trivial center, Le., is a factor). Proposition 2.9.1. Let H, K be Hilbert spaces. We denote by S2(K, H) the space of all Hilbert-Schmidt operators x: K ---+ H and we denote its HilbertSchmidt norm simply by Ilx112. Consider finite sequences (ai) in B(H) and (b i ) in B(K). Then we have
Proof. Note that B(H) 0 m in B(K) has the norm induced by B(H 02 K), and if we identify H 02 K with S2(K, H), then the operator E ai 0 bi can be identified with x ---+ E aixbi . •
2.10. Opposite Let u: Hi ---+ H2 be a bounded linear mapping between Hilbert spaces. In this section, we will denote by tu: Hi ---+ Hi the tmnsposed mapping. Note that the dual spaces Hi and H;' are Hilbert spaces that can be canonically (Le., in a basis free manner) identified respectively with the conjugates Hi and H 2. Let E C B(H) be an operator space. The opposite is the operator space EOP associated to the same space E but equipped with the following norms on Mn(E). For any (aij) in Mn(E) we define
2. The Minimal TellSor Product. Ruan's Theorem. Basic Operations Equivalently, for any
ai E
IC,
IILai Q9Xill
Xi E
65
E we have
K®llIinE01}
=
IILtai Q9Xill
JC®lIlill E
.
In other words, EOP can be defined as the operator space structure on E for which the transposition: X - 4 tx E B(H*) defines a completely isometric embedding of EOP into B(H*). Let A be a C*-algebra. In that case, note that both A C B(H) and AOP C B(H*) are C*-subalgebras and, moreover, AOP and A are C*-isomorphic; hence they are completely isometric. Indeed, the involution x -4 x* is an antilinear isomorphism between A and AOP and hence a linear one between A and AOP. Note however that this is special to C*-algebras: When E is a general operator space, we have EOP 'f!. E, for instance, Rand R are completely isometric but ROP is completely isomorphic to C, and hence not to R.
2.11. Ruan's Theorem and quantization Ruan's Theorem illustrates very well the idea of quantization (see [E2]). This usually refers to the basic idea of quantum mechanics t.o replace (real or complex) numbers by operators (hermitian or not) on a Hilbert space. Here our field of scalars is replaced by IC (or by B(f2)) and t.he product is replaced by the tensor product. Seen through this light, the formulas become quite familiar. While a Banach space structure can be summarized by a norm on a vector space E ~ CQ9E, an operator space can be summarized by a norm (verifying Rl and R 2 ) on IC Q9 E. Let I be any set. Given a vector space V, we denote by Vel) the space of all finitely supported families (Vi)iEI in VI. Let (ei )iE I be an algebraic basis for E. If we identify E to C( 1), then
which shows that the elements of IC play the role of the "scalar coordinates" of a vector: The norm of E is defined on the linear combinations L::A;ei iEI
(Ai)
E C(I)
66
Introduction to Operator Space Theory
while the norm
0:
is defined on the tensors
Seen with this interpretation, the various identities concerning operator spaces become much more transparent: For all the formulas defining the dual, the quotient, the interpolated space, and so on, the same observation is relevant. Let us illustrate this principle for the dual: Let E c B(H) be an operator space. Suppose E is finite-dimensional to simplify. Let el, ... , en be a basis for E. For all coefficients al, . .. , an in K let
(2.11.1)
Then, by Ruan's Theorem, there are operators T}, ... , Tn in B(H) (for H = £2) such that, for all ai in K,
and the dual operator space E* can be identified to span[T}, ... , Tn] C B(H) ([ER2, BPI]). Actually, (2.11.1) remains valid in arbitrary dimension: Let (ei)iEI be a linear (Le. algebraic) basis for an operator space E and let (eniEI be the biorthogonal system in E*. Then, we can write for all "coefficients" (ai)iEI, ai E K with only finitely many nonzero terms,
II L
ai
0 e; IIK®lUlnE'
= sup{11
L ai 0 billK®llIillK I bi
E
K,
IILbi0eiIIK®rnIIlE::; I}.
(2.11.2)
Moreover, if we wish, we can replace K by B(H) in (2.11.1) and (2.11.2). For example, the o.s. structure of Rand C is described by the following formulas (which, by Exercise 2.1.4, are dual to each other in the sense of (2.11.2)) established in Remark 1.13:
IILXi 0 elillK®mlnR = IIL Xix ;11 2 IILXi 0eilllK®mlnC = IILX;Xill~/2. 1
/
2. TIle Minimal Tensor Product. Ruan's Theorem. Basic Operations
67
Note that these two expressions are not equivalent, which reflects the fact that Rand C are not completely isomorphic (compare with (1.3)). In particular, note for the record the following inequality (see Exercise 2.1.4) "Ix;, Yi E B(H) (1 ~ i ~ n)
II LX; ® Yillmin
~ II L xix:1I1/211 Ly:y;ll i / 2.
(2.11.3) This follows directly from (1.11) using ai = Xi ® I, bi = I ® Yi. We note in passing that (2.11.3) implies VXi E B(H) (1 ~ i ~ n)
II LX; ®X;II~i~' ~max{11 Lx;x:ll i / 2, II
L x:x;11
1/
2}.
(2.11.4)
2.12. Universal objects
Every Banach space E embeds into a space of the form £00 (1) for some I, for instance, with I = BE', When E is separable we can take I = N. The operator space counterpart is obvious: Every o.s. E embeds into B(H), and if E is separable, we can take H separable. l\Jore precisely, we have Proposition 2.12.1. Let M = EBn>i AIn- Then any separable operator space embeds completely isometricallyinto M. Afore generally, for any operator space E, tllere is a family of integers (n;)iEI (possibly n'itl1 repetitions) SUell that E embeds completely isometrically into EB;EI Alni . Proof. Let I be the collection of all maps v: E ---+ Afn ,. with 111'llcb ~ 1. We define J: E ---+ EBvEI M n ,. by J(x) = EBvEI v(x). Then, by (2.1.6) and (2.6.2), J is a complete isometry. Now, when E is separable, eaeh space CB(E, Jl1n) = Aln(E*) is weak-* separable, and hence the supremum can be restricted a countable collection I, so we can assume I = N. Finally, adding zero entries whenever necessary, we can assume niH> n; for all i E N. Thus we obtain an embedding into M. • Remark. In the Banach space case, every separable space embeds into C[O, 1], which itself is separable. Moreover, any n-dimensional space can be 1 + c-embedded into £~ for some N = N(c, n) suitably large (see Exercise 2.13.2). The analogous results for operator spaces fail dramatically: No single separable o.s. contains all of them, and the possibility of (1 + c)-embedding a space E into lv!N for some N is very restrictive. This is related to the notion of exactness that will be studied in Chapter 17. The situation for quotients is a bit simpler: At least we do have a separable universal object. In the Banach space case, every Banach space E is isometric to a quotient of i 1(1) for some set I, with I = N in the separable case. The o.s. analog is as follows.
68
Introduction to Operator Space Theory
Proposition 2.12.2. Let M* = f1({M~ I n 2': I}). Then every separable operator space is completely isometric to a quotient of M*. More generally, for every o.s. E, there is a set I and a family of integers (ni)iEI so that E is a quotient Off1({M~i liE I}). Proof. Let I be the collection of all possible maps v: M~" ---+ E with IIvll cb ::; 1. Let XI = f1({M~; liE I}). We define q: XI ---+ E by q((~V)VEI) = LVEI v(~v)' Then IIqllcb ::; 1. By construction, for any n, any x E Mn(E) = CB(M~, E) admits a lifting with the same norm in CB(M~, XI). Hence q is a complete metric surjection. (Alternatively, a simple argument shows that the adjoint J = q* is a complete isometry.) Hence q is a complete surjection onto E, so that E ~ XI! ker(q). Again, when E is separable we can restrict to I countable and niH> ni, and hence E is a quotient of M.. • The next corollary explains why 8 1 is viewed as the operator space analog of fl. Corollary 2.12.3. Every separable operator space is completely isometric to a quotient of 8 1 = (K)*. Proof. Since Co ( {Mn I n 2': I}) is a subspace of K, by Exercise 2.6.2 and by (2.4.3), M* = f1 ({M~ I n 2': I}) is a quotient of K* = 81. so this follows from the preceding statement. • 2.13. Perturbation lemmas
We end this chapter with several simple facts from the Banach space folklore that have been easily transferred to the operator space setting. We start by a well-known fact (the proof is the same as for ordinary norms of operators). Lemma 2.13.1. Let v: X ---+ Y be a complete isomorpllism between operator spaces. Tllen clearly any map w: X ---+ Y with Ilv - wllcb < Ilv- 1 11;;-bl is again a complete isomorphism, and if we let ~ = IIv - wllcbllv-11Icb, we have
Recall that the c.b. distance between two n-dimensional operator spaces E 1 , E2 is defined as follows:
where the infimum runs over all possible complete isomorphisms w: E1
---+
E2•
2. Tlle Alinimal Tensor Product. Ruall's Theorem. Basic Opera.ti01ls
(j9
Lemma 2.13.2. Fix 0 < c < 1. Let X be an operator space. Consider a biorthogonal system (Xi,Xi) (i = 1,2, ... ,n) with Xi E X, xi E X* alld let YI, ... , Yn E X be such that
L
Ilx; 1IIIxi - Yi II < c.
Tl1en tl1ere is a complete isomorphism w: X
--+
X sucll that w(x;) = Yi and
Ilwlleb :::; 1 + c In particular, if El = span(xI,"" xn) and E2 = spall(YI, ... , Yn), we have
Proof. Recall (Proposition 1.10 (ii)) that any rank one linear map v: X --+ X satisfies IIvll = Ilvlleb. Let 8: X --+ X be the map defined by setting 8(x) = LXi(X)(Y; -x;) for all x in X. Then 11811eb:::; L IlxilillYi -xiii < c. Let·w = 1+8. Note that w(x;) = Yi for all i = 1,2, ... , n, Ilwlleb :::; 1 + 11811eb :::; 1 + c, and by the preceding lemma we have Ilw- 1 1leb :::; (1 - c)-I. • Corollary 2.13.3. Let X be any separable operator space. Then, for any n, the set of all the n-dimensional subspaces of X is separable for the distance associated to deb. Proof. Let (xI(m), ... ,xn(m)) be a dense sequence in the set of all linearly independent n-tuples of elements of X. Let Em = span(xl (m), ... , xn(m)). Then, by the preceding lemma, for any c > 0 and any n-dimensional subspace E c X, there is an m such that deb(E, Em) :::; 1 + c. • Lemma 2.13.4. Consider an operator space E and a family of subspaces Eo: c E directed by inclusion and such that UEo: = E. Then, f£r any c > 0 and any finite-dimensional subspa.ce SeE, there exists 0' alld S C Eo: SUell tllat deb (S, S) < 1 + c. Let u: FI --+ F2 be a linear map between two operator spaces. Assume that u admits tlle factorization FI ~E~F2 with c.b. maps a, b sucll tllat a is of finite rank. Tllen for eadl c > 0 tl1ere exists 0' and a
factorization FI ~Eo:lF2 witll Iiallebllbileb < (1 finite rank.
+ c)llallebllbll eb
and
a of
Proof. For the first part, let Xl, ... , Xn be a linear basis of S and let xi be the dual basis extended (by Hahn-Banach) to elements of E*. Fix c' > O. Cho~e 0' large enough and Yl, ... ,Yn E Eo: such that L IIx; 1IIIxi - Yi II < c'. Let S = span(yl, ... , Yn). Then, by the preceding lemma, there is a complete
70
Introduction to Operator Space Theory
isomorphism w: E --+ E with Iiwllcbliw-11lcb < (1 + c;')(l - C;,)-l such that w(S) = S c Eo. In particular, dcb(S, S) ::; (1 + c;')(l - C;')-l, so it suffices to adjust c;' to obtain the first assertion. Now consider a factorization Fl ~E~F2 and let S = a(Fl)' Note that S is finite-dimensional by assumption. Applying the preceding to this S, we find a and a complete isomorphism w: E --+ E with Ilwlicbllw-llicb < 1 + c; such that w(S) C Eo. Thus, if we take a = wa: Fl --+ Eo and b = bWIE~' we obtain the announced factorization. • Exercises Exercise 2.13.1. Let 0 < c; < 1. Let {Xl"",XN} be an c;-net in the unit sphere of a finite-dimensional Banach space E. Let u: E --+ F be a linear map into another Banach space with Ilull ::; 1. Assume that c;' ~ 0 satisfies c; + c;' < 1 and IiU(Xi)11 ~ 1 - c;' Vi=1,2, ... ,N. Then
liu(x)11
~ (1 - c; - c;')lixii
VxE E,
and u defines an isomorphism from E to u(E) such that Ilu- 1: u(E) (1- c; - C;,)-l. Consequently, d(E, u(E)) ::; (1- c; _ C;,)-l.
--+
Ell ::;
Exercise 2.13.2. Let E be a finite-dimensional Banach space. Show that, for any 8 > 0, there is an integer N and a subspace E c e! such that d(E, E) ::; 1 + 8. Exercise 2.13.3. Let A be the (commutative) C*-algebra of all continuous functions on a compact set K. Show that there is a net of maps u a : A --+ A admitting factorizations of the form A~e~~A with Ilvoliliwoll ::; 1 that tend pointwise to the identity on A.
Chapter 3. Minimal and Maximal Operator Space Structures
Let E be a normed space. Then each linear embedding of E into B(H) with H Hilbert defines an operator space structure on E. If the embedding is isometric (Le., preserves the original norm of E), then we will say (in this section only) that the o.s.s. is admissible. Note that the associated norm 0: on IOl)E then satisfies, for any a E K and any e E E, a(a®e) = Ilallllell. Blecher and Paulsen observed that the set of all admissible o.s.s. on a given normed space E admits both a minimal and a maximal element. The minimal one is easy to describe: Simply embed E isometrically into a commutative C*algebra C (for instance, we can take C = C(T) the algebra of continuous functions on the compact set T = (BE" a(E*, E))). Let us denote by min(E) the resulting operator space. By Proposition 1.lO(ii) , min (E) does not depend on the choice of C. Moreover, by (1.9), we have V(aij) E Mn(min(E))
Equivalently, the norm O:min on K ® E associated to min(E) coincides with the injective tensor norm (see Remark 1.11) on K ® E. Clearly (see Proposition 1.lO(ii)), for any operator space F, any linear map u: F --+ E satisfies lIu: F
--+
min(E)lIcb
= lIull·
(3.2)
The maximal tensor product can be described as follows: Let I be the collection of all maps u: E --+ B(Hu) with lIuli :::; 1. (Warning: A "collection" is not necessarily a set. Throughout these notes, we deliberately ignore this set theoretic difficulty wherever it appears since it is obvious how to fix it.) Consider the embedding j:
E EeB(Hu)C B(EeHu) --+
uEI
uEI
defined by VxEE
j(x)
= Eeu(x). uEI
Then (as was explained in §2.6) this embedding (which clearly is isometric) defines an admissible o.s.s. on E. We denote by max(E) the resulting operator space. We have, by (2.6.2), V(aij) E Mn(max(E)) II(aij)IIM.. (max(E»
= sup{ll(u(aij))IIM.,(B(H
u ))
I u:
E
--+
B(Hu ),
lIuli :::;
I}. (3.3)
72
Introduction to Operator Space Tlleory
Clearly, for any operator space F, any linear map u: E Ilu: max(E)
--->
--->
F satisfies
Fllcb = lIull·
(3.4)
Let O:max be the norm on K ® E, corresponding to the operator space max(E) . (Warning: It is not the projective norm in Grothendieck's sense!) For any admissible o.s.s. on E, the identity on E defines completely contractive maps max(E) ---> E ---> min(E). In other words, the set of norms 0: on K ® E satisfying Axioms (R 1) and (R 2 ) and (say) o:(eu ® e) = Ilell Ve E E admits a minimal element O:min and a maximal one O:max and O:min ::; 0: ::; O:max· We have completely isometrically (see [BPI, B2]) min(E)* = max(E*)
and
max(E)* = min(E*).
(3.5)
For the proof, see Exercise 3.2. In particular, we have for any set I and any measure space (0, JL)
where £00(/) or Loo(O, JL) are equipped with their o.s. structures as commutative C* -algebras. In [Pa2]' Paulsen observed that the norm of the space Mn(max(E)) can be described as follows: Theorem 3.1. Let x E Mn(E). Tllen IIxIlM.,(max(E» < 1 iff x admits for some N a factorization of tlle form x = 0:0 Do: 1 , wllere 0:0, 0:1 are (rectangular) scalar matrices of size n x Nand N x n, respectively, and wllere DE AfN(E) is a diagonal matrix sucll tllat 110:011 IIDII 110:111 < 1. Proof. Let Illxlll n = inf{IIO:oIIIIDllllo:dl}, where the infimum runs over all possible factorizations as above. It is not hard to check that Axioms (R 1 ) and (R 2 ) hold in this case, so that, by Ruan's Theorem (see §2.2), these norms come from an operator space structure on E. Let E be the resulting operator space so that IIxIlM,,(E) = Illxllln- By examining the case n = 1, ~e see that
E is isometric to E. Moreover, by the very definition of III Illn' it is easy to see that the identity defines a complete contract~n from E to max(E) , but then by the maximality of max(E) we conclude E = max(E) completely • isometrically. A variant of the preceding argument yields the following:
3. Minimal and Maximal Operator Space Structures
73
r
be an arbitrary set, let f 1 (r) be the classical = max( e1 (r). We denote by Eo C E tlw linear span of the canonical basis vectors (e"Y )"YEr ill E. Consider x in Mn(Eo). TllenllxIlM,,(E) < 1 ijfx admits (for some integer N) a factorization of the form x = aODal, wllere ao, al are scalar matrices of size n x Nand N x n, respectively, and where D is a diagonal matrix with entries of the form (e"Yl' ... , e"YN) for some N -tuple '/'1, ... ,'/'N in r.
Theorem 3.2. ([Pa2}) Let
eI-space over the index set r, and let E
Naturally, we will say that an operator space is "minimal" (resp. "maximal") if min(E) = E (resp. max(E) = E) completely isometrically. A minimal (resp. maximal) operator space E is characterized by the following property: "IF "Iu : F ---; E (resp. "Iu : E ---; F)
Ilulicb
= lIull.
Indeed, this property clearly holds if E is minimal (resp. ma..ximal), and the converse is easy to show: Take F = min(E) (resp. F = max(E» and let
u=IE · Obviously, minimality passes to subspaces, but it need not pass to quotients (s('e, e.g., §9.1). On the other hand, maximality does not pass in general to subspaces, but it does to quotients, as follows. Proposition 3.3. Any quotient of a maximal operator space is itself maximal. More precisely, let E = max(E) be a maximal operator space and let SeE be a closed subspace. Then the following completely isometric identity holds: max(EjS) = max(E)jS. In particular, a.ny maximal operator space E = max( E) is completely isometric to a quotient of max(f 1 (1» for some set I, and if E is separable, we can take I = No Proof. Let F be an arbitrary operator space and let u: E j S ---; F be an arbitrary map. \Ve denote by q: E ---; E j S the canonical surjection. Clearly, it suffices to show that Ilu: max(EjS) ---;
Fllcb =
Ilu: max(E)jS ---;
Fllcb.
(3.6)
But the left side of (3.6) is equal to lIull and, by (2.4.2), the right side is = Iluqllcb = Iluqll = lIull· Thus we indeed have (3.6). Since any Banach space is isometric to a quotient of £1 (1), with I = N in the separable case, the last assertion is clear. • Let E be a Banach space. Recall we denote by IE the identity mapping on E. Paulsen [Pa2] introduced the constant
a(E)
=
IIIE: min(E) ---; max(E)lIcb.
74
Introduction to Operator Space Theory
It is easy to show (exercise) that a( E) is equal to the supremum of the ratio when u runs over all possible maps u: E ---+ F with E isometric to E and F arbitrary. The finite-dimensional case is quite interesting:
lIullcb/llull
Proposition 3.4. ([Pa2J) Let E = space. Then
(en, II
liE) be allY n-dimensional normed (3.7)
wllere the supremum runs over all sets al, ... , an E B(H), bI , ... , bn E B(K) SUell that
(H, K arbitrary Hilbert spaces)
and Vt E
en
Proof. By the definition of a( E), we have a(E)
= supsup{llxIlM",(max(E» IllxIlMm(min(E» :::; I}.
(3.8)
m
Consider x = (Xij) E Mm(E) with IIxIlMm(min(E» :::; 1. Let {ed be the canonical basis of E = en and let ~k E E* be the dual basis. Let bk = Ei,j eij ® ~k(Xij) E Mm· By (3.1), we have sup {IILtkbkIIM",
I(tk)
E
en,
IILtk~kIIE. :::; I} = IIxIlMm(min(E)) :::; 1. (3.9)
On the other hand, by (2.6.2) and the definition of max(E), we have IIxIlMm(max(E» = sup IIvll:51
IILe
i .1
®v(xij)11
M",(B(Ht ,»
.
(3.10)
Now fix v: E ---+ B(H) with IIvll :::; 1 and let ak = v(ek). Note that, for any in en, we have IIEtkakl1 :::; IIvlllltllE :::; IltiIE. Moreover,
IlL eij
® v(xij)11 =
= IIEbk®akIlM",(B(H» =
E eij~k(Xij) ® v(ek)
i,j,k
IIEak®bkIlB(H)®mlnM",'
Combining this with (3.8), (3.9), and (3.10), we obtain
t
3. l\finimal and l\,faximal Operator Space Structures
75
with the supremum as in (3.7). Moreover, we have equality when the supremum is restricted to J( finite-dimensional. Then, invoking (2.1. 7), we obtain the equality in (3.7), as announced. • Remark. By (2.1.7), the supremum is the same in (3.7) if we restrict Hand both to be finite-dimensionaL
J(
Corollary 3.5. ([Pa2}) For any n-dimensionalllormed space E, we llmre neE)
= o(E*).
(3.11)
Note that this also can be deduced from (3.5) and (2.3.3). We now return to the number o(n) considered at the end of Chapter 1. Proposition 3.6. ([Pa2J) Let o(n)
where the supremum
= sup {I//~~t HillS
I u:
E
-4
F, rk(u) ::::;
n} ,
over all possible operator spaces E, F. Tllfm
o(n) = sup{o(G) / dimG::::; n}.
(3.12)
The proof is left as an exercise. To majorize
0: (n),
we will need the following classical lemma.
Lemma 3.7. (Auerbacll's Lemma) Let E be all arbitrary n-dimensional Ilormed space. There is a biorthogollal system Xi E E, ~i E E* (i = 1,2, ... , n) Stich that I/xJ = I/~il/ = 1 for all i = 1, ... , n.
.
Proof. Choose XI, ... ,X n in the unit sphere of E on which the function -4/det(xl, ... ,Xn )/ attains its maximum, supposed eql\al to C > o. Then let ~i(Y) = C- I det(xI, ... , .'1:i-l, y, Xi+I, ... , xn). The desired properties are
X
~~~.
Theorem 3.B. ([Pa2J) For allY n 2: 1, we llave
n/2 ::::;
o(e~)
::::; (t(n) ::::; n
and
(n/2)1/2 ::::; o(e~) ::::;
-.Iii.
Proof. We first show that o(n) ::::; n. Consider E with dim(E) = n. Let (Xi'~i) be as in the preceding lemma. We have Je(x) = L~ Ui(X) with
76
Introduction to Operator Space Theory
Ui(X) = ~i(X)Xi' Hence o:(E) = IIIe: min(E) -+ max(E)llcb ::; 2::~ Ilui: min(E) -+ max(E)llcb. But since Ui is of rank::; 1, we have (see Proposition 1.1O.(ii)) Iluilicb = lIuill = IIxill II~ill = 1, whence o:(E) ::; n, and by (3.12) o:(n) ::; n. This bound is improved later in Theorem 7.15. We now turn to the converse estimate. Let (U l , ... ,Un) be a "spin system," that is, an n-tuple of unitary self-adjoint operators such that
A simple calculation shows that, for any
x in en, we have
Indeed, let T = 2:: XiUi; we then have T*T IITI12 ::; 22:: l~iI2.
+ TT* =
22:: IXi 12 . I, and hence
Moreover, it is easy to see that such a system generates a finite-dimensional C*-algebra (indeed the linear span of all the 2n products Uit Ui2 .. .uik with il < i2 < ... < ik is a C* -algebra). Therefore, such a system can be realized in B(H) with dim(H) < 00 (actually with dim(H) = 2n). From this, we will deduce the following claim:
(3.13)
Indeed, let (e p ) be an orthonormal basis of H. Let t = 2:: ep 0 ep be the tensor in H 02 H corresponding to the identity of H. Since Uk is self-adjoint and unitary, a simple calculation shows that Uk 0 tUk(t) = t, and hence 2::~ Uk 0 tUdt) = nt, whence 112::~ Uk 0 tUkl1 2:: n. The converse is obvious, whence the claim (3.13). Applying (3.7) with ak = tb k = 2- l / 2Uk, we obtain
n/2 ::;
o:(e~)
(3.14)
and a fortiori n/2 ::; o:(n). We now turn to the second line. The lower bound follows again from (3.7) with ak = Uk/(2n)l/2 and bk = tUk. The remaining upper bound is left as an exercise (Hint: Use, e.g., the following factorization of the identity min(er) -+ en -+ max(er), where the first arrow has c.b. norm = 1 and the ~~~=~.
•
The next result comes from [Pa2] for dim(E) > 4 (we could include dimensions 2 and 3 thanks to exercise 3.7).
3. Minimal and IIfaximal Operator Space Structures
77
Corollary 3.9. Any infinite-dimensional Banach space E satisfies o:(E) l\.[oreover, we have o:(E) > 1 as soon as dim(E) > 2.
=
00.
Proof. Observe that, if F is another Banach space, we have obviously
o:(F) :S d(E, F)o:(E).
(3.15)
In particular, 0:(£2) :S d(E,£2)0:(E). By a well-known result due to F. John (see, e.g., [P8] or [T.ll]), if dim(E) = n, then d(E, £2) :S Vii. Thus, together with Theorem 3.8, this gives us that any n-dimensional space E satisfies Vii/2 :S o:(E). In particular, o:(E) > 1 whenever n > 4. Using Exercise 3.7, we can extend this n > 2. When E is infinite-dimensional, we use a variant of F. John's Theorem: For any n, there is a subspace En C E with dim(En ) = n and an isomorphism u: En ---7 £2 with Ilu-111 :S 1, IIuli :S Vii but, moreover, such that u admits an extension u: E ---7 £2 with IIuli :S .;n. Then again we claim that o:(E) ~ 0:(£2)/Vii. Indeed, the identity min(£;) ---7 max(e 2 ) factors as -1
-
min(£2)~ min(E)~ max(E)~ max(£2)' which implies that (recalling (3.2) and (3.4)) 0:(£2)
= III: min(£2)
Letting n go to
00,
---7
max(C2)IIcb:S IIu-Illo:(E)IIuli :S foo:(E).
we obtain o:(E) =
00
by Theorem 3.8.
•
Remark. By Corollary 3.9, any infinite-dimensional Banach space E can be embedded isometrically into some B(H) so that there is a bounded lllap (and actually an isometric one) u: E ---7 B(H) that is not c.b. There are some interesting questions left open in the dimensions n = 2,3. or 4 that are aimed at measuring how small (or how large) the set of all admissible o.s.s. on E can be. For instance, what is exactly its diameter? Which Banach spaces E admit a unique admissible operator space structure? Equivalently, for which E do we have o:(E) = I? By Corollary 3.9, the dimension of such an E must be :S 2, but the only known examples of dimension> 1 are those of the following result: the spaces £~ and £r (i.e. the two-dimensional versions of £00 and Ct). Are these the only examples? Proposition 3.10. {Pa2}. There is a unique operator space structure (respecting the norm) on the spaces £r and £~. Equivalently, we have o:(£r) = o:(£~) = 1. Proof. Let el, e2 be the canonical basis of £r. The space min( £r) can be realized in C(8D) (here D = {z Eel Izl < I}) as the linear span of the pair [1, z], so that we have for any pair (XI, X2) in Jl,fn : IIXI
® el
+ X2 ® e211M,,(min(en>
=
sup
zE8D
Ilxl + zX2IiM".
(3.16)
78
Introduction to Operator Space Theory
On the other hand, note that the data of a contraction u: £~ --+ B (H) boil down to that of a pair of contractions Tb T2 in B(H) with Ti = u(ei)' Therefore we have
where the supremum runs over all H and all pairs (Tb T2 ) of contractions in B(H). Since any contraction is in the closed convex hull of the unitaries (by the Russo-Dye Theorem [Ped, p.4l), the last supremum remains the same if we restrict ourselves to pairs of unitary operators (TI' T 2 ). But then we may multiply by Til (say on the left), and we find that we can restrict ourselves further to pairs of the form (1, T) with T unitary. Then, the pair (1, T) lies in a commutative C*-algebra that can be identified with C(a(T», a(T) being the spectrum of T, and a(T) c aD. Thus we have clearly Ilxl ® 1 + X2 ® Til = sup Ilxl zEq(T)
+ ZX2 II JIl"
sup IIxl
:::;
zE8D
+ zx21IJ1l ... ·
Recalling (3.16) and (3.17), the previous discussion shows that II min(£n --+ max(£nllcb = 1, whence o:(£~) = 1. By (3.11), we have o:(£~) = 1. • Remark. In [Pa2], Paulsen actually shows that
max{ yn, n/2} :::; o:(£~) :::; n/h.
(3.18)
(n/2)1/2 :::; o:(£~) = o:(£n :::; min{ yn, n/2},
(3.19)
and Note that, by (3.15), 0:(£2) :::; yno:(£~). That v'n :::; 0:(£2) follows easily for instance from (1.5). The upper bound 0:(£2) :::; n/h follows from Proposition 3.10. Indeed, let (ak) and (bk) be as in Proposition 3.4 with E = £2' By (3.15), we have 0:(£2) :::; J2. Hence for any k 1:- £ we have and, on the other hand, 2(n - 1)
L ak ® bk = L ak ® bk + al ® blj k#l
hence by the triangle inequality 2(n -1) IIEak ®bkll :::;
E lIak ®bk +al®alll k#
:::; n(n -1)h,
3. l\Iinimal and l\Jaximal Operator Space Structures
79
and thus we obtain IIL:ak ® bkll ::; n/V2. Hence, by Proposition 3.4, o:(€~) ::; n/V2. A similar argument shows that 0:(t'1) ::; n/2. Finally, the bound 0:( t'f) ::; ..;n follows from (say) the factorization min(t'l) ~ en ~ max(t'1) defined by v(ei) = en and w(en) = e;, since it is easy to check that Ilvllcb = 1 and IIwllcb = ..;n. Paulsen [Pa2] eonjectures that o:(t'~) = n/V2 and o:(t'~) = In/2, but (contrary to this) we believe that the lower bound appearing in Exercise 3.7 is sharp. We refer the reader to Paulsen's paper [Pa2] and to the work by Arias, Figiel, Johnson, and Schechtman [AF JS] for more on this theme. Open problems 1. Are the spaces C, £?, and t'~ the only operator spaces E -I- {O} admitting a unique operator space structure, that is, for which o:(E) = 1 ? 2. What is the value of o:(t'~) for n > 2? What about o:(t'~) for n > 2? Note that, by (3.18) and (3.19), we have o:(t'~) = V2 and O'(€~) = 1. lVIore importantly: 3. What are the values of: lim sup 0'( t'~) / n 1/2? n---+oo
limsupO'(t'~)/n? n---+oo
lim sup o:(n)/n? n---+oo
4. What are the n-dimensional normed spaces E such that o:(E)
= o:(n)?
We discuss more open problems concerning maximal operator spaces and their subspaces in Chapter 18. Exercises Exercise 3.1. (Principle oflocal reflexivity [LiR].) Let E, G be Banach spaces v
v
with dim(G} < 00. We then have (E ® G)** = E** ® G (isometrically). As a consequence, for any v: E* --+ G there is a net of weak-* continuous maps Va: E* --+ G with Ilvall ::; Ilvll that tend to v in the topology of simple convergence. Exercise 3.2. Let E be an arbitrary Banach space. Prove that min(E*) = max(E)* and max(E*) = min(E)* ([BPI]). Exercise 3.3. Show that E is minimal (resp. maximal) iff its bidual E** is also ([BPI]). Exercise 3.4. Show that the direct sum of a family of minimal spaces is again minimal. Exercise 3.5. Let {Ei liE I} be a family of maximal operator spaces (for instance, we could have Ei = C for all i in 1). Show that €1 ( {Ei liE I}) is again a maximal operator space.
Introduction to Operator Space Tlleory
80
Exercise 3.6. Let U1 , ••• , Un be a spin system. Show that the spaces E
= span{Ui
® Ui
11 :::; i
:::; n}
and
F
= span{Ui
® Ui
11 :::; i
:::; n}
are minimal operator spaces. Exercise 3.7. Let Ci be the creation operators (i = 1, ... , n) on the antisymmetric Fock space (see §9.3) associated to e~. Show that IIE~ XkCk11 = (E IXk 12) 1/2 for any x in en and deduce that
Show that and that if n is even. Exercise 3.8. Let X be a maximal o.s. Let XI C X be a separable subspace. Show that there is a subspace X 2 with XI c X 2 C X that is separable and maximal.
Chapter 4. Projective Tensor Product Since the minimal tensor product is analogous to the (Banach space) injective tensor product, it is natural to search for an analog in the operator space setting of the (Banach space) projective tensor product. This question is treated in [BPI] and [ER2] independently. Effros and Ruan went further and considered a version for operator spaces (in short, o.s.) of Grothendieck's approximation property. They proved the o.s. analog of many of Grothendieck's Banach space results, introduced integral operat.ors and absolutely summing operators, and proved a version for o.s. of the Dvoretzky-Rogers Lemma. This program meets several interesting difficulties and leaves open several problems, mostly related to the absence of local reflexivity for operator spaces. We will limit ourselves here to a brief description of the o.s. version of the projective tensor product of two operator spaces E, F, which we will denote by E @" F. The latter space is defined in [BPI] as the natural "predllal" of the space CB(E,F*). The equivalent but more explicit definition of [ER2] is as follows. Let t be an element of the algebraic tensor product E @ F. Of course, t admits a (non-unique) representation of the form t= i,j$.e p,q$.m
where e, m are integers and where x rectangular scalar matrices.
E
Afe(E),
y
E
Mm(F), and a, (3 are
Then the "projective" tensor norm (in the sense of o.s.)
IItllE01\F is defined
as:
IItllE01\ F = inf{ lIaIIHSllxIL'IIdE) IlylLu", (F) 11/31IHS}, where IIIIHS is the Hilbert-Schmidt norm and where the infimum nms over all possible representations (actually, by adding zeroes \ve may restrict attention to the case e= m if we wish). \Ve denote by E @" F the completion of E @ F with respect to this norm. More generally, this space can be equipped with an operator space structure corresponding to the norm defined (for each n) on lIfn (E @" F) as follows: Let t = (t rs ) E Mn(E @ F) and assume t
= a . (x @ y) . (3,
(4.1)
where the dot denotes the matricial product and where x E lIf(E), y E lIfm (F), and a (resp. (3) is a matrix of size n x (em) (resp. (em) x n). Note that x @ y is viewed here as an element of the space lIffm (E @ F). I\Iore explicit ly, we may index [1, ... , mt'] either by a pair (ip) (1 :S i :S e,1 :S p :S m) or by a pair (jq) (1 :S j :S t',1 :S q :S m), so that by dropping the parentheses we may write
Introduction to Operator Space Theory
82
Then (4.1) can be rewritten more explicitly as trs =
L
O:r,ip(Xij
18)
(1 ~ r ~ n, 1 ~ s ~ n).
Ypq){3jq,s
(4.2)
i,p,j,q
Then (following [ER2]) we can define
We then obtain an operator space structure on E
18)11
F. Moreover (see [ER2,
BPI]): Theorem 4.1. We have completely isometrically (E 18)11 F)*
= CB(E, F*).
(4.4)
More generally, for any operator space G, we have completely isometrically CB(EI8)II F,G)
= CB(E,CB(F,G».
(4.5)
Proof. Note that (4.4) is a particular case of (4.5) (take G = q. Therefore we concentrate on (4.5). We will show that (4.5) is an isometric identity. It follows "automatically" (replacing G by Mn (G» that it is completely isometric. Let cp: E 18) F -+ G be a linear map with associated mapping u: E -+ CB(F, G) defined by u(x)(y) = cp(x 18) y). To show that (4.5) is isometric, it clearly suffices to show that I = I I, where 1= IIcpllcB(E®"F,G)
and
II
=
IlullcB(E,CB(F,G»'
By (4.2) and (4.3), we have
I = sup {
[t=
O:r,ipCP(Xij
18)
ypq){3jq,s]
Jpq
} , r,s
/If.. (G)
where the supremum runs over all n and all m, e, 0:, (3, x, y with 110:11.1IJ" ml ~ 1, 11{3I1M",£ ... ~ 1, IIxilllldE) ~ 1, IIYIIMm(F) :::; 1. Let a E Mlm(G) be the £1n x em matrix with entries as follows:
a(i,p;j,q) = cp(Xij 18)ypq). Clearly, the matrix product o:.a.{3 satisfies (4.6)
4. Projective Tensor Prodllct
83
where the supremum runs over all Ct, (3 in the unit ball respectively of Jlfn . l1l € and JlIl1lc .n . In addition, when n = the choice of n, {3 equal to the identity shows that (4.6) is an equality. A fortiori, we have equality whenevpr n ::; tm. Hence we conclude from this that
em
(4.7) where the supremum runs over all C, m and all x, y as before. On the other hand, we have by definition II
= sup{II(IJlle 0
u)(x)II,'11,(CB(F,G»
I C ~ 1, lI:rlllll,(E) ::; I}.
Then, using life (CB(F, G))
= CB(F, Me(G)), we find
11(l1ll, 0
=
u)(x)lllIldCB(F,G»
sup{IIIlIl m 0 [(lM,
o u)(x)](y) II JII." (CB(F,lIldG»)
I m ~ 1, lIyIIJlI",(F)::; I}.
Now again lIfm(CB(F, lIfe(G))) ~ CB(F, MI1l(Me(G)) Moreover, in this correspondence
~
CB(F, lIIlm (G)).
is nothing but the Cm x em matrix with entries
•
Thus by (4.7) we obtain I = I I as announced. Remark. In particlIlar, taking F = CB(E, lIIn ), (4.4) implies
M~
and recalling that Mn(E*)
(4.8) The projective tensor product is a "good" one in the sense that, for any operator spaces E i , F i , if U1: E1 --+ F1 and U2: E2 --+ F2 are c.b., then U1 0 U2: E1 0" E2 ---+ F1 0" F2 is c.b. and we have
The projective nature of this tensor product can be seen through the following property: If B c F is a closed subspace, then we have a completely isometric identification (4.9) E 0" (FIB) = (E 0" F)IN,
84
Introduction to Operator Space Theory
where N is the kernel :)f the natural (surjective) mapping from E ®i\ F onto E ®i\ (FjS). More generally, if in the preceding situation Ul and U2 are complete surjections, then Ul ® U2: El ®i\ E2 ----+ Fl ®i\ F2 is also one. This is fairly easy to check with the definition of E ®i\ F. However, of course, just like in the Banach space case, injectivity fails in general: Ul ® U2: El ®i\ E2 ----+ Fl ®i\ F2 may even fail to be injective when Ul and U2 are complete isometries. In the Banach space case, the projective tensor norm is the largest reasonable tensor norm. An analogous property also holds in the operator space case (see [BPI]). Recall that, following Sakai ([Sa]), one can define a von Neumann algebra as a C* -algebra Al that is the dual of a Banach space, which we call its predual (see §2.5). This predual is unique up to isometry, and we denote it by Al*. It is what is usually called a "noncommutative L 1-space." Actually this terminology is quite abusive, since the commutative case (which corresponds to the standard L 1-spaces) is not really excluded! Let AI, N be two von Neumann algebras equipped with their natural o.s. structure, and let AI* and N* be their duals equipped with the dual structure, as defined in (2.3.1). We can equip AI* and N* with the o.s. structure induced by Af* and N*. One can then show ([ER8]) that the projective tensor product AI* ®i\ N* is completely isometric to the predual of the von Neumann algebra AJ®N generated by the algebraic tensor product M ® N. In other words (see [ER8]): Theorem 4.2. We have a completely isometric identification (4.10)
Proof. By Theorem 2.5.2, we already know that Af®N ~ CB(AI*, N). Thus, the fact that AI. ®i\ N. ~ (AI®N). isometrically can be deduced from (4.4) and the unicity of the predual of M®N. (Actually, since the unicity of the predual is also valid in the completely isometric sense, this argument yields a • proof that AI. ®i\ N. ~ CB(M., N) completely isometrically.) Therefore, if E, Fare "noncommutative Ll-spaces," then E ®i\ F is also one. This is analogous to Grothendieck's classical result ([GrD that the (Banach space) projective tensor product of L 1(f..l) and L1(V) is isometric to an L 1space, namely, to the space L1 (f..l xv). It is possible to check that the natural morphism
E ®i\ F
-+
E ®min F
4. Projective Tensor Prodllct
85
is a complete contraction, but in general it is not injective. Its injectivity is related to the operator space version of the approximation property for E or F, as follows. Following [ER8J, an o.s. E is said to have the OAP if there is a net of finite rank (c.b.) maps Ui: E -+ E such that the net I®ui converges pointwise to the identity on K ®min E. This is the o.s. analog of Grothendieck's approximation property (AP) for Banach spaces. When the net (UI) is bounded in CB(E, E), we say that E has the CBAP (this is analogous to the BAP for Banach spaces). To quote a sample result from [ER8]: E has the OAP iff the natural map E* ®I\ E -+ E* ®min E is injective. The class of groups G for which the reduced C*-algebra of G has the ~AP is studied in [HK] (see also §9 in [Ki8]).
Remark. It should be emphasized that the AP for the underlying Banach space is totally irrelevant for the OAP: Indeed, Alvaro Arias [A2] recently constructed an operator space isometric to €2 but failing the ~AP! Building on previous unpublished work by T. Oikhberg, Oikhberg and Ricard fORi] obtained more dramatic examples of the same nature. In particular, they constructed a Hilbertian operator space X such that a linear map T on X is c.b. iff it is the sum of a multiple of the identity and a Hilbert-Schmidt map, or iff it is the sum of a multiple of the identity and a nuclear map in the o.s. sense (they can even produce finite-dimensional versions of the space X). In particular, every T E CB(X) has a nontrivial invariant subspace. The ideas revolving around the OAP or the CBAP are likely to lead to a simpler and more conceptual proof of the main result of [Sz] (that B(H) fails the AP), but unfortunately this challenge has resisted all attempts so far. We will return to these topics (the OAP and the CBAP) when we discuss "exactness" in Chapter 17. We refer the reader to [BP1, ER2,5,6,8] for more information on all of this.
Chapter 5. The Haagerup Tensor Product Curiously, the category of operator spaces admits a special kind of tensor product (called the Haagerup tensor product) that does not really have any counterpart (with similar properties) in the Banach category. This tensor product leads to a very rich multilinear theory (initiated in [CS2j), the equivalent of which does not exist for Banach spaces. The Haagerup tensor norm was introduced by Effros and Kishimoto [EKJ, who, in view of its previous use by Haagerup in [H3], called it this way. They only considered the resulting Banach spaces; but actually, it is the operator space structure of the Haagerup tensor product that has proved most fruitful, and the latter was introduced in [PaS], extending the fundamental work of Christensen and Sinclair [CS2] in the C*-algebra case. Basic properties. Let E 1, E2 be operator spaces. Let Xl E K ® E 1, X2 E K ® E 2. We will denote by (Xl, X2) --+ Xl 8 X2 the bilinear mapping from (K ® E 1) x (K ® E 2) to K ® (E1 ® E 2) that is defined on rank 1 tensors by (k1 ® et) 8 (k2 ® e2)
= (k1k2)
® (e1 ® e2).
Let us denote for any Xi in K ® Ei
Qi(Xi)
= IIxillK®mlu E i
(i = 1,2).
Recall that Ko denotes the linear span of the system {ei.j} in K. Then, for any X in Ko ® (E1 ® E 2), we define
Qh(X)
= inf
{t Q1(x{)Q2(~)}
,
(5.1)
3=1
where the infimum runs over all possible decompositions of X as a finite sum n
X=LX{8X~
(5.2)
j=l
with
x{
E
Ko ® E 1 ,
x~
E
Ko ® E 2 •
In the particular case when X EEl ®E2 ~ M 1(E 1®E2) ~.Af1 ®(E1 ®E2), this definition means that
where the infimum runs over all possible ways to write X as a finite sum X = L:~ ai ® bi with ai EEl, bi E E2' Hence, if Ei C B(Hi ), we have (cf. Remark 1.13)
5. Tlle Haagerup Tensor Product
87
Notation. Let E be an operator space and let x = L ak i8l ek be in K i8l E (ak E K, ek E E). Then we will denote for any a, bin B(f2)
and also of course
One surprisingly nice property of (5.1) is that it suffices to take n (5.1), that is, we have
=
1 in
(5.3) Indeed, let x{, x~ be as in (5.2). Let Sj: £2 -+ £2 be a sequence of isometries with orthogonal ranges, that is, such that sj S j = I and si s j = 0 for all i 1= j. Then let n
and
X2 =
LSk' x~. k=l
Clearly, we have x = Xl 0 X2 and, moreover, if we embed E; into a ('* -algebra A (completely isometrically), then we have in the C*-algebra K i8l m in Ai, n
0:1 (xt}2 =
IIx1 Xi 11x::®",;uAI
=
L
X{ X{*
(5.4)
L X~*X~
(5.5)
j=l
and similarly n
0:2(X2)2 =
Ilx2X211x::®",;uA2
=
j=l
which in particular yields
Now by a homogeneity trick this clearly yields the identity of the right sides of (5.1) and (5.3), respectively. It is very easy to check that the norm O:h satisfies the axioms of Ruan's Theorem. Hence, after completion, we obtain an operator space denoted by El i8lh E2 and called the Haagerup tensor product. But actually, we do not need Ruan's Theorem here; the fact that E1i8lhE2 is an operator space follows from Theorem 5.1.
88
Introduction to Operator Space Theory Note that, by Remark 2.1.6, we have for any x in /Co ® EI ® E 2 ,
Therefore, we have a completely contractive mapping
By an entirely similar process we can define the Haagerup tensor product of an N-tuple E 1 , .•. , EN of operator spaces. Once again, for any x in /Co ® (EI ® E 2 ••• ® EN), we can define a(x)=inf{al(xI)a2(x2) ... OW(XN)
I
x=xI0 x 2 ... 0xN: XiE/Co®Ed··
(5.6) This coincides with the extension of (5.1). Again this satisfies Ruan's axioms, so that we obtain an operator space denoted by EI ®h E 2 ... ®h EN' The very definition of the norm (5.6) clearly shows that this tensor product is associative, that is, for instance, we have
However, it is important to underline that it is not commutative (Le., EI ®hE2 can be very different from E2 ®h Ed. It is immediate from the definition that EI ®h E2 enjoys the classical "tensorial" properties required of a decent tensor product; that is, for any operator spaces FI, F2 and any c.b. maps Ui: Ei --+ Fi (i = 1,2), the mapping UI ® U2 extends to a c.b. map from EI ®h E2 into FI ®h F2 with
Moreover, this remains valid with N factors instead of 2. The great power of the Haagerup tensor product stems from the fact that it admits two distinct descriptions, a "projective" one, as above, and an "injective" one, as follows. Although we will soon abandon this notation, for any x in /Co ® EI ® E2 of the form say x = L ak ® el ® e~, we define (5.7)
where the supremum runs over all Hilbert spaces 1f. and possible maps O"i: Ei --+ B(1f.) with 1I00ilicb :5 1. This norm clearly corresponds to an operator space structure on EI ® E2 associated to the mapping
L el ® e% EB L>1(el)0"2(e%), --+
0'1,0'2
5. The Haagemp Tensor Prodllct
89
where the direct sum runs over the collection of all possible pairs of complete contractions 0'1: El ----+ B(H) and 0'2: E2 ----+ B(H) on the same Hilbert space. vVe will denote by EI Q9 f E2 the resulting operator space (after completion). We will denot.e by 0'1 . 0'2: EI Q9 E2 ----+ B(H) the linear mapping taking el Q9 e2 to 0'1(ed0'2(e2). Thus, we have
In other words, if we denote
and
where the direct sum runs over all pairs as above, and if we denote by H the Hilbert space on which these are acting, then the mapping
(5.7)' is a complete isometry. 1Ioreover, this definition clearly extends to an arbitrary number of factors E I , ... , EN and leads to an operator space denoted by EI Q9f ... Q9f EN. We will show that, if we have E; C A; (completely isometrically) for some C* -algebra Ai as above, then EI ® f E2 naturally embeds into the /7'ee product Al * A2 of the two C* -algebras, whence our notation. But first we will show that the tensor products El Q9h E2 and EI Q9 f E2 are actually identical, as follows. Theorem 5.1. For allY operator spaces E I , E2 we have
ah = Clf.
Proof. Let x = XI 8·7:2 with .7:; E KoQ9E;, (i = 1,2) and Then, for any pair of complete contractions
x E Ko®(EI Q9E2)'
we clearly have
where the product on the right is the product in the C*-algebra KQ9 m ill B(H). Hence
Introduction to Operator SpRce Theory
90 This proves that
(5.8) To prove the converse, we use the Hahn-Banach Theorem. Let E2)*. We will show that aj(e) :::; ah(e)·
eE (K ® El 0
This clearly yields the converse to (5.8). Assume ah(e) = 1. Then, by (5.4) and (5.5), for all finite sequences (xfh~j~n in Ko ® Ei (i = 1,2), we have
For simplicity, we denote B
= B(f2)
in the rest of this proof.
By a standard application of the Hahn-Banach Theorem (see Exercise 2.2.2) this implies the existence of states It and 12 on B®minAl and B®minA2' respectively, such that, for all Xi in Ko ® E i ,
If we now consider the GNS (unital) representations 7Ti: B ®min Ai --+ B(Hi) associated respectively to It and 12, we find elements 6 E HI, 6 E H2 of norm 1 such that (5.9) Clearly, (5.9) implies the existence of an operator such that
T:
H2
--+
HI with
IITII :::;
1
and hence we find (5.10) Now let ri: B --+ B('H) and Pi: Ai determined by the identities
--+
B('H) (i = 1,2) be the representations
Note that ri and Pi have commuting ranges. From (5.10) we deduce that, for any i,j and any al E Et, a2 E E 2, we have (using eij = eilelj, hence eij ® al ® a2 = (eil 0 at) 8 (elj ® a2)) (5.11)
with kj, fi
E
11. defined by k j = r2(elj)6
and
li = rl(eit}*el.
5. Tlle Haagerup Tensor Prodllct
91
Note that
L
IIkj ll 2 =
L(r2(elj)*r2(elj)6,6) ~
11611 ~
1
j
and similarly
L II£i11 2~ 1.
Let us define 0"2: E2 -> B(H2' Hd by setting 0"2(a2) = Tp2(a2) Va2 E E 2· Now consider an element x in /Co 0 (El 0 E 2) of the form m
X = L eij 0tij i,j=1 By (5.11) we have
~(x) = L(Pl· 0"2)(tij)kj ,£i). i,j Hence (note that we may as well assume HI
= H2
if we wish)
This proves that o:j(~) ~ 1; hence, by homogeneity, we have o:j(~) ~ and the proof is complete.
0:;; (~), •
Remark 5.2. Using the factorization of c.b. maps (Theorem 1.6) it is easy to check that the tensor product El 0f E 2 0/ ... 0f EN is associative. Indeed, using the complete isometry (5.7)', Theorem 1.6 implies that any c.b. map u: El 0f E2 -> B(H) can be written as U(XI 0 X2) = 0"1(Xt}0"2(X2) with 1I0"111cbll0"211cb = Ilulicb. From this it is then very easy to verify, for example, that (El 0/ E 2) 0f E3 = El 0/ E2 0/ E 3. Now, using the associativity of both 0/ and 0h, we can extend Theorem 5.1 (by induction on N) to the case of an arbitrary number N of factors. \Vhellce the following statement. Corollary 5.3
(i) Let E 1 , ..• , EN be arbitrary operator spaces. Then El 0/···0/ EN = El 0h··· 0h EN (as operator spaces). (ii) There are complete isometries lfJ i : Ei -> B(H) on some Hilbert space H such that
is a complete isometry.
92
Introduction to Operator Space Theory
(iii) Consider an element x in /( ® El ® ... ® EN of the form x =
L Ai ® xt ® ... ® xf i
with Ai E /( and xt EEl,'" ,xf E EN. Then we have
II X II 1C 0",
II>
[E! % E2 ... 0h EN 1
{I
~ 'up ~~, ® u (xllu'(xl) ... uN (xi'll mJ ' 1
where the supremum runs over all possible choices of H and of complete contractions 17 1 : El ---+ B(H), ... ,aN: EN ---+ B(H). Multilinear factorization. The next corollary is the fundamental factorization of multilinear maps due to Christensen and Sinclair [CS1]. It was extended to operator spaces in [PaS]. See [LeM1] for an extension with Hand Hi replaced by Banach spaces. Corollary 5.4. Let E l , ... , EN be operator spaces and let u: El ® .. ·®EN B(H) be a linear mapping. The following are equivalent.
---+
(i) The map u extends to a complete contraction from El ®h ... ®h EN into B(H). (ii) There are Hilbert spaces Hi and complete contractions ai: Ei ---+ B(Hi+l, Hi) with HN+1 = H and HI = H such that
Moreover, if El"'" EN are all separable (say) and if dim(H) then we can take Hi = H for all i.
=
00,
Remark 5.5. Recall that, by the factorization Theorem 1.6, if Ei is completely isometrically embedded in a C*-algebra Ai, each ai itself admits a factorization of the form ai(xi) =
Vi 1Ti(Xi)Wi
with II Vi II II Wi II ~ 1 and some representation 1Ti: Ai implies a decomposition
---+
B(iii)' Thus (5.12)
(5.13) with contractive "bridging maps" Ii: iii+1 ---+ iii with iiN+! = H, iio = H. Remark 5.6. If we assume H and all the spaces Et, ... , EN finite-dimensional, it is easy to see that in Corollary 5.4(ii) we can assume (by suitably restricting) that all the spaces Hi also are finite dimensional.
5. The Haagerup Tensor Product
93
'*
Proof of Corollary 5.4. (i) (ii) follows from the factorization Theorem 1.6 using Corollary 5.3 (ii), as in Remark 5.2. Conversely, assume (ii). Then clearly by definition of ®f, u extends to a complete contraction on El ®f' .. ®fEN, but by Corollary 5.3(i) this coincides with El ®h'" ®h EN. The last assertion is immediate since, if dim(H) = 00, any countable direct sum of copies of H can be identified with H. • Injectivity /projectivity. Quite strikingly, the Haagerup tensor product turns out to be both injective and projective, as explicited in the next corollary. Concerning the projective case, we recall that a linear mapping u: E --+ F between Banach spaces is called a metric surjection if u is surjective and the associated map it: E /ker( u) --+ F is an isometry. When E, F are operator spaces, recall (see §2.4) that u is a complete metric surjection iff h®u: K®min E --+ K ®min F is a metric surjection. Corollary 5.7. Let E 1 , E 2, F 1 , F2 be operator spaces.
(i) ("Injectivity") If E; C Fi completely isometrically, thell El ®h E2 C Fl ®h F2 completely isometrically. (ii) ("Projectivity") If qi: Ei --+ Fi is a complete metric surjection, then ql ® q2: El ®h E2 --+ Fl ®h F2 is also one. Proof. (i) Observe that, by the Arveson-Wittstock extension theorem, any pair O'i.: Ei --+ B(Ji) (i = 1,2) of complete contractions admits completely contractive extensions 0\: Fi --+ B(Ji). Hence, by definition of ®f, we have El ®f E2 '----> Fl ®f F2 completely isometrically; whence the result by Theorem 5.1'.
(ii) This is immediate from the very definition of El ®h E 2.
•
Self-duality. In some sense, Corollary 5.4 describes precisely the dual operator space (El ®h ... ®h En)*. However, this information can be inscribed into some very nice formulas that reflect the self-duality of the Haagerup tensor product, as follows. For simplicity, we state this only for two factors, but it trivially extends to any number of factors by the associativity of ®h. This fact is apparently due to Blecher, Effros, and Ruan (see [ER4]). Corollary 5.B. Let E 1 , E2 be two finite-dimensional operator spaces. Then (El ®h E2)*
= Ei ®h E~ completely isometrically.
(5.14)
Actually, it suffices for this to have one of the spaces E 1 , E2 finite-dimensional. Aforeover, in the infinite-dimensional case, we llave a natural completely isometric embedding (5.15)
Introduction to Operator Space Theory
94
Proof. We need a preliminary observation. Consider maps a;: E; ~ Ko (i = 1,2) and let X; E Ko 0 be the corresponding tensors. Let u = a1 . a2: E1 0 E2 ~ Ko be defined as before by U(X1 0 X2) = a1(xt)a2(X2), and let X E Ko 0 Ei 0 E2 be the tensor corresponding to u. Then a simple verification shows that
E;
Assume E 1, E2 finite-dimensional. We will identify for simplicity (EI 0 E2)* with Ei 0 E 2. Recall the notation Ko = Un AIn C K. Recall that, by definition, we have
Hence, by Remark 5.6 and our preliminary observation, for any x in Ko 0 (E1 0h E2)* (say x E Mn 0 (E1 0h E2)* for some n) with Ilxllmin ::; 1, there are x; E Ko 0 with IIx;IImin ::; 1 such that Xl 8 X2 = x. Hence, we have Oh(X) = IIxII1C®lIliU(E;®"Ei) ::; 1. Conversely, if IIxII1C®miu(E;®"E2) ::; 1, then we have Oh(X) ::; 1 (with E; instead of E; in the definition of Oh), and the (easy) implication (ii) =} (i) in Corollary 5.4 then shows that IIxllcB(El®"E2 X) ::; 1. This establishes (5.14).
E;
We now easily verify (5.15). Consider Y E Mn(Ei 0 E2). Then there are finite-dimensional subspaces G; C E; with Y E Aln(GI 0 G 2 ). We have G; = (Ed F;)* with F; C E; equal to the preannihilator of G;. Now, by Corollary 5.7, G10hG2 ~ Ei0hE2 is a complete isometry and q: E10hE2 ~ Ed FI 0h E2/ F2 is a complete metric surjection. Hence (see §2.4) q* is a complete isometry from (Ed FI 0h E2/ F2)* into (EI 0h E2)*' But, by the first part of the proof, (Ed Fl 0h E2/ F2)* = G 1 0h G 2. Hence we have
whence (5.15). Note in particular that if either El or E2 is finite-dimensional. we have equality in (5.15). •
Remark. More generally, using a classical result of Bessaga and Pelczynski (cf. [LT1, p.22]) it is not too hard to show that, if the space Ei Q3;h E2 does not contain Co as a Banach subspace, then (5.14) holds. See Exercise 5.1 for more details. In the next statement (and from now on) we use an obvious extension of our previous notation: If A is any operator algebra (for instance if A = B(H)), consider elements Yll Y2 of the form Yl = Li ai(l) 0 ei(l) E A ® Ell Y2 = Lj aj(2) ® ej(2) E A ® E2 (finite sums). Then we denote Yl 8
Y2
=
L ai(1)aj(2) ® ei(l) ® ej(2) E A 0 El 0 E i,j
2.
5. TIle Haagerup Tellsor Product
95
Corollary 5.9. Let E 1 , E2 be arbitrary operator spaces. Consider Y in B(H) ® El ® E 2 • If dim (H) = 00, tllell tIle following are equiva.lent:
(i) IlyIIB(H)0I11iU(E101oE2) ::; 1. (ii) Tll€re are Yi in B(H) ® E i. witll IIYillmin ::; 1 (i = 1,2) sucll tllat Y = Yl 0 Y2· Proof. By Remark 2.1.6, (ii) ~ (i) is valid in full generality (for any H), so we concentrate on the converse. Assume (i). By the injectivity of ®h we can assume that E 1 , E2 are both finite-dimensional. Then, let u: Ei ® E:; ~ B(H) be the mapping associated to y. By (5.14), we have Ilulleb = IlyIIB(H)0I11iU(E101o E2) ::; 1. Hence, by the last assertion in Corollary 5.4, we can write u(6 ® 6) = 0'1(~1)0'2(6) (~i E with IhIICB(E;.B(H» ::; 1. Let Yi E B(H) ® Ei be associated to O'i' We then have IIYillmill = II00ilieb ::; 1 and Y = Yl 0 Y2· •
En
At this stage, it is not hard to check the following completely isometric identities (dual to each other): (5.16)
which are valid for any integers 17" k. The infinite-dimensional analogs are also valid. l\Iore generally, for any Hilbert spaces H, J(, we have the following completely isometric identities:
He ®h
J(e
= (H ®2 J()e and
Hr
®h J(r
= (H ®2 J()r.
(5.16)'
For the proof see Exercise 5.2. The following identity is a much more significant and useful result (cr. [BPI, Proposition 5.5]. See also [ER4]). Corollary 5.10. Let E be an arbitrary operator space. TIlen, for allY illteger 17" we Ilave completely isometric isomorpllisms (5.17)
and (5.18)
via tile mappillg
L eil ®
Xij
® elj ~
L eij ®
Xij'
Afore generally, for any pair E, F of operator spaces, we lIave a complete isometry (5.19)
Proof. That (5.17) or (5.18) is completely isometric is an easy exercise left to the reader (see Exercise 5.3). Then (5.19) follows by associativity. • By duality we obtain
Introduction to Operator Space Theory
96
~ Mn(E)* and hence Mn(E)** ~ Afn(E**) completely isometrically. Moreover, we have completely isometric isomorphisms
Corollary 5.11. We have Rn Q9h E* Q9h en
Rn Q9h E Q9h en ~ M~ Q9" E R Q9h E Q9h e ~ K* Q9" E
(5.20) (5.21 )
via the mapping where eij
E
K* is the functional defined by eij (a) = aij'
Proof. By (5.14), (5.17) implies e~ Q9h E* Q9h R~ ~ Mn(E)*, and hence (since R* ~ e and e* ~ R by Exercise 2.3.5) Rn Q9h E* Q9h en ~ Mn(E)*. Iterating this, we find Mn(E)** ~ en Q9h E** Q9h Rn and hence by (5.17) again ~ Mn(E**).
We now turn to (5.21). Note that M~ Q9" E (resp. Rn Q9h E Q9h en) is completely isometrically embedded in K* Q9" E (resp. R Q9h E Q9h e), and the union over n of these spaces is dense in K* Q9" E (resp. R Q9h E Q9h e). Hence it actually suffices to prove (5.20). Then by the first part of the proof we have Rn Q9h E** Q9h en ~ Mn(E*)*, and hence by (4.8) Rn Q9h E** Q9h en ~ (M~ Q9" E)**. Since the inclusion X c X** is completely isometric (see Exercise 2.3.1 and (2.5.1)) both for X = Rn Q9h E Q9h en and for M~ Q9" E, we obtain (5.20). • In particular, when E is one-dimensional, the two preceding corollaries yield and
More generally, for any Hilbert spaces H, K, the product He Q9h Kr (resp. Hr Q9h Ke) can be canonically identified, completely isometrically, with the space of all compact (resp. trace class) operators from K* to H. Here, the space of trace class operators is equipped with its operator space structure as the predual of B(H, K*), as explained in §2.5. This illustrates the noncommutativity of Q9h. This shows (since K and K* are not isomorphic) that the spaces El Q9h E2 and E2 Q9h El may fail to be even isomorphic as Banach spaces. However, it is easy to check (using the "opposite" spaces as defined in §2.1O) that, if E!, E2 are arbitrary operator spaces, the space E2 Q9h El is completely isometric to the space (E? ®h E~P)OP, via the linear mapping that
5. The Haagerup Tensor Product takes X2 ® Xl to Xl ® X2. Equivalently, we have (E1 ®/t E 2 )OP ~ E~P completely isometrically. (This was observed in [BPI].)
97 ®h E~P,
For the complex conjugates (see §2.9), the analogous question has a simpler answer: \Ve have completely isometrically. We leave the proofs of these last two identities as (easy) exercises for the reader. The next result originates (essentially) in Haagerup's unpublished work [H3] (see [EK, BS] for more on the same theme).
Theorem 5.12. Let A C B(H) and Be B(K) be C*-algebras. We have a natural completely isometric embedding
J:
A®h
B
---->
CB(B(K,H),B(K,H))
defined by J(a ® b): T
---->
aTb.
In particular, when A = B = JI/n (n:::: 1), this map is a completely isometric isomorphism Afore generally, for all integers, n, 'In, r, s, the same map defines a completely isometric isomorphism
Proof. We only prove the last assertion and leave the rest as an exercise. We have, by Corollary 5.10 (and associativity),
JI/n,m ®h JlJr,s = C h ®h Rm ®h Cr ®h Rs
= Cn ®h (Rm ®h Cr) ®h Rs = Mn,s(M:n"r) ~
CB(Mm,n lIfn,s)'
•
Remark. Let Sn C CB(Mn , Mn) be the subspace of all the Schur multipliers of Mn (Le., maps taking each eij to a multiple of itself) and let Dn C Mn be the subalgebra of all diagonal matrices. It is easy to check that J(Dn ® Dn) = Sn and, since Dn ~ £:::0 (completely isometrically), we obtain completely isometrically
£:::0
®h £~ ~
Sn C CB(Mn, Mn).
Actually, Haagerup proved that, for any Schur multiplier T: lIJn ----> lIJn , we must have IITlicb = IITII (cf., e.g., [PlO, p. 100]); hence we also find an
Introduction to Operator Space Theory
98
isometric embedding of £~ Q9h £~ into B(Mn , Mn). See [ChS] and references there for more on this theme. Free products. Let Al,A2 be two C*-algebras (resp. unital C*-algebras). We will denote by Al *A 2 (resp. Al *A 2) their free product (resp. free product as unital C*-algebras). This is defined as follows (cf., e.g., [VDN]): We consider the involutive algebra (resp. unital algebra) A (resp. A) that is the free product in the algebraic sense. By the universal property of free products, for any pair 1I'i: Ai ~ B (i = 1,2) of homomorphisms (resp. unital ones) into a single algebra (resp. unital algebra) B, there is a unique homomorphism (resp. unital) 11'1 *11'2: A ~ B (resp. 11'1 * 11'2: A ~ B) that extends 11'1 and 11'2 when Al and A2 are embedded into Al *A 2 (resp. Al * A 2) via their natural embedding. Now assume B = B(H) and 11'1,11'2 representations. Then 11'1 * 11'2 and 11'1*11'2 are *-homomorphisms, so that we can introduce a C*-norm on A (resp. A) by defining for all x in A (resp. A)
where the supremum runs over all pairs (resp. unital pairs) (11'1,11'2) of representations of Al,A2 on an arbitrary Hilbert space H. The completions of A (resp. A) for these norms are C* -algebras denoted, respectively, by Al *A 2 and Al * A 2. Clearly, a similar definition leads to the free products Ai and iEI i~I Ai for an arbitmry family (Ai)iEI of C* -algebras (resp. unital ones).
*
Note that we have a canonical bilinear map (aI, a2) ~ al ·a2 from Al x A2 into Al *A 2 (resp. Al * A 2), and hence a canonical linear map from Al Q9 A2 into Al *A 2 (resp. Al *A 2). More generally, for any N-tuple AI, ... , AN of C*algebras, we have a canonical linear map Al Q9 ... Q9 AN ~ Al *A N (resp. Al * ... * AN) that takes al Q9'" Q9 aN to ala2." aN (where the product on the right is the product in the free product viewed as containing canonically each algebra AI, . .. , AN).
*...
The next result shows that the Haagerup tensor product is intimately connected to free products. Theorem 5.13. Let El"'" EN be a family of operator spaces given with completely isometric embeddings Ei C Ai (i = 1, ... , N) into C* -algebras (resp. unital ones) AI, . .. , AN. Then the canonical map, restricted to El Q9 ... Q9 EN, defines a completely isometric embedding of El Q9h ... Q9h EN into Al *AN (resp. into Al * ... * AN).
*...
This is proved in [CES] in the nonunital case. The unital case is worked out in detail in [Pl1]. The proof requires a dilation trick in order to replace Ntuples of complete contractions (as in the definition of Q9J above) by N-tuples of C* -representations.
5. The Haagerup Tensor Product
99
Lemma 5.14. ConsiderC*a1gebras AI,"" AN and operator spaces Ei C Ai
(completely isometrically). The equivalent assertions considered in Corollary 5.4 are equivalent to: (iii) There are a Hilbert space 1{, representations 7ri: Ai - 4 B(1{), lind operators~: 1{ -4 H and 'I: H -4 1{ with II~II :::; 1 and 11'111 :::; 1 such that
Moreover, if the C* -algebras are all unital, the representations 7ri can also be assumed all unital. Proof. (iii) =:;. (ii) is obvious. Conversely, assume (ii). By Remark 5.5, we can assume (5.13). Replacing the spaces iii by suitable enlargements, we can = K for all i = 2, ... , N. Since IITili :::; 1 in (5.13), assume (5.13) with it is easy to show that there are unitary elements Ui in 1I12(B(K)) such that
k
Ti = (Ui )l1. (Indeed, one just considers for each contraction T the matrix
(1 - TT*)1/2) ) T* . Let p: K $ K -4 K (resp. j: K -4 K $ K) be the canonical projection (resp. injection) for the first coordinate. Let
Then it is very easy to verify that
Without loss of generality, we may assume that there are unitaries V; such that U l = VI V2*' U2 = V2V3* , ... , UN -1 = VN -1 VN (indeed, we just choose VI arbitrarily, say, VI = 1; then the relations determine V2, V3,'" successively). Then, if we replace riO by 7ri = V;*ri(')V;, we obtain (5.22) with 1{ = K$K, ~ = TOPVl' and 'I = VNjTN . To check the unital case, we need to go one step further. So we assume that the Ai are unital and that (5.22) holds but with a priori nonunital representations 7ri. Let Pi = 7ri(I). Note that Pi is a projection on 1{ commuting with the range of 7ri. By associativity (see below), we may and do assume that N = 2. Note that (by considering, e.g., At Q9min A 2 ) we know that there exists a pair Pi: Ai -4 B(L) (i = 1,2) of unital (and faithful) representations on the same Hilbert space L.
100
Introduction to Operator Space Theory
Note also that, by suitably augmenting H, we can assume that (l-pt}(H) and (1- P2)(H) are of the same Hilbertian dimension, and that they are both isometric to some direct sum of copies of L. This allows us to define (using PI and P2) *-homomorphisms Pi: Ai ---t B(H) (i = 1,2) such that
Then we define for ai E Ai
+ (1 = P2 7r2(a2)P2 + (1 -
a1(ad = Pl 7l't(a1)P1
pdPI(at}(l - pd
a2(a2)
P2)P2(a2)(1 - P2).
but now a1,a2 are unital representations (Le. *-homomorphisms); hence this yields finally, for Xi E E i ,
which completes the unital case of (5.22) when N = 2. It follows from (5.22) that E I @h E2 is naturally embedded into the unital free product A1 * A2 (see below). Then, if N = 3, we may apply what we just proved to (E1 @hE2)@h£3 and this yields (5.22) for N = 3 and so on for larger N. •
Remark. Lemma 5.14 is proved in [eES] (see also [YJ) for the nonunital free product, but nothing is said there about the unital case, which is verified in
[Pll]. Proof of Theorem 5.13. By the factorization Theorem 1.6, the equivalence of (i) and (iii) in Lemma 5.14 means, in the unital case, that a mapping u: E 1 @h ... @h EN ---t B(H) is a complete contraction iff it extends to a complete contraction on the free product Al * ... * AN. Equivalently, this means that we have a completely isometric embedding
In the nonunital case, the proof is identical.
•
The next result (inspired by [ER10, EKRJ) on the "tensor shuffling" is one of the many applications of the free product connection.
5. The Haagerup Tensor Product
101
Theorem 5.15. Let Ei (i = 1,2,3,4) be operator spaces. lVe have a natural completely contractive map
that reduces to a permutation on the algebraic tensor prodllct. Proof.
Assume E j C Ai for some C*-algebra Ai' Let B = A 1 *A 3 and Consider the C*-algebra A = B ®lTIin C. Clearly the product map p: A x A ----+ A defines a complete contraction from A ®h A to A. Now, if we view Al and A3 (resp. A2 and A4) as subalgebras of B (resp. C), we have
C
= A 2 *At .
which yields that p restricted to (El ®lTIin E 2 ) ®h (E3 contractive. But the latter restriction takes value in
®lTIin
E 4 ) is completely
where the products are meant in Al *A 3 and A 2*A 4 , respectively. By Theorem 5.13, span(EI . E 3 ) ~ El ®h E3 and span(E2 . E 4 ) ~ E2 ®h E 4 • so we obtain the announced result. • Factorization through R or C. Let X be a fixed operator space. A linear map u: E ----+ F (between operator spaces) is said to factor through X if there are c.b. maps v: X ----+ F and w: E ----+ X such that u = tlw. We will denot.e by r x (E, F) the set of all such maps, and we define
where the infimum runs over all possible such factorizations. Note that. in general 1x (.) is not a norm and r x (E, F) may fail to be stable under addition, but when X is "nice" (as below), this "pathology" disappears. In the following section, we will study the cases X = R and X = C. Proposition 5.16. ([ER4j) Let E, F be operator spaces. Tllen the natllral mappings E* ® F ----+ C B (E, F) and E ® F* ----+ C B (F, E) extend to isometric em beddings E*
®h
Fer R(E, F)
and
E
®h
F* C
r c(F, E).
Aforeover, if either E or F is finite-dimensional, these embeddings are actually isometric isomorphisms. Similarly, we have isometric embeddings E ®h F C rR(E*, F) and E ®h Fe rc(F*, E). Proof. Consider x E E* ® F and let u: E ----+ F be the associated finite rank linear map. Let E~R~ F be a factorization of u through R. Since
102
Introduction to Operator Space Theory
u has finite rank, say, equal to n, and R is "homogeneous" (which means that (1.7) holds), we can replace R by a suitable subspace (and project onto it) to find a modified factorization E~Rn~F with IIWl11cb = Ilwllcb and IIVl11cb = Ilvllcb. Identifying WI with a = E eli ® ai E Rn ® E* and VI with b = E eil ® bi E C n ® F, we find x = a 8 b with
Conversely, to any a, b as above corresponds a factorization' of u so that we ' have
'YR(u) = inf{llallminllbllmin I a E Rn ® E*, bE Cn ® F,x = a 8 b}, which means precisely that
The other assertions are proved similarly.
•
Remark. If we equip rR(E,F) and rc(F,E) with appropriate o.s.s., then the preceding isometries become completely isometric. The appropriate structure on r R (E, F) (resp. r c( F, E)) is the one that gives to the space Mn(rR(E, F)) (resp. Mn(rc(F, E))) the norm of the space rR(Rn(E), Rn(F)) (resp. rc(cn(F),Cn(E))), where we have denoted
Symmetrized Haagerup tensor product. It is natural to investigate what happens to ®f if one restricts the maps (ab (72) to have commuting ranges. This produces a symmetrized version of the Haagerup tensor product that is studied extensively in toP]. We will describe only one basic result from toP] and its consequences. Let (Eb E 2) be operator spaces. We denote by C the collection of all pairs a = (171, (72) of complete contractions a i: Ei ~ B (H(f ) (into the same Hilbert space) with commuting ranges. Let cI>: El ® E2 ~ E9 B(H(f) be the embedding defined by cI> = EB(fECal . 172 or, more explicitly: (fEC
cI>(XI ®X2) = Efjal(xl)a2(x2). (fEC
This linear embedding induces an operator space structure on El ® E 2; we denote by El ®,.. E2 the resulting operator space after completion. A similar construction makes sense for N-tuples of operators (Eb . .. , EN) and produces an operator space denoted by (El ® .•. ® EN),.. (so that (El ®
5. The Haagemp Tensor Product
103
E 2 )p, is the same as E1 ®p, E2). This new tensor product is projective, and, by its very construction, it is commutative, but it fails to be either associative (see lOP] for Le Merdy's argument for this) or injective.
Theorem 5.17. ([OP)) Let E 1, E2 be two operator spaces. Consider tIle mapping
defined on the direct sum of tlle linear tensor products by Q( tt ffi v) = u + tv. Tllen Q extends to a complete metric surjection from (E1 ®h E 2)EBI (E2®hEJ) onto EI ®p,E2. In particular, for anytt in E1 ®E2 , we llmre: Ilttllp, < 1 ifftllere are v, 'W in E1 ® E2 sucll tllat u = v + 'Wand IIvllE)0"E2 + IIlwllE20"E) < 1. The preceding statement means that E1 ®'L E2 is completely isometric to the "sum" E1 ®h E2 + E2 ®h E1 (in the style of interpolation theory; see §2.7) in analogy with the space R + C described latf'r in §9.8. The case of N-tuples is also considered in lOP], but only the complet.f'ly isomorphic analog of the preceding result is proved. l\Ioreover, lOP] also contains a different result (complet.ely isometric, this time) for N-tuples of completely contractive maps O"i: Ei -> B(H(j) with "cyclically commuting" ranges (see lOP, Theorem 19]).
By Corollary 5.4, one can see t.hat the following statement is a dual reformulation of Theorem 5.17. Theorem 5.1S. ([OP)) Let E1,E2 be two operator spaces, and let
B(1t) be a linear mapping. Tlle following are equivalent:
(i) II B(1t, H), and (31: E1 -> B(1t, H), (32: E2 B(H,1t) sudl tllat
Q9
->
->
(iii) For some Hilbert space H, tllere are complete contractions O"i: Ei -> B(H) (i = 1,2), with commuting ranges, and contractions V: 1t -> H and lV: H -> 1t sudl tllat
Proof. Assume (i). Then
104
Introduction to Operator Space Theory
0'1: E1
1---+ B(H1 EB H2 EB H 3) and 0'2: E2 notation, as follows:
-+
B(H1 EB H2 EB H 3) using matrix
0 fil~,) ~0 a+»0!1 (Xl)
0 0
0'1(X1) =
)
f32(X2)
u,(x,)
0 0
Then, by (ii), we have
hence 0'1,0'2 have commuting ranges and are complete contractions. Therefore, if we let W: HI EBH2EBH3 --+ 1t be the projection onto the first coordinate and V: 1t --+ HI EB H2 EB H3 be the isometric inclusion into the third coordinate, then we obtain (iii). Finally, the implication (iii) implies (i) is obvious by the very definition of E 1 ® f.I. E 2 • •
Proof of Theorem 5.17. By duality, it clearly suffices to show that, for any linear map cp: E1 ® E2 --+ B(1t), the norms lIifllcB(El®"E2 -+B(1tll and IIcpQllcb are equal. But this is precisely the meaning of the equivalence between (i) and (ii) in Theorem 5.18. Thus we conclude that Theorem 5.18 implies • Theorem 5.17.
u:
Remark. Let X = R EB C. Let E1 --+ E2 be a finite rank map with associated tensor u E Ei ® E 2 • Using Proposition 5.15, it is easy to check that Theorem 5.17 implies (5.23)
Moreover, it is easy to see that the identity of X factors completely contractively through K, and hence /1(.(u) ~ /x(u).
(5.24)
Corollary 5.19. ([OP]) Let E be an n-dimensional operator space. Let iE E E* ® E be associated to the identity of E and let
5. Tlle Haagerup Tensor Product
105
Then (5.25)
Moreover, /1(E) = 1 iff eitller E = Rn or E = C n (completely isometrically). Proof. Note that (5.25) clearly follows from (5.23). Assume that IL(E) = 1. Then, by Theorem 5.17 and Proposition 5.16 (using also an obvious compactness argument), we have a decomposition IE = Ul + U2 with (5.26) In particular, this implies that 12(IE) = 1 (where 12(') denotes the norm of factorization through Hilbert spacej see, e.g., [P4, chapter 2] for more background), whence that E is isometric to C~ (n = dimE). Moreover, for any e in the unit sphere of E we have
Therefore we must have (5.27) Let ai = Iluil!, so that (by (5.26» al + a2 = 1. Assume that both al > 0 and a2 > O. We will show that this is impossible if n > 1. Indeed, then Ui. = (ai)-lui (i = 1,2) is an isometry on e~, such that, for any e in the unit sphere of E, we have e = al U l (e)+a 2 U2 (e). By the strict convexity of f~, this implies that Ul(e) = U2 (e) = e for all e. Moreover, by (5.27) we have IR(Ut} = 1 and Ic(U2 ) = 1. This implies that E = Rn and E = C n completely isometrically, which is absurd when n > 1. Hence, if n > 1, we conclude that either al = 0 or a2 = 0, which implies either IC(IE) = 1 or IR(IE) = 1, or, equivalently, either E = C n or E = Rn completely isometrically. The remaining case n = 1 is trivial. • Remark. Here is an alternate argument: If /1(E) = 1, then, using (5.23), (5.24) and the reflexivity of E, we see that, for both E and E*, the identity factors through K**j therefore E is an injective operator space as well as its dual. Now, in [Ru2], Ruan gives the complete list of the injective operator subspaces of finite-dimensional C*-algebras (see also [EOR] for more on this theme). Running down this list, and using an unpublished result of R. Smith which says that a finite-dimensional injective operator space is completely contractively complemented in a finite-dimensional C* -algebra (see [B2]), we find that Rn and C n are the only possibilities. For the isomorphic version of the last statement, we will use the following result from [01]:
106
Introduction to Operator Space Theory
Theorem 5.20. ([01]) Let E be an operator space SUell that IE can be factorized completely boundedly through the direct sum X = He EBl Kr of a column space and a row space (i.e., there are c.b. maps u: E ---+ X and v: X ---+ E such that IE = vu). Then there are subspaces El C He and E2 C Kr SUell that E is completely isomorpllic to El EBl E 2 • More precisely, if we have Ilullebllvileb ~ c for some number c, then we can nnd a complete isomorphism T: E ---+ El EBl E2 such tllat IITllebllT-1lleb ~ f(c), where f: 1R+ ---+ 1R+ is a certain function. Theorem 5.21. ([OP]) The following properties of an an operator space E are equivalent: (i) For any operator space F, we have F ®min E = F ®,.. E isomorpllically. (ii) E is completely isomorphic to the direct sum of a row space and a column space.
Proof. The implication (ii) =} (i) is easy and left to the reader. Conversely, assume (i). Then, a routine argument shows that there is a constant K such that for all F and all u in F ® E we have lIull,.. ~ Kllullmin. Let SeE be an arbitrary finite-dimensional subspace, let is: S ---+ E be the inclusion map, and let ts E S* ® E be the associated tensor. Then we have by (5.23) sUPs 1'RE!)lc(is) = sUPs IItsll,.. ~ K. By a routine ultraproduct argument, this implies that the identity of E can be written as in Theorem 5.20 with c = K; thus we conclude that E ~ El EBl E 2, where El is a row space and E2 is a column space. Note that we obtain an isomorphism T: E ---+ El EBl E2 such that IITllebllT-11leb ~ f(K). In particular, if E is finite-dimensional, we find T such that
•
Complex interpolation. The Haagerup tensor product also has a nice "commutation property" with respect to complex interpolation (defined in §2.7) that can be briefly described as follows. Let (Eo, E 1 ) and (Fo, F 1 ) be two compatible pairs of operator spaces. Then the couple (EO®hFO, El ®hFt) can be viewed as compatible. Indeed, assume that (Eo, E 1 ) (resp. (Fo, Ft}) are continuously injected into a Banach space £ (resp. F), which allows us to view the couple as compatible. Then, it is easy to check that the tensor product of the suitable injections defines a continuous injection from Eo ®h Fo into the v
injective tensor product £ ® F. Similarly, El ®h Fl is continuously injected into the same space £ F. (Note: The injectivity of the tensor product of two injective mappings can sometimes fail, but not for tensor norms as nice as the injective one or the Haagerup one. See [PI, Theorem 6.6] for a general result related to this technical point.) Thus we may view the couple (Eo ®h Fo,El ®h Ft) as compatible. Then we have (cf. [PI, Theorem 2.3]):
0
5. The Haagerup Tensor Product
107
Theorem 5.22. Let (Eo, E l ) and (Fo, F l ) be two compatible couples of operator spaces. Then, viewing the couple (Eo Q9h F o, El Q9h F l ) as compatible as explained above, we llave a complete isometry
Exercises Exercise 5.1. Let E l , E2 be operator spaces such that Ei Q9h E2 does not contain a subspace isomorphic (as a Banach space) to co. Then
Exercise 5.2. Prove (5.16) and (5.16)'. Exercise 5.3. Prove (5.17) and (5.18). Deduce from (5.18) that C Q9h E = C Q9min E and E Q9h R = E Q9min R for any operator space E. Use this to give an alternate solution to Exercise 5.2. Exercise 5.4. Let Ei C B(Hi } (i = 1,2) be operator spaces. Then any x in El Q9h E2 can be written as a series of the form L~ an Q9 bn , where an E El bn E E2 with the series L~ ana~ and L~ b~bn converging in norm in B(Hd and B(H2 }, respectively. Moreover, in that case, the series L~ a n Q9bn converges in El Q9h E 2 , and, if Ilxlih < 1, we can choose (an) and (b n ) so that
Exercise 5.5. Show that for any operator space E we have completely isometrically
Exercise 5.6. Show that if E is finite-dimensional, we have completely isometrically (K Q9min E)** C:::' K** Q9min E. Moreover, if A is any C*-algebra, we have (K
Q9min
A)**
C:::'
K**Q9A**.
Exercise 5.7. Let E l , E2 be maximal operator spaces. Then for any x in El Q9 E2 we have
Introduction to Operator Space Theory
108
where the infimum runs over all n and all possible ways to decompose x as n
X=
L
aijXi ®Yj'
i,j=l
Exercise 5.8. Let A, B be arbitrary C* -algebras and let A*B be their (nonunital) free product. For any k ;::: 1, let T2k = A ®h B ®h ... ®h A ®h B (where A and B each appear k times) and T2k+1 = T2k ®h A. For any d > 1, let 1/Jd: 7d ---+ A*B be the completely contractive map induced by the product map in A*B. Show that 1/Jd is a completely isomorphic embedding of 7d in A*B. l\.fore precisely, the range 1/Jd(7d) is the closed subspace Wd C A*B spanned by all the "words" of length d that begin in A, and we have
lI1/Jdi!vJcb:::; (d _1)d-l. Exercise 5.9. Fix an integer n. The aim of this exercise is to produce explicit completely isometric embeddings of C n and Rn into .e~ ®h.e~. We will denote by (ei) the canonical basis of .e~. Let W = (Wik) be an n x n matrix with unimodular entries (Le., IWikl = 1 for all i, k) such that
Iln- 1/ 2 wIIMn = 1. In
.e~ ®h.e~
we consider the vectors Xi
= ei ® L
Wikek
k
Yi
=L
Wikek ® ei'
k
(Note that Yi = t Xi .) Let Ex c .e~ ®h.e~ (resp. Ey C .e~ ®h .e~) be the operator subspace spanned by {Xi I i = 1,2, ... , n} (resp. {Yi Ii = 1, ... , n}). Prove that Ex :::; C n (resp. Ey := Rn) completely isometrically.
Exercise 5.10. The aim of this exercise is to produce an explicit completely isometric embedding of Mn into .e~ ®h .e~ ®h .e~. Let w, Wi be two n x n matrices with the same properties as in the preceding exercise. We introduce . the following elements of the operator space .e~ ®h.e~ ®h i'~: Zij
= ei ®
(~WikW~jek) ® ej.
Then the span of {Zij I i, j = 1, ... , n} in .e~ ®h .e~ ®h .e~ is completely isometric to Mn. Let H be an infinite-dimensional Hilbert space. Show that a matrix (aij) in Mn(B(H» is in the unit ball iff it can be factorized as a matrix product Dl(n-l/2w)D2(n-l/2w')D3, where Db D 2, D3 are three diagonal matrices in the unit ball of B(H).
Chapter 6. Characterizations of Operator Algebras In the Banach algebra literature, an operator algebra is just a closed subalgebra of B(H). A uniform algebra is a closed unital subalgebra of the space C(T) of all continuous functions on a compact set T, which is (usually) assumed to separate the points of T. In the 1970s, the theory of uniform algebras gave birth to the notion of Q-algebras (quotients of a uniform algebra by a closed ideal) and ultimately uncovered some surprising stability properties of the class of operator algebras. Specifically, in the late 1960s, B. Cole (see [We]) discovered that Q-algebras are necessarily operator algebras, and soon after that, G. Lumer and A. Bernard proved that the class of operator algebras is stable under quotients (see Theorem 6.3). One of the natural problems considered dming that period was to characterize the Banach algebras that are isomorphic to a uniform algebra, or a Q-algebra, or an operator algebra. Among the many contributions from that time, those that stand out are Craw's Lemma characterizing uniform algebras and Varopoulos's work on operator algebras ([Vl-2]). Varopoulos discovered that, if the product mapping PA: x Q9 y ---4 xy of a Banach algebra A is continuous on the tensor product A Q9 A equipped with the ')'2-norm, then A is an operator algebra. (From this he deduced, using Grothendieck's inequality, that any Banach algebra structure on a commutative C* -algebra is necessarily an operator algebra.) Several authors then tried to characterize operator algebras by a property of this type but with a different tensor norm than the ')'2-norm. Variations on this theme were given by P. ChaTpentier and A. Tonge (see [DJT, Chapter 18]) until K. Carne [Cal somewhat closed that chapter by showing that operator algebras cannot be characterized by the continuity of the product map on a suitable tensor product. More precisely, he showed that there is no reasonable tensor norm ')' such that the continuity of PA: A Q9')' A ---4 A characterizes operator algebras. In sharp contrast with Carne's result, it turns out that in the category of operator spaces the situation is much nicer. We have Theorem 6.1. ([BRS]) Let A be a unital Banad1 algebra witll unit 1A given witll an operator space structure sucll tllat IlIA II = 1. TIle following are equivalent.
(i) TIle product map is completely contractive. (ii) Tllere is, for some Hilbert space H, a completely isometric unital 110momorpllism j: A ---4 B(H). Remark. Let us denote here simply x . y for PA(X, y). Note that, by the definition of ®h, (i) is clearly the same as:
110
Introduction to Operator Space Theory
(i)' The natural matrix product
ab =
(L.:
aik . bkj) ij
k
defines a unital Banach algebra structure on Mn(A) for all n In other words, (i) holds iff Ilell = 1 and
~
l.
The implication (i)' (¢:> (i)) ~ (ii) shows that this implies that, for any unital operator algebra B, B ®min A is equipped with a natural unital operator algebra structure corresponding to the tensor product of PB and PA. Curiously, the isomorphic version of this theorem resisted until Blecher [B4] recently proved the following statement. Theorem 6.2. ([B4J) Let A be a Banach algebra given with an operator space structure. The following are equf.,lalent. (i) The product map PA: A ®h A ---- A is completely bounded. (ii) There is, for some Hilbert space H, a homomorphism j: A ---- B(H), which is a complete isomorpllism from A to j(A). Remark. Note that (i) in Theorem 6.2 holds iff K[A] = K ®min A is a Banach algebra up to a constant for its natural matrix product, in other words, iff there is a constant C such that Ilabll ::; Cllalilibil for all a,b in K[A]. The original proofs ofthese theorems do not use the results from the 1970s, and can be used to give new proofs of the latter. However, we have recently found a new approach: We will deduce the preceding two theorems from the stability under quotients of operator algebras. Note, however, that, both for Theorems 6.1 and 6.2, some trick evocative of the original proofs of [BRS, B4] still remains. Our approach is based on the following statement (due to A. Bernard, to G. Lumer independently in some form, and originally due to Cole when A is a uniform algebra). Theorem 6.3. Let A c B(H) be a closed subalgebra of B(H) and let I c A be a closed ideal. Then there is, for some Hilbert space 1-1., an isometric (and completely contractive) homomorphism cp:
AI I
----+
B(1-I.).
In other words, the class of operator algebras is stable under quotients. Proof. We essentially follow Dixon's exposition in [Dix1]. Consider x E AII with IIxll = 1. We claim that there is, for some ii, a homomorphism
111
6. Characterizations of Opera. tor Algebras
'Px: All
--+ B(ii) with II'Pxllcb ::; 1 and II'Px(x)11 = Ilxll = 1. From this claim the conclusion of Theorem 6.3 follows immediately: We simply let 'P = EBx 'Px (with x running over the unit sphere of A). We now briefly justify this claim. Consider ~ in the unit sphere of (AI1)* such that ~(:l:) = 1. Let us denote by q: A --+ AJI the quotient map and let [ E A* be defined by [(a) = ~q(a», so that II~IIA* = 1. Then there is a representation 7r: B(H) --+ B(H) and elements s, t in the unit ball of ii such that
[(a) = (7r(a)s, t).
(6.1)
\;fa E
A
(This is entirely elementary and well known, but we can view it as a very special case of Theorem 1.6 since II[IIA' = 1I€'IlcB(A.q.) Now l('t E1 C ii and E2 C E1 be defined by
E1 = span[s, 7r(a)s I a E A] E2 = span[7r(i)s liE I]. Then E1 and E2 are clearly invariant under 7r(A); therefore the subspace
is semi-invariant with respect to 7r(A), so that (by Sarason's well-known ideas, see, e.g., [PlO, Theorem 1.7]) the compression ii': A --+ B(E) defined by
is a homomorphism. Indeed, for a, bE A, we have PE7r(a)PE2 = 0, and hence ii'(a)ii'(b) = ii'(ab). lVloreover, ii' is clearly contractive (actually completely contractive). Observe that since [ vanishes on I, we have t E E;}-. Thus, it is now easy to check that (6.1) yields
[(a) = (ii'(a)PEs, t),
(6.2)
and also, since 7r(I)E1 C E 2 , that ii': A --+ B(E) vanishes on I. Hence, if we let 'Px: AI I --+ B(E) be defined by ii' = 'Pxq, we have (take a with q(a) = x in (6.2» 1 = IIxll = ~(x) = ('Px(x)PEs, t), whence 1 ::; II'Px(x)ll. Since (by (2.4.2» II'Pxllcb claim.
=
11ii'llcb ::; 1, this proves our ..
Corollary 6.4. In the situation of Tlleorem 6.3, there is, for some Hilbert space 11., a completely isometric homomorphism
1/;: AI I
--+
B(11.).
112
Introduction to Operator Space Theory
Proof. Two simple proofs of this can be given. A first proof consists of applying Theorem 6.3 for each integer n to the algebra Jl.Jn(A)/Mn(I) ~ Mn(A/ I). Assuming A and r.p: Mn(A)/ Mn(I) --+ B(1i) unital, it is easy to check (by elementary algebra) that r.p is necessarily of the form Iu" ® '¢n for some homomorphism '¢n: A/I --+ B(1i). A direct sum argument (considering '¢ = EEln'¢n) then completes the proof. A second (and better) proof consists of repeating the proof of Theorem 6.3, but this time with in the unit sphere of CB(A/I,B(L» for some Hilbert space L. The same proof (using the fundamental Theorem 1.6 to factorize [: A --+ B(L) instead of (6.1» then yields Corollary 6.4. •
e
Remark. The analogs of Theorems 6.1 and 6.2 for dual operator algebras are proved in [LeM6] (see also [B5]). Remark. See [LeM2] for an extension of the preceding two results to the case when the class of Hilbert spaces is replaced by that of subspaces of quotients of Lp, with p =I 2. The idea of the proofs of Theorems 6.1 and 6.2 is very natural: We will represent A as a quotient of the "free operator algebra" generated by A. The latter is defined as follows. We will distinguish between the unital and nonunital cases. Let E be an operator space. Let T(E) (resp. Tu(E» be the tensor algebra (resp. the unital tensor algebra) over E, that is, T(E)
= E EEl (E ® E) EEl .. ·
and
We view T(E) as a subspace of Tu(E). Then T(E) (resp. Tu(E» is an algebra (resp. a unital algebra) containing E as a linear subspace and characterized by the property that any linear map v: E --+ B from E into an algebra (resp. a unital algebra) B uniquely extends to a homomorphism (resp. unital homomorphism) T(v): T(E) --+ B (resp. Tu(v): Tu(E) --+ B). We will denote the free operator algebra associated to E by OA(E), and we define it as follows: Let C = {v: E --+ B(Hv) I IIvllcb ~ I}, where Hv is an arbitrary Hilbert space with dimension at most, say, the density character of E. We define a linear embedding J: Tu(E)
--+
ffiB(Hv) C B (ffiHV) vEe
vEe
by setting J(x) = ffiTu(v)(x). vEe
6. Characterizations of Operator Algebras
113
Note that J is a unital homomorphism. We define OAu(E) (resp. OA(E)) as the closure of J(Tu(E)) (resp. J(T(E))) in B (ffiVEC Hv). Thus OAu(E) is a unital operator algebra and OA(E) C OAu(E) is a closed subalgebra. It is easy to check that OAu(E) is characterized by the following universal property: For any map v: E ---+ B(H) with Ilvllcb :=:; 1 there is a unique unital homomorphism OAu(E) ---+ B(H), extending v, with Ilvllcb :=:; 1.
v:
The elements of Tu (E) can be described as the vector space of formal sums:
where each Pj (1 :=:; j :=:; N) is "homogeneous" of degree j, that is, Pj is of the form
with Ao(j) E C and e'j'(j) E E (1 :=:; i :=:; j). Similarly, any element Q of T(E) can be written as Q = PI + ... + PN . For any v E C, let
v(Pj and let v(P)
)
= L Aa(j)v(er(j))··· v(ej(j))
= AoI + v(P1 ) + ... + v(PN )
and v(Q)
= v(P1 ) + ... + v(PN ).
Clearly v(P) and v(Q) are nothing but Tu(v)(P) and T(v)(Q). Then we have 1IPIIoA,,(E) = sUPvEC IIv(P)11 and IIQlloA(E) = sUPvEC Ilv(Q)II· Explicitly, this means that N
AoI + L
IlPlloA,,(E) = sup vEe:
Ao(j)v(er(j))v(e~(j)) ... v(ej(j))
L
j=l
a
B(H,.)
l\'1ore generally, let us now assume that the coefficients Ao, Ao(j) are all in K (or even in B(f2)) and let P be an element of K 0 Tu(E) of the form N
P
= Ao 01 + LLAa(j) 0er(j) 0··· 0ej(j). j=l
a
Then the above definition of OAu(E) means that
N
= sup
vEC
Ao 0 I
+L j=l
Aa(j) 0 v(er(j)) ... v(ej(j))
(6.3) K®miuB(H,,)
Of course a similar formula holds for OA(E) with the constant term omitted.
114
Introduction to Operator Space Theory
Remark 6.5. Let E be any operator space. Consider the iterated Haagerup tensor product X = E ®h E . .. ®h E (N times). Let x be arbitrary in IC ® E ® ... ® E. Then x can be written as a finite sum
x= L
Ai ®
xI ® ... ® xi"
i
with Ai E IC and
xl, ... ,xi" E E.
By Corollary 5.3(iii), we have
where the supremum runs over all possible choices of H and of complete contractions 0'1: E --+ B(H), ... ,aN: E --+ B(H). We claim that this supremum is actually attained when 0'1, ..• , aN are all the same; more precisely, we have
where the supremum runs over all possible H and all complete contractions --+ B(H). Indeed, this follows from a trick already used by Blecher in [B1] and which seems to have originated in Varopoulos's paper [V2]. The trick consists of replacing 0'1, ..• ,aN by the single map v: E --+ B(H EB ... EB H) ~ N+l times of the form v: E
o v(e) =
o
o
(More precisely, v(e) is the (N + 1) x (N + 1) matrix having (al(e), ... , aN (e)) above the main diagonal and zero elsewhere.) Then it is easy to check that
Ilvllcb = SUPj lIa j IIcb, 'v'xl, ... ,x N
E
E,
From this our claim immediately follows. Proposition 6.6. Let E be any operator space. Fix N ~ 1. Let EN = E ® '" ® E (N times). Consider EN as embedded into OA(E). Then the identity mapping on EN extends to a completely isometric embedding of
6. Characterizations of Operator Algebras
115
E ®h ... ®h E (N times) into OA(E). l\!oreover, there is a completely tractive projection from OA,,(E) onto the range of this embedding.
COll-
Proof. For simplicity let us again denote X = E®h·· ·®hE (N times). Since the algebraic tensor product E®·· ·®E is dense in X, to show the completely isometric part it suffices to prove that, for any element G in K ® E ® ... ® E, we have
IIGI1K0IHinX = IIGI1K0IHiuOA(E). But this is immediate by (6.3) and (6.4). We now show that we have a "nice" projection QN from OA(E) onto the closure of EN in OA(E). Let x be a typical element of T(E), of the form x = L::=o Xn with Xn E En viewed as included in OA,,(E) (so J is viewed as an inclusion) and (say) N ~ 1n. Note that by (6.3) we have
IlxIIOA(E) = for any real t and XN =
IlL eintXn I
OA(E)
(6.5)
f e-iNt(L: k eiktxk)dt/27f, which implies (6.6)
This proves that the projection defined by QN(X) = XN is contractive. But by an obvious modification (using operator coefficients instead of scalars) one easily verifies that it actually is completely contractive. • Let E be an operator space equipped with a Banach algebra structme. Let p: E ® E --+ E be the linear map associated to the product map. Then, by associativity, p uniquely extends to a mapping
p:
T(E)
--+
E,
which is the identity on E and satisfies for any N ;::: 2 and any Xi in E
To prove Theorem 6.2, we will need the following simple lemma (already used in [B4]). Lemma 6.7. For any z in C, let us denote by a z: T( E)
--+ T( E) the homomorphism that, wIlen restricted to E, is equal to z IE. Then, if the product map extends to a mapping on E ®h E with IlpllcB(E01,E.E) = c, we 118ve, for any z with Izi < lie,
~ II Izl I paz cb:S 1 -Izlc
(6.7)
116
Introduction to Operator Space Theory
Proof. By an obvious iteration p defines a product mapping from E ®h ... ®h E (n times) into E with c.b. norm::; cn-l. By Proposition 6.6, the natural embedding of E ®h ... ®h E (n times) into OA(E) is a complete isometryj hence the restriction of p to E ® ... ® E c T(E) c OA(E) has c.b. norm ::; cn - l . We have
and hence by (6.5)
or, equivalently,
IipO'z11 ::;
",00
L.."l
Izlnen-l =
1-
Izl . Izle
By a simple modification (left to the reader) we obtain the same bound for the c.b. norm of PO'z, whence (6.7). •
Proof of Theorem 6.2. The implication (ii) ::::} (i) is immediate by Corollary 5.4. We now prove the converse. Assume (i). We denote by E the operator space underlying A. Fix z with Izl < l/e. Clearly pO'z is a surjective homomorphism from OA(E) onto E. Let I be its kernel and let w: OA(E)/I --+ E be the associated isomorphic homomorphism. By (6.7) and the definition of the quotient operator space structure (cf. §2.4), we have Ilwllcb ::; l~i;,c' Let i: E --+ OA(E) be the canonical inclusion and let q: OA(E) --+ OA(E)/ I be the quotient map. Clearly we have O'zi(x) = zi(x)j hence pO'zi(x) = zx or (wq)i(x) = zx, so that (l/z)qi = w- l . Whence
Thus if we choose for instance the value z = 1/2e, we find a homomorphism liw-llicb ::; 2c and liwllcb ::; l/e. By Corollary 6.4, this completes the proof. • w: OA(E)/ I --+ E with
Paradoxically, in our approach the proof of Theorem 6.1 is a bit more complicated. The crucial point is the following.
Lemma 6.B. Let E be an operator space that is also a Banach algebra with a unit e. Let p: E ® E --+ E be associated to the product map. Then, if lieli = 1 and if p defines a complete contraction from E ®h E into E, the mapping p introduced before the preceding lemma extends to a completely contractive unital homomorphism Pu from OAu(E) onto E.
6. Characterizations of Operator Algebms
117
Proof. Let Eo = C and Ej = E®·· ·®E (j times). Let x be a typical element in K®Tu(E). Then;1: can be written as;r = L;l=O Xj with ;1:0 = t ® 1 E K®C and with each x j expanded as a finite sum Xj = L::>j(O:)0ef® .. ·®ej a
with tj(O:) E K, ej E E, 1 :::; j :::; n. For el,"" ej E E, we denote simply below by ele2 ... ej their product in E. Note that h:®pu(x) = Lj La tj(a)® ele2' ... ej, so that the content of Lemma 6.8 is the inequality
L L tj(O:) ® (ele2' ... ej) j
(6.8)
a
This is somewhat reminiscent of von Neumann's inequality for polynomials in a contraction. Observe that (as in the proof of Theorem 6.2) we clearly have (6.8) if x is reduced to a "homogeneous" term of some degree, that is, if x = Xj for some j. Indeed, by iteration the product map p defines a complete contraction el ® ... ® ej ~ ele2 ... ej from E 0h ... 0h E into E, and E ®h ... ®h E = E ® f ... ® f E clearly naturally embeds completely isometrically into 0 Au (E). Thus we immediately get (6.8) in the homogeneous case. For the general case, the idea will be to replace a general element. x = L x j as above by another one y(N) that is homogeneous of degree N and satisfies moreover Ily(N)llmin :::; Ilxllmin and limN---+oo 11(h: 0 pu)(y(N))llmin = 11(1}( IZl pu)(x)llmin. The conclusion will then be easy to reach. Note that we will denote simply by II Ilmin the norms appearing on both sides of (6.8). This should not bring any confusion. Note that K ®min OAu(E) is an operator algebra embedded in the unit.al operator algebra B = B(£2) ®min OA,,(E). For convenience we introduce the following elements of B: We denote ~k
=
1 B (£2) ®~.
k-times Then we define l:!!.j(N)
=
1 N-j N- '+1 L~k'Xj'~N-k-j, J k=O
where the dot denotes the product in B. Observe that (6.9)
Introduction to Operator Space Theory
118
~
We claim that we have, for all N
n, (6.10)
To verify this let N
cp(t)
= (N + 1)-1/2 L
N
1jJ(t)
and
eikt(,k
= (N + 1)-1/2 L
k=O
Note that since B(H), we have
II!
Ilell =
cp(t)cp(t)* ::
1 we have
t(H)
e-ilt(,N_e.
(=0
$ 1
II(,kll
and
=- 1 for all
II!
k; hence, if we view B c
1jJ(t)*1jJ(t) ::
t(H)
$ 1.
(6.11)
Let us compute
(6.12) A simple verification shows that
y(N) = (N + 1)-1 k+j=e
O::;j::;n, O::;k,e::;N
(6.13)
=
t ~j(N) .(N;!: 1) . J=O
The preceding formula shows that y(N) is "homogeneous of degree N," that is, we have y(N) E K Q9 EN. As mentioned above after (6.8), Pu is clearly completely contractive when restricted to K Q9 EN. Therefore we have (6.14) and by (6.12) we have (by a variant of (1.12»
whence by (6.11) and (6.5) (6.15)
6. Characterizations of Operator Algebras
119
Finally, we have by (6.9) and (6.13)
(h @Pu)(y(N»
=~ n
(N-j+1) N + 1
(h @Pu)(Xj);
hence by (6.14) and (6.15) we obtain our claim (6.10). Then, from (6.9) and (6.10), we have
:s; lim sup lI(h @pu)(y(N»lImill :s; Ilxllmill' N->oo
min
This concludes the proof of (6.8) and hence of Lemma 6.8.
•
Proof of Theorem 6.1. Again (ii) =} (i) is immediate by Corollary 5.4. We turn to the converse. We denote by E the operator space underlying A. The proof is analogous to that of Theorem 6.2, but instead we need to know that, if (i) holds, the product map el @ ••• @ en ----> el . e2 ... en extends to a completely contractive unital map from OAu(E) to E. This is provided by Lemma 6.8. Given this fact, we conclude as before: Assume (i). The map Pu: OAu(E) ----> E yields after passing to the quotient by its kernel I a unital bijective homomorphism w: OAu(E)/ I ----> E with Ilwllcb :s; 1. As before, we have w- 1 = qi, where i: E ----> OAu(E) is the natural injection and q: OAu(E) ----> OAu(E)/I is the natural quotient map. Hence we conclude that IIw- 1 ll cb :s; IIqllcbllillcb = 1, that is, w is a completely isometric isomorphism from OAu(E)/I onto E. By Corollary 6.4, E is unit ally and completely isometrically isomorphic to a unital operator algebra .
•
Remark. We refer the reader to [BLM] for a detailed study of various possible operator algebra structures on "classical" examples such as £p or Sp (the Schatten p-class). Operator spaces that are also modules over an operator algebra (in other words, "operator modules") can be characterized in a similar way (see leES, ER9], and see also [Ma1-2] for dual modules). Let X c B(H) be an operator space and let A, B be two subalgebras of B(H). If X is stable under left (resp. right) multiplication by elements of A (resp. B), then X can be viewed as a concrete submodule of B(H) with respect to the actions of A and B by left and right multiplication. The next statement characterizes the "abstract" bimodules that admit such a concrete realization in B(H).
Theorem 6.9. (fCES}) Let A, B be C* -algebras and let X be an (A, B)bimodule, that is to say, X is both a left A-module (witl} action denoted
Introduction to Operator Space Tbeory
120
(a,x) ---4 a.x) and a rigbt B-module (witb action denoted (x, b) ---4 x.b), so tbat we bave a "module multiplication" map m: A x X x B ---4 X defined by m(a, x, b) = (a.x).b = a.(x.b). For simplicity we write m(a, x, b) = a . x . b. We assume tllat tbe sets {a.x I a E A,x E X} and {x.b I x E X,b E B} are dense in X. Tben, given an operator space structure on X, tbe following are equivalent: (i) m defines a complete contraction from A Q9h X Q9h B to X. (ii) Tbere exists a completely isometric embedding j: X ---4 B(11.) and representations PI: A ---4 B(11.) and P2: B ---4 B(11.) sucb tbat, for all x in X, a in A, and b in B, we bave j[m(a,x,b)] = PI(a)j(x)p2(b). Proof. Assume X C B(H). Applying Lemma 5.14 to the trilinear map m: A x X x B ---4 B(H), we obtain a complete contraction 0': X ---4 B(11.); representations 11"1: A ---4 B(11.), 11"2: B ---4 B(11.); and contractive maps V, W: H ---4 11. such that (\Ix E X, a E A, bE B) (6.16) Let PI (resp. P2) be the orthogonal projection onto span [11"1 (A)V H] (resp. span[1I"2(B)WH]). Let s = PlO'(a· X· b)P2 and t = pl11"1(a)O'(x)1I"2(b)P2' We claim that s = t. Indeed, it suffices to check that
for all
0:
in A, f3 in B, and h, k in H. But this equality is the same as
and indeed by (6.16) both sides are equal to (o:a· our claim that
X·
b(3)(h), k), which proves (6.17)
Note that PI (resp. P2) commutes with 11"1 (resp. 11"2) so that Pl(-) = PI1I"1(-) and P2(-) = 1I"2(')P2 are representations. Setting j(x) = PlO'(X)P2, we have IIjllcb ~ 1 and we obtain by (6.17)
j(a· X· b) = pl(a)j(x)p2(b). It remains to show that j is completely isometric. In the unital case this is immediate because (6.16) then allows us to write x = V*j(x)W, and since V, Ware contractive (and IIjllcb ~ 1), j must be completely isometric. In general, our density assumption guarantees that if (ai) (resp. (bk» is an
6. 911aracterizations of Operator Algebras approximate unit in the unit ball of A (resp. B), then ai . X· bk by (6.16) we have
121 ---->
x. Hence,
and hence we have IIxll ::; Ilj{x)1I and similarly for the matrix norms, so again we conclude that j is completely isometric. • Suitably modified versions of the Haagerup tensor product are available for operator modules (see [Bl\·IP, ChS, Ma2, AP, Pop]). Operator modules playa central role in [Bl\lP], where the foundations of a Morita theory for non-self-adjoint operator algebras are laid. 'See [BMP], [B4], Blecher's survey in [Kat], and references contained therein for more on all of this. Exercises Exercise 6.1. Let A be an operator algebra. Let E 1 ,. , • , EN be operator spaces. Then the product map A x ... x A ----> A induces a complete contraction
A fortiori, for N = 3, the product map PA: A 0h A contraction
---->
A defines a complete
Chapter 7. The Operator Hilbert Space Hilbertian operator spaces. In operator algebra theory and in quantum physics, numerous Hilbertian operator spaces have appeared. Let us say that a subspace E c B(H) is Hilbertian if it is isometric (as a Banach space) to a Hilbert space. We have already met the spaces Rand C, for instance. A more sophisticated example is the linear span of the Clifford matrices or, equivalently, of the generators of the Fermion algebra (this is sometimes called a spin system). Essentially the same example appears with the linear span of the creation (or annihilation) operators on the antisymmetric Fock space (see §9.3). A different example is the linear span of the generators of the Cuntz algebra 0 00 (see §9.4) and also the generators of the reduced C*-algebra C>:(Foo ) on the free group with infinitely many generators (this one is only isomorphic to £2; see §9.7) or the "free" analog of Gaussian variables in Voiculescu's "free" probability theory (see §9.9). Actually it is possible to show that there is a continuum of distinct (Le. pairwise not completely isomorphic) isometrically Hilbertian operator spaces. Furthermore (we will come back to this in Chapter 21), if n > 2, it can be shown that the set of all Hilbertian operator spaces of a fixed dimension n equipped with the "complete" analog of the Banach-Mazur distance is not compact, and not even sepamble. As mentioned in Chapter 3, Blecher and Paulsen [BPI J observed that any separable infinite-dimensional Hilbertian operator space 1{ sits "in between" the extreme cases min(l'2) and max(l'2), that is, we have isometric and completely contractive inclusions
Moreover, the interval between the two extremes is in some sense very "broad." Existence and unicity of OH. Basic properties. Despite the multiplicity of (different) examples of Hilbertian operator spaces, it turns out that there is a centml object in this class, that is, a space that plays tlle same central . role in the category of operator spaces as Hilbert space does in the category of Banach spaces. To motivate the next result, we recall that, if H is a Hilbert space, we have a canonical (Le. basis free) identification H=H*. Moreover, this characterizes Hilbert spaces in the sense that, if a Banach space E is such that there is an isometry i: E -+ E* that is positive (Le., i(x)(x) ~ 0 \Ix E E), then E is isometric to a Hilbert space. In the category of O.s. this becomes (see [PIj) Theorem 7.1. Let J be any set. Then there is an operator space OH(J)
7. The Operator Hilbert 8pace
123
tlwt is isometric (as a Banach space) to £2 (J) and which is such that tIle canonical identification £2(1) = £2(1)* induces a complete isometry from OH(I) to OH(1)*. AIoreover, OH(1) is the unique operator space up to a complete isometry witIl this property. Let K be any Hilbert space and let (li)iEl be any Ol·tllOnormal basis of 0 H (1). Tllen for allY finitely supported family (x;);El in B(K) we have (7.1 )
Notation. If J = N, we denote the space OH(1) simply by OH (we call it "the operator Hilbert space"). If J = {l, 2, ... , n}, we denote the space OH(1) simply by OHn . Note that any n-dimensional subspace of OH is completely isometric to OHn . To prove Theorem 7.1, it suffices to produce a norm 0: on K>g;£2(1) with a suitable self-dual property. The formula is extremally simple: \:Ix = LiE! Xi Q9 e; E K Q9 £2(1) (finite sum) with X;. E K, we set
o:(x) = IILXi Q9Xillll2 'El
mm
The proof that this is indeed an operator space structure on £2(1) with the required property is based on a version of the Cauchy-Schwarz inequality due to Haagerup ([HI, Lemma 2.4]), as follows:
(7.2)
valid for all Xi, Yi E K (or, more generally, for all Xi, Y;. E B(£2)). These forn1Ulas are one more illustration of the idea of "quantization" (see the discussion in §2.1l).
Proof of (7.2). By Proposition 2.9.1 we have (7.3)
where the supremum runs over all a, b in the unit ball of 8 2 8 2 ((2, (2) and where ( , ) denotes the scalar product in the Hilbert space 8 2, that is, (x;ayi, b) = tr(xiaYib*). Note that lIall2 S 1 iff a can be written as a = ala2 with trlall 4 S 1 and trla21 4 S 1, and similarly for b. Using this we can write
124
Introduction to Operator Space Theory
Hence, by Cauchy-Schwarz
but by Proposition 2.9.1 again I : Ilbixia111~
= I:tr(bixialaix;bl) = I:tr(Xi a l ai x ;b1bi)::; III:Xi ®Xillmill'
and similarly for the other term. Hence we finally derive (7.2) from (7.3) . •
Remark. The preceding proof also establishes the following two additional formulas: (7.3)'
III: Xi ® Xillmill
= sup {I: tr(xiax;b) I a ~ 0, b ~ 0, tra 2 trb 2
(7.3)"
III:Xi ® Xillmin
::;
::;
1,
1}
= sup { (I: lI,8xioll~) 1/2 I trlol 4
::;
1, trl,81 I}. 4 ::;
Proof of Theorem 7.1. We first show the "existence" of the space OH(I). Let K be a fixed infinite-dimensional Hilbert space. Let B = B(K) (actually, the proof works equally well if we set B = IC). Let us denote by C the set of all finitely supported families Y = (Yi)iEI in B such that IIEYi ® Yillmin ::; 1. Let Hy = K for all Y in C. For any fixed i in I, we define
Then, for any finitely supported family (Xi)iEI in B, we have
But by (7.2) we have (7.4)
since the supremum is attained for
125
7. The Operator Hilbert Space Thus we obtain
(7.5) In particular, if e E B is any element with Ilell supported familly of scalars (ai)iEI
= 1, we have for any finitely
Hence if we let E
= span[Ti liE JJ,
then E ~ €2(I) isometricaHy. We claim that E ~ E* completely isometrically. To verify this it suffices to check that, if (~i)iEI is the basis of E* biorthogonal to (Ti)iEI, we have for any x = (Xi)iEl as before (7.6) But by the definition of the complex conjugate (see §2.9) we have
where u: E ---> B is the mapping defined by u(x) = L (or, equivalently, U(Ti) = Xi for all i).
~i (x )Xi
for all x in E
By (2.1.9) (and by a density argument) we have lIulicb = sup {IILYiQ9 U(Ti) II}, where the sup runs over all finitely supported families (Yi )iE I in B with IILYi Q9 Tillmin ::; 1. Equivalently, lIullcb = sup {IILYi
Q9
Xiii
I YE C}.
Hence, by (7.4),
Thus (7.7) and (7.5) give us (7.6). This proves the existence ofa space with the properties in Theorem 7.1. We now address the unicity. Let F be an operator space isometric to €2(I) and such that F ~ F* completely isometrically. Let (Oi)iEI be any orthonormal basis in F. We will show that, for any (Xi)iEI as before, we have necessarily
This will show that the correspondence Ti ---> Oi is a completely isometric isomorphism, thus establishing the announced unicity. Let (T/i)iEI be the
126
Introduction to Operator Space Theory
basis of F* biorthogonal to (lJi)iEI. Reasoning as above, we see that F implies that, for all (Xi)iEI as before, we have
~
F*
where the supremum runs over all finitely supported families (Yi )iEI in B such that IIL:Yi ~ lJdlmin :::; 1. In particular, we must have by homogeneity
Taking Xi = Yi, this implies
In particular, IIL:Yi~lJillmin:::; 1 implies (Yi)iEI E C, and hence (7.8) and (7.4) imply the converse inequality
Thus, we conclude that for any finitely supported family (Xi)iEI in B we have
= B(I<)
Of course, this holds a fortiori when I< is finite-dimensional. Thus if we set OH(I) = E, we have checked all the properties in Theorem 7.1. • Definition. Let E be a (complex) vector space. We will say that a linear mapping J: E --+ E* (resp. i: E* --+ E) is ''positive definite" if it satisfies
'V.xEE
J(X)(X) ;::: 0
(resp.
'VeE E*
Remark. It is easy to see that a bounded linear operator J: E --+ E* (resp. i: E* --+ E) is "positive definite" iff there is a Hilbert space H and a bounded operator v: E --+ H (resp. v: H --+ E) such that J = tv v: E --+ H ~ H* --+ E* (resp. i = v tv: E* ~ H* ~ H --+ E). In the category of Banach spaces, the analog of Theorem 7.1 is the following obvious fact: If a Banach space E admits a positive definite isometry J: E --+ E* (resp. i: E* --+ E), then E is isometric to a Hilbert space, and J (resp. i)" is then surjective and coincides with the canonical (Le. basis free) correspondence between E* and E. Thus Theorem 7.1 says that the operator space E = OH(I) is characterized by the property that there is a positive definite map J: E --+ E* (resp. i: E* --+ E) that is a complete isometry. The next statement is mainly a useful reformulation of (7.1).
127
7. The Operator Hilbert Space
Proposition 7.2. (i) Let E be any operator space, I any set, and let v: OH(I) -+ E be a c.b. map. Let tv: E* -+ OH(I)* be its adjoint. Consider tile composition vtv: E* -+ OH(I)* -:.::= OH(I) -+ E.
We have then (7.9) (ii) Moreover, if (Ii) is an orthonormal basis of 0 H (I), we have
Ilvll~b = sup {II~ v(Td 0 V(Ti)11 tEJ
(iii) For any v: OH(I)
.I
J c I,
IJI < co} .
(7.10)
mm
-+
OH(I') (I, I' arbitrary) we have
IIvllcb = Ilvll.
Proof. (i) (Finite Case) Assume first that I is finite. Let ~i E OH(/)* be the biorthogonal basis to (Ti)' so that the mapping v corresponds to L ~i 0v(T;). Then, by (2.3.2) and (7.5), we have
Ilvllcb = 1II:~i 0 = III:Ti
0
tin 1II:~i v(Ti)tin V(Ti) tin = III: V(Ti) V(Ti)lI~i: .
V(Ti)
0
=
0
On the other hand, for any ~ in E* we have tv(~)
=
L(~,v(Ti))~;. Hence
and hence
which means that the map vtv: E* -+ E corresponds to the tensor L v(T;) 0 veT;). Thus, by (2.3.2) again we obtain
Ilvtvlleb =
III: V(Ti)
0 V(Ti)tin '
and we conclude by the first part of the prpof
= IIvll~b· (ii) Now let I be an arbitrary (a priori infinite) set. Let J C I be a finite subset and let E(J) = span[Ii liE J]. By a simple density argument, we have Ilvlleb = sup IlvIE(J) Ileb. J
Hence, by the first part of the proof we obtain (7.10).
128
Introduction to Operator Space Theory
(i) (General Case) Assume now that I is infinite but that E is finitedimensional. Then the series EiEI ~ Q9 v(Ti) and EiEI v(T;) Q9 v(T;} are convergent, respectively, in OH(I)* Q9 E and in E Q9 E, so that the preceding proof of (7.9) is easily extended to this case. Finally, in the general case, we can use Exercise 1.4 to show that (7.9) reduces to the case when E is finite-dimensional, which we just treated. (iii) By density, it suffices to check this when I is finite. Then, applying (i) with E = OH(I') we are reduced to showing Ilvtvllcb = IIvtvll. But now, since w = vtv is a non-negative Hermitian finite rank operator, it is diagonalizable; hence, for some orthonormal basis (Ti)iEI' in OH(I'), w is associated to a tensor of the form iEI'
with (Ai)iEI' finitely supported. The inequality to the following claim:
for any finite subset J
IIwllcb ::; IIwll
is thus reduced
c I'. But then, by (7.3)", we have
and by (7.10) applied to the identity map on OH(I') we know that
•
so we obtain the announced claim.
Corollary 7.3. Let I and J be arbitrary sets and let OH(I) and OH(J) be the corresponding operator Hilbert spaces. Then OH(I) Q9h OH(J)
~
OH(I x J)
completely isometrically.
Proof. For simplicity, we may and will assume that I and J are finite sets. It is easy to reduce the general case (by density) to that case. Let E = OH(I) and F = OH(J). Then, by Corollary 5.8, the following complete isometries hold: (E Q9h F)* '::! E* Q9h F* '::! E* Q9h F* ~ EQ9h F.
7. Tile Operator Hilbert Space
129
Clearly, the corresponding mapping E ® F --+ (E ® F)* is positive definite (by the same argument as the one showing that the tensor product of two scalar products is also one). Thus, by t.he preceding remark, we conclude that E ®h F is completely isometric to OH(I') with the cardinality of I' equal to dim(E ®h F), so that we can set I' = I x J. •
Remarks (i) By (7.1), the o.s.s. just defined on OH(1) is dearly independent of the choice of an orthonormal basis.
(ii) Thus, given a Hilbert space H, we may equip H with the o.s.s. corresponding to OH(I) with the cardinalit.y of I det.ermined by t.he Hilbertian dimension of H. Thus, Theorem 7.1 admits the following basis free ("canonical") reformulation: For any Hilbert. space H, there is a unique o.s.s. on H for which the canonical isometric identification H ~ H* becomes a complete isometry. \Ve will sometimes denote by Hoh the resulting operat.or space. (iii But actually, we will rarely lise the notation Hoh in the sequel. We prefer to make the following convention: We consider that t.his o.s.s. is the "nat.ural" one on a Hilbert space, so we always assume (unless explicitly stated otherwise) that. H is equipped with the o.s.s. of H oh .
(iv) In particular, if H ~ 1'2(1), t.he Hilbert. space S2(H)oh can be identified with OH(I x I), allowing us to view S2(H) as an operator space from now on. (v) It will be convenient to record here the following formula describing this o.s.s. on S2(H). For any x = L eij ® Xij in 1I/n (S2(H)) (Xij E S2(H)), we have (7.11)
where the supremum runs over all a, (3 in lIIn wit.h trlal 4 < 1 and trl(31 4 :::; 1. Indeed we have
i.,j
p,q
p.q
with Ypq = Li,j eijXij(P, q). Hence, since {e pq I p, q E I} is an orthonormal basis of S2(H), we have by (7.3)" .
Ilxllmin
= sup
(2: Ilaypq(3I1~)
1/2
p,q
= sup Iia . x
. (3lls2(e~02H)'
Introduction to Operator Space Theory
130
•
where the supremum is as above.
Lemma 7.4. Consider a, bin MN. We denote IIall4 = (trlaI 4)1/4 and IIbl14 = (trlbI 4)1/4. Let Ma,b: MN --+ sf)! be the operator defined by 1\fa,b(X) = axb, and let v: OHn --+ MN be any map. We llave tllen (note: here by sf)! we mean (Sf)!)oh) (7.12)
IIMa.bVIIHS :::; IIvll cbll a l1411bl1 4 IIMa,bllcb :::; IIal1411blk
(7.13)
More generally, the same estimates hold on any infinite-dimensional Hilbert space H, with Ma,b: B(H) --+ S2(H) defined in the same way with a, b in B(H) such tllat trlal 4 < 00, trlbl 4 < 00, and for any c.b. map v: OH --+ B(H). Proof. Let (Ti) be an orthonormal basis in OHn . We then have
IIMa,bVllHS = (l: IIMa,bV(Ti) 112) 1/2 = (l:
Ilav(Ti)bll~) 1/2.
Hence, by (7.3)" and (7.10), :::; IIall411bll411l: V(Ti) ®
V(Ti)ll~i: : :; Ila11411b11411vllcb.
This establishes (7.12). To verify (7.13), consider x = L, eij ®Xij in Mn(MN). We have (I®Ma,b)(X) = L,eij ®axijb. Hence, by (7.11), II (I ® Ma ' b)(x)IIMn (SN) = sup {11(a ® a) . ' " eij ® Xij . (/3 ® b)11 2 ~ S2(t 2ot f) x IIall4 :::; 1, 11/3114 :::; 1 } , which yields
:::; IIa®aI1411l:eij®Xijll Thus we obtain (7.13).
Af.,0mln M N
II/3®bI14:::; IIxILu,,(MN)II a ll41I blk
•
Finite-dimensional estimates. By a classical result due to Fritz John (1948), for any n-dimensional Banach space E, there is an isomorphism u: e~ --+ E such that IIuli = 1 and Ilu-111 :::;..;n. Equivalently, d(E, e~) :::; ..;n. In John's original proof, the mapping u is selected so that u(Be2 ) is the maximal volume ellipsoid included in BE. One can also majorize the projection constants: If E is an n-dimensional subspace of a Banach space X, it is known that there is a linear projection P: X --+ E with norm IIPII :::; n 1/ 2 (Kadec-Snobar, 1971; see IKTJl-2] for more references and more recent information) . In modern Banach space theory (see IP8, TJ1]), the following lemma due to Dan Lewis has turned out to be a very useful generalization of John's idea.
7. The Operator Hilbert Space
131
Lemma 1.5. Let E be an n-dimensional Banach space and let 0: be any norm 011 the space L(£2' E) of all linear maps from £2 to E. For any v: E ---> 1'2' we define o:*(v) = sup{ltr(vw)11 w: c~ ---> E, O'(w) ~ 1}.
Then there is an isomorpllism 11: Proof. Choose u:
£2
I det(u)1
--->
£2
--->
E with n (u)
E sucll that 0:( 11) = 1 and 0:* (11- 1 ) = n.
= 1 such that
= sup{1 det(w)11
'W:
1'~ --->
E. n(w)
~
1}.
Clearly det(u) -:I 0. By homogeneity, for any z in C and for any with DeW) ~ 1, we have
11':
f2
--->
E
Idet(u + zw)1 ~ Idet(u)I(1 + Izl)n. Hence, letting Izl
--->
0, we find
Idet(I + zu-lw)1 ~ (1 + Izl)n
=
1 + nlzl + o(lzl),
and since det(I + zu-l.w) = 1 + z tr(u-lw) +o(lzl), we obtain Itr(u-lw)1 ~ n .
•
Remark. If 0: is invariant under the unitary group U(n) (i.e., if we have O'(uw) = O'(u) for any n x n unitary matrix w), then the isomorphism u appearing in Lemma 7.5 is unique modulo U(-n); that is, if u\: f2 ---> E is another isomorphism with the same property, we must have u - \ U 1 E U (n). (In geometric language, the corresponding ellipsoid is unique.) Indeed, let U- l 11l = b. Let b = wlbl be the polar decomposition. By the choice of u ("maximal volume") we know that Idet(udl ~ Idet(u)I and hence I det(b)1 = Idet(lbl)l ~ 1. Therefore the eigenvalues AI, ... , An of Ibl satisfy IIAi ~ 1. Bllt, on the other hand, b- l = u 1l u, and hence L Ail = tr(b- l ) ~ 0'(u)a*(u 1l ) ~ n. Hence we must have (since the geometric mean of Ail is equal to its arithmetic mean) Ai = 1 for all i, which means that U-l.Ul is unitary. • Theorem 1.6. For any n-dimensionaJ operator space E and any E > 0, tIl ere is an isomorphism u: OHn ---> E with 1i11llcb = 1 such tlwt u- l : E ---> OH" admits for some integer N a factorization of tIle form E ~ ~MN ~O H n with Ilvlllcb = ..;n and IIv2licb < 1 + c. Proof. By the preceding lemma, applied with O'(u) = 1111llcB(OH".E)o there is an isomorphism u: OHn ---> E with Ilulicb = 1 and 0'*(u- 1 ) = n. In other words (see Chapter 4), the o.s. nuclear norm of u- 1 is equal to n. Therefore,
132
Introduction to Operator Space Tl1eory
for any 10 > 0 there is an N and a factorization of u -1 of the form u -1 as follows: VI II T sN W OHn u -1 : E -----+ 1. N -----+ 1 -----+
= WTV1
with IIv111cb = n 1/ 2, Ilwllcb < n 1/ 2(1 + c), and with T: MN ~ sf defined by T(x) = axb with a, b in the unit ball of sf. If we wish, we may as well assume (by the polar decomposition) that a ~ 0 and b ~ O. Then, we define T 1: MN ~ sf and T 2: sf ~ sf by T 1(x) = a 1/ 2xb 1/ 2 and T 2(y) = a 1/ 2yb 1/ 2, so that T = T 2T1. Consider the composition T1V1: E ~ sf. By Lemma 7.4, we have (7.14)
Let E2 = T1V1(E) c sf. Clearly, dimE2 = n (since u bijective implies T1V1 injective). Thus we may (and will) identify E2 with OHn . Let P: sf ~ E2 be the orthogonal projection. Clearly (see Proposition 7.2(iii)), 1IPlicb = IIPII = l. Recapitulating, we have
Now let W: E2 ~ OHn be the restriction of wT2 to E2 and let V PT1V1: E ~ E2. We may then write
u- 1 = WV, and V, Ware both invertible. By Lemma 7.4 (note that, with the proper identification, T:; = T 1 ) we have
and by (7.14) we also have
We claim that this implies IIW* - VullHS < fn(c)
(7.15)
with fn(c) ~ 0 when 10 ~ O. Indeed, using the scalar product of the HilbertSchmidt norm, namely, (x, y) = tr(xy*), the identity tr(WVu) = tr(u- 1u) = n means that (W*, V u) = n. Therefore we have IIW* - Vullhs
= IlVullhs + IIW*lIhs -
2 Re((W", Vu})
< n+n(l +10)2 - 2n,
from which the above claim becomes obvious. Since W(Vu) = I, (7.15) implies
7. The Operator Hilbert Space
133
hence, a fortiori IIWW* - III < f~(c:) with f~(c:) ----+ 0 when c: ----+ o. Therefore, a look at the polar decomposition of lV shows that there is a unitary operator U: E2 ----+ OHn such that
IITV - UII < f~(c:) with f;:(c:)
----+
0 when c:
----+
o.
We can now finally conclude: We define V2: AIN ----+ OHn by '1)2 = lVPTI. Note u- 1 = V2Vl. By Lemma 7.4, we have IITlllcb ~ 1; on the other hand, by Proposition 7.2(iii) and the preceding bound,
IITVllcb = IIWII < 1 + f~(c:)· Thus we finally obtain
• Corollary 7.7. Let E c B(H) be any n-dimensional operator space. There are an isomorphism u: OHn ----+ E with Ilulicb = 1 and Ilu- 1 l1cb ~ fo alld a projection P: B(H) ----+ E with 1IPlicb ~ fo. 111 particular,
Moreover, for any c: > 0, there is an integer N and a subspace that dcb(E, E) ~ (1 + c:)fo.
E cAIN such
Proof. By the preceding statement, for any c: > 0, there is a map Ue: OHn ----+ E with Iluelicb = 1 and lI(u e)-lllcb < fo(l+c:). By norm-compactness, the net (u e ) admits a cluster point U when c: ----+ o. Clearly, U has the desired property. Note that, by the extension property of A[N, the mapping VI appearing in Theorem 7.6 admits an extension 11t: B(H) ----+ MN with IIvdlcb = Ilvll1cb. Therefore, we may assume that (u e )-1 admits an extension Ve: B(H) ----+ OHn with IIvelicb < fo(l + c:). The mapping P e = UeVe: B(H) ----+ E is then a projection with IIPelicb ~ fo(l + c:). Taking a cluster point of (Pe ) for the pointwise convergence on B(H), we obtain a projection P with the desired property. Finally, with the notation in Theorem 7.6, let E = 'lh(E) C M N . Then the restriction of UV2 to E is the inverse of VI: E ----+ E, and hence dcb(E, E) ~ IIvIilcblluv211cb ~ IIVl11cbllv211cb ~ (1 + c:)fo. •
Remark. Fritz John's result can be recovered from Corollary 7.7 as a special case: Indeed, for any n-dimensional normed space E we have d( E, e~) ~ dcb(min(E), OHn) ~ fo.
134
Introduction to Operator Space Tlleory
Remark. For any operator space X C B(H), we define the c.b. projection constant of X as follows:
Acb(X) = inf{llPllcb
1
P: B(H)
-+
X, projection onto X}.
By Corollary 1. 7, this is invariant under a completely isometric isomorphism. With this notation, Corollary 7.7 implies that any n-dimensional operator space E satisfies Acb(E) :S In particular, Acb(OHn ) :S Surprisingly, however, this estimate is far from optimal. Indeed, Marius Junge [J2] proved quite recently that Acb(OHn ) was O( VnjLog n) when n -+ 00, and the paper [PiS] already showed that Acb(OHn ) was not asymptotically smaller. Thus, there is a constant J( > a such that, for any n, we have
,;no
,;no
See [KT Jl-2] for recent results on the Banach space analog of n-dimensional projection constants. Theorem 7.8. Let Tn: Rn -+ C n be tlle linear mapping taking For any n-dimensional operator space E, tllere is a factorization
eli
to
eil'
,;no
of Tn tllrougll E witllllwllcb = 1 and IIvllcb = Moreover, tllis is unique in tlle sense tllat if VI, WI is anotller sucll factorization, tllere is a unitary U on e~ c::: Rn sucll tllat WI = wU. Proof. We will apply Lemma 7.5 with o(w) = ilL eil 0 w(ei)lIcn®h E' Given in E*, let v: E -+ e~ be the mapping defined by v(e) = L~ ~i(e)ei' Then, by (5.14), since C~ c::: R n , we have o*(v) = ilL eli 0 ~iIlR"®I,E" By definition of 0h (and by homogeneity), o*(v) :S n iff we can write
6, ... '~n
,;n
with aik E C, T/k E E* such that (Li k laik 12) 1/2 :S and ilL ekI 0 T/k II min = Now let u: e~ -+ E and u . .\ e~ -+ E be as in Lemma 7.5. Let w: Rn -+ E be defined by w(eli) = u(ej). By (2.3.2) and Exercise 5.3, we have IIwllcb = o(u) = 1. Let~; E E* be such that u-I(e) = L~i(e) 0 ei' Let a = (aik) and (T/k) be as above. We then have
,;no
Vi
~i = Lk aikT/k·
By polar decomposition, we may assume without loss of generality that a is hermitian 2': O. Let v: E -+ en be defined by v(e) = LT/k(e)ekI (e E
7. The Operator Hilbert Space
135
E).
Then Ilvllcb = IILek101Jkllmin =.;n. Hence, by (1.5), Ilvwllm, = IlvwllcB(R",C,,) ~ IIvllcbllwllcb ~ .;n. Note that u-1u = I implies ~i(Xj) = Oij' Hence Lk aik1Jk(Xj) = oij. Let bkj = 1Jdxj). We then have b = a-I but also n = tr(ab) = (a*,b) and IIul12 ~.;n, IIbl1 2 ~ .;n. Thus n- 1/ 2b norms n-l/2a* in S"2. Hence we must have b = a*. Since b = a-I, a must be unitary, and since a ~ 0, we must have a = I. Thus we conclude that r,dxj) = Okj, so that vw = Tn. The unicity follows from the rem,ark appearing after Lemma 7.5 . •
Notation. In the direct sum RnEBCn , we consider the vectors c5i =. eliEBe;l. In the style of §2. 7, we will denote by Rn n C n the linear span of {c5i 11 ~ i ~ n}. More generally, let E be any operator space. We will denote by E n EOP the operator subspace of E EB EOP formed of all vectors of the form e EB e (e E E). Similarly, we may consider the mapping Q: E EBl EOP -+ E taking (x, y) to x + y. We will denote by E + EOP the operator space (E EBI EOP)/ ker(Q). Corollary 7.9. ([Jl]) For any n-dimensional operator space E, dcb(E, Rn n Cn) ~ 2.;n. Proof. We apply Theorem 7.8 to the operator space E factorization· Tn: Rn~E+EoP~Cn
+ EOP:
We have a
with Ilwllcb = 1, Ilvllcb = .;n. Since E -+ E + EOP and EOP -+ E + EOP are completely contractive, we have Ilv: E -+ Cnllcb ~ .;n and Ilv: E -+ Rnllcb = IIv: EOP -+ Cnllcb ~ .;n and hence Ilv: E -+ Rn n Cnllcb ~ ..;n. In the preceding, we of course think of R n , C n , and Rn n Cn as having £2 as commOI1 underlying vector space, via the obvious identifications. On the other hand, let w(eli) = Xi E E. Since Ilw: Rn -+ E + EOPllcb ~ 1, we can write Xi = ai +b; (ai, bi E E) with IlL eil 0 aillc,,0m;n E ~ 1, ilL: eil 0 bill c,,0n,;u E "" ~ 1. Hence, if we assume E c B(H), we have IIL:a;adI 1/2 ~ 1 and IIL:bibiII1/2 ~ 1. Therefore, for any Ci in J( we can write
hence, by (1.11),
~ IlL aiai 1 1/211L c;cill '/2+ IlL bibiII1/21ILaia;II'/2 ~ 211Lc5; 0 Cill, and we conclude IIw: Rn n C n -+ Ellcb ~ 2. Thus if we define u(8;) = w(eh), we find lIu: Rn n Cn -+ Ellcb ~ 2 and we have Ilu-l: E -+ Rn n Cnllcb = II v: E -+ Rn n Cnll cb ~ .;n, so that finally Ilullcbliu-11lcb ~ 2.;n. •
Complex interpolation. After we observed the existence of OH(l), we developed in [PI] the theory of these spaces in analogy with what is known in Banach space theory. In these developments, complex interpolation has played a crucial role. For instance, we have
136
Introduction to Operator Space Theory
Theorem 7.10. Let E be an arbitrary operator space equipped with a continuous injection i: E* -+ E with which we may view the couple (E*, E) as compatible. Assume tllat i is ''positive definite," that is to say, such that i(~)(e) ~ 0 for all in E*. Then (E* , E)! is completely isometric to 0 H (1) for some index set I.
e
Proof. In the first part of the proof we treat E as a Banach space. We denote Eo = E* and El = E. Note that, by Lemma 2.7.2, we have an isometric identity (7.16) Let Fo
= E and Fl = E*.
We apply Lemma 2.7.5 to the sesquilinear mapping
u: Eo n El x Fo n Fl
-+
C
defined by u(~, i(r;» = i(~)(1J) or, equivalently (since ~ and i(~) are identified by the "compatibility"), u(i(~),1]) = i(~)(1J). Thus,
lu(i(~), 1])1
::;
Ili(~)IIEII1JIIE·.
Moreover, by the polarization identity, positive definiteness implies selfadjointness, that is, we have i(~)(1J) = i(r;)(e). It follows that, if x, y belong to Eo n El = Fo n F 1 , we have simultaneously
(Note that we do not assume that i is contractive.) Let X = (Eo,Elh/2 = (E*, Eh/2. Note that by symmetry X = (Fo, F 1 h/2. By Lemma 2.7.5, u . extends to a contractive sesquilinear form (still denoted by u) on X x X. Hence we have v x,y E X I'l/(x, y) 1::::; Ilxlix lIyllx. But for all x = i(~) in E* we have u(x, x) = i(~)(e) ~ o. Hence, since i(E*) is dense in X, we have u(x,x) ~ 0 for all x in X, so that u is a scalar product on X. Taking x = y, we obtain VXEX
U(X,X)1/2::;
IIxllx.
(7.17)
We now claim that we have a "natural" isometric embedding X -+ X*. Indeed, by (2.7.6), we have an isometric embedding of (Eo,Eih/2 into X* or, equivalently, of (Eo, Eih/2 into X*. But Eo = E** and Ei = Eo; hence (7.16) ensures that (Eo, Eih/2 = X and we indeed obtain Xc X* isometrically. In particular,
v x = i(~) E i(E*) IIxlix = sup{l(e,y}11 y E i(E*) IIYllx ::; I}.
7. The Operator Hilbert Space
137
Hence (7.17) implies by Cauchy-Schwarz 'V x E i(E*)
IIxllx::; sup{lu(x,y)11 u(y,y) ::; 1} ::; U(X,X)I/2.
Thus we conclude finally that IIxllx
= U(X,X)I/2,
which shows that X is isometric to a Hilbert space. \Ve now turn to the operator space case. By Theorem 2.7.4 (and the above observation (7.16) that takes care of the bidual E** and replaces it by E) the inclusion X ---+ X* is completely isometric, and since X is Hilbertian by what precedes, it is actually surjective. Thus, by the unicity part of Theorem 7.1, we conclude that X must be completely isometric to OH(I) for a suitable • set I. Remark. See [Watl-2] for a more general discussion of the complex interpolation space (with () = 1/2) between a Banach space and its antidual. Corollary 7.11. We have completely isometrically (R,Ch2
= OH
Remark. Let H be a Hilbert space. Recall (see (1.7)) that we have denoted by He (resp. Hr) the space H equipped with the column (resp. row) O.S.s. In the same vein, we sometimes denote by Hoh the space H equipped with the o.s.s. corresponding to the OH-space of the same (Hilbert ian) dimension as H. \Vith this more canonical notation, the first part of Corollary 7.11 says that we have a complete isometry
More explicitly, we have Corollary 7.12. Fix n with the norm
~
1. Let Ao = B(H)n (resp. Al = B(H)n) equipped
lI(xl> ... ,xn)llo = IIEx;x:III/2 (resp.lI(xI,oo.,xn)III
=
IIEx; Xi
l
l
/
2 ).
Then (Ao,Ad! coincides with B(H)n equipped with the norm
Introduction to Operator Space Tlleory
138
Proof. It is easy to reduce to the case H = 1!.2. Then, by (2.7.3), Corollary 7.11 implies (K ®min Rn,K ®min Cnh/2:= K ®min OHn· On the other hand, for any finite-dimensional operator space E, by Exercise 5.6 we have (isometrically)
(/C ®min E)** := B(1!.2)
®min
E.
Hence, by Theorem 2.7.4, we have isometrically
(B(1!.2)
®min
Rn, B(1!.2)
®min
Cnh/2 := B(1!.2)
which is the content of Corollary 7.12.
®min
OHn,
•
This has been generalized to the case when B(H) is replaced by a general von Neumann algebra ([P7]). These ideas have applications to the study of completely bounded projections P: M -+ N from a von Neumann algebra M onto a subalgebra N C M. See Chapters 15 and 23. Vector-valued Lp-spaces, either commutative or noncommutative. Let (0, J.L) be a finite measure space. The natural inclusion map Loo(J.L) -+ Ll (J.L) allows us to view the pair (Loo (J.L), Ll (J.L)) as compatible in the most classical sense. We feel it is worthwhile to first explain why the classical isometric identity (L oo (J.L),L 1 (J.L))1.2 := L 2(J.L) follows from Theorem 7.10. For any y in Loo(J.L), let
In the operator space case, exactly the same argument gives us (using Exercise 3.2) (min(L oo (J.L)),max(L 1 (J.L))h/2:= L 2(J.L)oh. We now turn to the noncommutative case. Let M be a von Neumann algebra, equipped with a faithful, normal, finite trace r. The space Lp(r) (1 ~ p < 00) is defined as the completion of M with respect to the norm IIXllLp(r)
= (r(lxIP»l/p.
7. The Operator Hilbert Space
139
In the case p = 1, it is well known that Ll (r)* ::: ]I.{ (isometrically) with respect to the duality (y,x) = r(xy) (x E L 1 (T),y E 1II). For any y in 1\.1, let t.py E Ll(T)* be the functional now defined by t.py(x) = T(XY). Let i: Ll(r)* -------> Ll(r) be the linear mapping defined by i (t.py)
= y*.
Again i is clearly positive definite, so that, if the "compatibility" is defined using i, we have
Moreover, since the (linear) mapping Uo: Ll (r)* ~ 111 defined by uo( t.py) = y* is clearly isometric, we may invoke Remark 2.7.8 again and we obtain
where the compatibility of the pair (]I.{, Ll (r)) is now defined using the natural inclusion of 111 into Ll (r). More generally, we have
Clearly, we can extend all this to operator spaces as before. There is however an important point that must be emphasized. \Ve must equip Ll (r) with the appropriate o.s.s., namely, the one for which Uo is completely isometric. Unfortunately, this point is insufficiently explained in [P2] (actually page 37 in [P2] is definitely misleading) and I am very grateful to M. Junge and Z. J. Ruan for pointing this out to me, thus allowing me to correct this "loose end" here. So we want Uo: Ll(T)* ~ 111 to be completely isometric. Equivalently, we want Uo: Ll(r)* ~ 111 (defined by uo(t.py) = y*) to be completely isometric, and since lIJ ::: lIfO P completely isometrically (via the map y ~ y*), this means that we want Ll (r) * ::: lIIop, completely isometrically, via the mapping taking t.py to y. In other words, we must equip Ll(r) with the o.s.s. of (M*)OP. Provided we make this convention (and we will stick to it throughout) everything ticks just as before and Remark 2.7.8 yields
(M, L 1(T)h/2 ::: L 2(r)oh
(completely isometrically).
Although we assumed r finite (for simplicity) in the preceding discussion, the situation is the same if r is only assumed semi-finite, which means that the subspace V = {x EM I T(lxl) < oo} is weak-* dense in AI. The space L 1(T) can then be defined as the completion of V with respect to the norm x ~ r(lxl). Let P be the collection of all projections in M with finite trace. We can still view the couple (wI, Ll (r)) as compatible using the natural inclusions
Me EBpMp pEP
and
Ll(r)
C
EBL1(TlpJ\;/p)· pEP
140
Introduction to Operator Space Theory
Then the preceding discussion remains valid (see [Watl-2j for more information). Consider in particular the case when M = B(£2) equipped with the usual (semi-finite) trace denoted simply "tr". We prefer to denote the resulting noncommutative L 1-space by 8 1 . This is the "trace class," formed of all the compact operators x on £2 such that !lxllsl = tr(lxl} < 00. More generally, if 1 ::; p < 00, when r = tr the space Lp(r) is the Schatten class 8 p of all compact operators x such that trlxl P < 00, equipped with the norm (7.18)
When p = 2,82 is the classical "Hilbert-Schmidt class." The previous discussion leads to
or equivalently, by Lemma 2.7.2, (811 K,h/2 c:= (82 )oh
(completely isometrically),
where the compatibility is meant with respect to the natural inclusion 8 1 C K,.
Remark 7.13. As is well known, we have K,* c:= 8 1 isometrically. In this isometric identity, the duality can be defined by setting either
(x, y)J = tr(xy)
or
(x, y)Il = tr(xty).
In other words, this gives us two distinct isometric identifications between K,* and 8 1 . Note that the resulting two o.s. structures on 8 1 (for which K,* c:= 8 1 becomes completely isometric) depend on the choice of this duality: Each one is the opposite of the other. But since K, c:= K,op (completely isometrically, via x --+ t x ), the resulting operator spaces are completely isometric. In the preceding discussion, we chose to work with (-'.)J, but actually, since we equip 8 1 with the o.s.s. for which (K,*)OP c:= 8 11 the resulting o.s.s. on 8 1 is exactly the same as the one that we would obtain on 8 1 by requiring that the isomorphism K,* c:= 8 1 associated to (., -)Il be completely isometric (which is the convention adopted in [P2, §1]). The preceding identities allow us to equip Lp(r) with an o.s.s. More generally, the author has developed ([P2]) a theory of noncommutative vector-valued Lp-spaces, which extends the so-called Lebesgue-Bochner theory of the spaces
where 1 ::; p ::; 00, (fl, J.l) is a measure space and E is a Banach space. Recall that, in the discrete case, that is, when (fl, J.l) is just the integers (resp.
7. The Operator Hilbert Space
141
= {I, ... , n}) equipped with the counting measure, Lp(Q, p; E) is usually denoted by Cp(E) (resp. C; (E». In this new theory, E has to be equipped with an operator space structure. Then, using interpolation, the spaces
Q
can also be given a natural operator space structure. When p = 2 and E = C this reduces to the theory of the already mentioned space OH.
In the noncommutative case we replace Loo(Q,p) by an injective von Neumann algebra M C B(rt). We assume that l\f admits a faithful normal semi-finite trace T so that, as before, the predual 111* can be identified with Ll (T) in the usual way (cf., e.g., [Di3, Se, Ne, Ta3]). Then let E C B(H) be an operator space, and we set
o (T; E )def Loo = M
@min
E.
Note that, in the commutative case, this space is in general smaller than the classical space Loo(Q, It; E) in Bochner's sense but coincides with it if E or M = Loo(Q, p) is finite-dimensional. The case p = 1 has already been treated by Effros and Ruan, who introduced more generally the analog of the projective tensor product, denoted by @A. We simply set L 1 (T;
E)~fLl(T) @A E.
Then we can define (following [P2])
(7.19) Of course, when E space Lp(T).
= C, the space Lp( T; E) reduces to the previously defined
For example, consider 111 = B(C2) equipped with the usual trace (somewhat discrete) "tr". Then, if p < 00, Lp(T) is the Schatten class Sp of all compact operators such that trlTIP < 00 equipped with the norm defined in (7.18). In this particular case, the space Lp(T; E) is denoted by Sp[E]. The space Sp (resp. Sp[E]) is analogous to a discrete or atomic Lp-space. It is analogous to Cp or C; (resp. Cp(E) or e;(E» in the commutative case. These spaces satisfy all the basic properties of the classical Lebesgue-Bochner spaces Lp(Q, p; E), such as duality, interpolation, injectivity, and projectivity relative to the space of "values" E. The reader should observe that our theory yields an isometric embedding Sp
c B(H)
Introduction to Operator Space Theory
142
that is highly nonstandard! Just as in the case of OB (when p = 2), we have no explicit description of this embedding. We will only give here a brief introduction to that subject and refer the reader to the extensive text [P2] devoted entirely to this topic. To avoid any technicality, we will discuss here only the spaces S;[E]. The spaces Sp[E] behave very much the same, and in fact the embeddings ... c S;[E] c S;+1[E] c ... are completely isometric, so that Sp[E] can be viewed as the completion of the union of the spaces S;[E]. Throughout this discussion, E will be an arbitrary operator space. The normed space S;[E] can be described as follows. It is the same vector space as Mn(E) but equipped with the following norm (1 :::; p < 00): For any x in Mn(E) we set (7.20) where the infimum runs over all possible factorizations of x of the form x = a· x· b with a, bE Mn and x E Mn(E). Actually, the same definition (7.20) makes sense for Sp[E], but Sp[E] is now defined as the subspace of K ®min E formed of all the x that can be written as x = a . x· b with a, b E S2p and x in K ®min E. It is natural to denote Soo = K and Soo[E] = K ®min Fj clearly the elements ofSoo[E] can be viewed as bi-infinite matrices with entries in E. This definition is reminiscent of the obvious fact that (
L
IIYnll~ )
l/P
= inf{llclle"lIylle",,(E)}
over all factorizations of the form Yn = cnYr;. But here the noncommutativity forces us to consider separately both left and right multiplications, and thus the space S!jp makes its appearance (note that Ilells;; :::; 1 iff c can be written c = ab with Iialls"2p :::; 1 and Ilblls"21) < 1). If x E Spn[E] is a diagonal matrix, then we find simply Ilxlls;:[Ej isometrically.
= (I: Ilxiill~)l/p, so that e;(E) c S;[E]
In the case p = 1, the norm appearing in (7.20) coincides with the norm of Sf ®" E as defined in Chapter 4. Therefore we have Sf[E]* = Mn(E*) isometrically. More generally (see [P2j) we have, for any 1 < p < 00, S;[E]* = [E*] isometrically. Equivalently, we have another description of the norm of S;[E] as follows:
S;,
IIxlls;:[Ej
= sup{lIa· X· bllsf[Ej},
(7.21)
where the supremum runs over all a, b in the unit ball of S!Jv. Thus, with (7.20) and (7.21), we have two a priori different ways (dual to each other) to look at the spaces S;[E] (or Sp[E]), and this is (to a large
143
7. The Operator Hilbert Space
extent) the reason why their theory works so nicely. Note that (7.20) is projective (if E is a quotient of F, then S;[E] appears as a quotient of S;[F]) while (7.21) is injective (if E c F, then S;[E] c S;[F]). Having defined the norm in the spaces S:[E], we can now define their o.s.s. as follows: For any x in .lI1n (S:[E]) we define Ilxll n = sup{lla· x· blls;:N[EJ}, where the supremum runs over all a, b in the unit ball of S~p and where we have identified .lI1n (.lI1N (E)) with .lI1nN (E). Then it can be checked that this sequence of norms satisfies Ruan's axioms (Rt) and (R2); hence it defines an o.s.s. on St'[E] (and also on Sp[E]). With this structure, if 1 < P < 00, we have (see [P2]) completely isometrically:
Since e;(E) c S;[E] (and epee) c Sp(E]), the preceding structure induces an o.s.s. on e;(E) and epee). More generally, if (0,11) is any measure space, we can equip Lp(O, 11; E) with an o.s.s. by introducing the following norm on .lI1n (Lp(O,/I;E)): Viewing an element x in 1IIn (Lp(0,/L;E)) as one in Lp(O, JL; Mn(E)), we let (7.22)
Ilxll n = sup{lla· x· bll L,.(fl.JL:S;: [E])}, where the supremum runs over all a, b in the unit ball of
S~p.
The resulting o.s.s. on Lp(O, JL; E) is called the natural one. It coincides with the one described above using complex interpolation in (7.19). The introduction of these vector-valued (actually operator space-valued) noncommutative Lp spaces opens the way for noncom mutative analogs of a number of natural questions originating in either Banach space theory, classical analysis, or probability. To name a few of those: \Ve can study uniform convexity and the Radon-Nikodym property, maximal inequalities for martingales or the unconditionality of martingale differences (U1'lD) for operator spaces. One can also ask which Fourier (or Schur) multipliers are com.pletely bounded on L p , and this is of interest already for the most classical groups such as Z or '['. We will return to this subject in §9.5, and we illustrate it with "concrete" examples in the subsequent §§9.6-9.9. See [P2] for a full treatment. To give only a sample result, recall that, among Banach spaces, Hilbert space is characterized by the Jordan-von Neumann (parallelogram) inequality
\/x,Y
E
E
144
Introduction to Operator Space Theory·
In other words, if we denote by
T: £~(E)
---->
£~(E)
the operator defined by
then E is isometrically Hilbertian iff IIT1! ::; 1. In our theory, if E is an arbitrary operator space, the space £~(E) (denoted by E(f32E in Remark 2.7.3) is also equipped with a natural operator space structure (this is a particular case of (7.22), with (0, /-l) a two-point space). For that structure we have
Theorem 7.14. set I.
IITllcb::;
1 iff E is completely isometric to OH(I) for some
We end this chapter with a surprising estimate (improving a bound in Theorem 3.8) that uses the space OH. This is due to Eric Ricard.
Theorem 7.15. ([Ril}) Let E be any n-dimensional operator space. Then
IIIE: min(E)
--+
max(E)lIcb ::; n/2 1/ 4 •
Equivalently, a:(n)::; n/2 1/ 4 • Proof. By a well-known result on 2-summing operators (cf., e.g., [P4, p. -1
15)), IE can be factorized as E~£2~E with 1T2(U) By a well-known estimate, we have lIuIlCB(min(E),R,,) ::;
1T2(U) =
= 1 and
lIu- 1 11
= .;n.
1,
and similarly IluIICB(min(E),C,,) ::; 1;
hence
On the other hand, the identity map on £'2 (denoted below by I) defines a completely contractive inclusion I: Rn n en --+ OHn. This follows from Corollary 7.11, but it can also be verified directly as follows. Let (Ti) denote an orthonormal basis in OHn . Then, for any n-tuple (Xi) in B(£2), we have by (7.5) and (2.11.3)
7. The Operator Hilbert Space
145
hence a fortiori
which proves our claim that III: RnnCn -> OHn Ilcb ~ 1. Furthermore, we also have (by (7.9) and Exercise 3.2) III: OHn -> max(£2)llcb = III: min(t2) -> max(£2)11!~2. Hence, by Theorem 3.8
Finally, we have lin-I: max(£2) conclude
->
max(E)llcb
=
lin-III
• Corollary 7.16. ([Ril}) An'y linear mapping spaces E, F with rank ~ n satisfies
11:
E
->
F between operator
The best constant (instead of 2- 1/ 4 ) in the preceding bound is not yet known.
Remark. An important and classical fact in Banach space theory is that £2 embeds isometrically into L 1 . This embedding can be obtained by mapping an orthonormal basis to a sequence of independent complex Gaussian random variables (see §9.8). It is natural to wonder about the o.s. analog of this fact for the space OH. This question was recently solved by l\'1arius .Junge [J2]. Curiously, we cannot use Gaussian variables, either "classical" or "free" (see §§9.8 and 9.9) to embed OH completely isomorphic ally into an L)-space, either commutative or not. Nevertheless, Junge [.J2] proved that OH embeds completely isomorphically into the predual 11[* of a von Neumann algebra 11[. The completely isometric case remains open (at the time of this writing). Actually, Junge proved more generally that any quotient of RtBC embeds into l'vf* for some 11[. Since, by Exercise 7.9, OH embeds completely isometrically in a quotient of R tB C (this fact has been known for some time independently to Junge and the author), we recover the embedding of OH as a corollary. Junge's construction uses for 11/ a von Neumann algebra "of type III," that is, one that does not admit any semi-finite and faithful trace. After learning of Junge's results, we were able to show (using [PiS]) that "type III" cannot be avoided: If OB C M* completely isomorphically, then 1I[ cannot be semi finite (see [P22]).
146
Introduction to Operator Space Theory Exercises
Exercise 7.1. Let E isometrically.
= OH(I). Prove that E
~
EOP and E
~
E completely
Exercise 7.2. In a suitable realization of OH(1) there is an orthonormal basis «()i)iEI of OH(J) formed of self-adjoint operators. Exercise 7.3. For any orthonormal basis (Ti)iEI of OH(1) (with have
III =J 0)
we
~~~ III:Ti ® Til = 1.
IJI
tEJ
Exercise 7.4. Let E = OH(1). Show that Ilull = Ilulicb for any u: E ---- E, so that C B (E, E) ~ B (£2 (1» isometrically. Show that this identity is not completely isometric. Exercise 7.5. Let E have
= OH(1).
IIxIIM,,(E)
=
Show that for all x
= (Xij)
in Mn(E) we
11«(xij,Xke)lli~~2
or, more precisely,
I: (Xij' Xke)eij
1/2
® eke
i,j,k,e
min
Exercise 7.6. Let E be any operator space. Then, for any hI"'" hn in OH and any Xl, ... ,Xn in E, we have 1/2
II I:h'®X·11 OH0mlu E t
t
i.,j
min
Exercise 7.7. Let (Tlo ... , Tn) be an orthonormal basis in OHn . Compute and
IIL:
n
1
eli
® ~II
R ... 0mln OH"
.
Let u: Cn ---- OHn (resp. v: Rn ---- OHn) be any isometric map. Show that
Ilulicb = IIvllcb = nl / 4 • Exercise 7.8. Let E C B(H) be any n-dimensional operator space. For any x = (Xi)i
7. Tile Operator Hilbert Space
147
Then we define for 0 < () < 1 IlxllE(I;l)
=
Il x ll(E(O),E(1))n'
Show that there is a basis (Xi) in E with biorthogonal basis that
(~i)
in E* such
In particular, for () = 1/2, deduce from this a new proof that dcb(E, OHn) :::; y'n. Hint: Use Theorem 7.8 applied to (E, EOP)o, like in the proof of Corollary 7.9. Exercise 7.9. Show that OH is completely isometric to a quotient of a subspace of R EB C.
Chapter 8. Group C*-Algebras. Universal Algebras and Unitization for an Operator Space We first recall some classical notation from noncommutative abstract harmonic analysis on an arbitrary discrete group G (we restrict ourselves to the discrete case for simplicity). Let 7l': G --+ B(Ji) be a unitary representation on G. We denote by C;(G) the C*-algebra generated by the range of 7l'. Equivalently, C;(G) is the closed linear span of 7l'(G). In particular, this applies to the so-called universal representation of G, a notion that we now recall. Let (7l'j)jEI be a family of unitary representations of G, say,
in which every equivalence class of a cyclic unitary representation of G has an equivalent copy. Now one can define the universal representation Uc: G--+ B(Ji) of G by setting Uc
= tBjEI7l'j on Ji = tBjE1Hj .
Then the associated C* -algebra CUG (G) is simply denoted by C* (G) and is called the C*-algebra of the group G (so this is the closed linear span of {Uc(t) I t E G}). We will sometimes call it the full C* -algebra of G to distinguish it from the reduced one, which is described in the following. Let 7l' be any unitary representation of G. By a classical argument, 7l' is unitarily equivalent to a direct sum of cyclic representations; hence, for any finitely supported function x: G --+ C, we have
Equivalently,
IlL x(t)7l'(t) I ~ IILX(t)Uc(t)lI· ilL x(t)Uc(t) II = sup {IILX(t)7l'(t)II},
(8.1)
where the supremum runs over all possible unitary representations 7l': G --+ B(H) on an arbitrary Hilbert space H. More generally, for any finitely supported function x: G --+ B(Ji) we have
where the sup is the same as before. These formulas explain why C* (G) is sometimes called the "maximal" C*-algebra of G. We denote by
AC: G --+ B(£2(G))
[resp.
Pc: G --+ B(£2(G»]
the left (resp. right) regular representation of G, which means that AC(t) (resp. pc(t) is the unitary operator of left (resp. right) translation by t
8. Group C*-Algebras. Universal Algebras and Unitizatioll
149
(resp. t-I) on £2(G). Explicitly, if we denote by (OdtEC the canonical basis of £2(G), we have Ac(t)Os = Ots (resp. pc(t)os = 0st-1) for all t, s in G. Actually, we will often drop the index G and denote simply by U, A, p the representations Uc, Ac, Pc. We denote by CA(G) (resp. C;(G)) the·C*-algebra generated in B(£2(G)) by Ac (resp. pc). Equivalently, CA(G) = span{A(t) I t E G} and C;(G) = span{p(t) I t E G}. Note that A(t) and p(s) commute for all t, s in G. \Ve will study the maximal tensor product of two C* -algebras AI, A2 in more detail in Chapter 11. For the moment we will content ourselves with its definition. For any x in Al Q9 A 2, we define
Ilxll max = sup 117r(x)IIB(H)l where the supremum runs over all possible Hilbert spaces H and all possible *-homomorphisms 7r: Al Q9 A2 ---> B(H). The completion of Al Q9 A2 with respect to this norm is a C*-algebra called the maximal tensor product of Al and A2 and is denoted Al Q9max A 2. Let G I, G2 be two discrete groups. It is easy to see (exercise) that C*(GJ)
Q9max
C*(G 2) c::= C*(G I
G 2),
(8.3)
CHGJ)
Q9min
C A(G 2 ) c::= CHG 1 x G 2),
(8.4)
X
and similarly for the free product G I * G 2 (the free product of C* -algebras is defined before Theorem 5.13):
These identities can be extended to arbitrary families pair (G I , G 2 ). In particular, we have
*
iEI
C*(Gd c::= C* (
*
iEI
Gi )
(Gi)iEI
in place of the
(8.5)
.
The following very useful result is known as Fell's "absorption principle." Proposition 8.1. For a.ny unitary representation
AC
Q9
7r c::= Ac Q9 I
7r:
G
--->
B(H), we have
(unitaryequimlence).
Here I stands for the trivial representation of G in B(H) (i.e., I(t) = IH ' B(£2), we have
II L II L
a(t)Ac(t)
b(t)
Q9
Ac(t)
Q9
7r(t)11 = II L
a(t)Ac(t)ll,
Q9
7r(t) II = II L
b(t)
Q9
Ac(t)lI·
Introduction to Operator Space Theory
150
Proof. Note that AG ® 71" acts on the Hilbert space K = f2(G) ®2 H :::' f2(G; H). Let V: K -4 K be the unitary operator taking x = (X(t))tEG to (71" ( t- 1 )x( t) )tEG. A simple calculation shows that V- 1 (AG(t) ® IH)V = AG(t) ®
•
7I"(t).
The next result illustrates the usefulness of this principle.
Theorem 8.2. We have an isometric (C*-algebraic) embedding
J: C*(G)
c
C,\(G)
®max C~(G)
taking UG(t) to AG(t) ® AG(t) (t E G). Moreover, there is a (completely) contractive projection P from C~ (G) ®max C~ (G) onto the range of J.
Proof. Let x: G x G following claim:
Il Lt
X,(t,
-4
C be finitely supported. It suffices to show the
t)UG(t)ll.
::; IlL x(s, t)AG(S) ® AG(t)11
C (G)
s,t
(8.6) max
Indeed, in the converse direction we have obviously
Hence (8.6) implies at the same time that J defines an isometric *-homomorphism and that the natural ("diagonal") projection onto its range is contractive. We will now prove this claim. Let 71": G -4 B(H) be a unitary representation of G. \Ve introduce a pair of commuting representations (71"1,71"2) as follows:
7I"1(AG(t)) = AG(t) ® 7I"(t)
and
7I"2(AG(t)) = PG(t) ® I.
Note that both 71"1 and 71"2 extend to continuous representations on C,\(G). For 71"1, this follows from the preceding absorption principle. For 71"2, it follows from the fact that PG:::' AG (indeed, if W: f2(G) -4 f2(G) is the unitary taking 1St to ISt - l , then W*AG(')W = PGO)· Since 71"1 and 71"2 have commuting ranges, we have
8. Group C* -Algebras. Universal Algebras and Unitization
151
hence, restricting the left side to 8e 0 He t2(G) 02 H, we obtain (note that (Ac(s)pc(t)8 e ,8e ) = 1 if s = t and zero otherwise)
II
::; IILX(S' t)AC(S) 0 Ac(t)11
LX(t, t)7r(t)11 t
s.t
8(H)
max
Finally, taking the supremum over 7r, we obtain the announced claim (8.6). The same argument (using operator valued coefficients (x(s, t)) instead of scalar ones) shows that 1IPlicb ::; 1. • Theorem 8.3. ([BF, Jol]) Let G be a discrete grollp and H a Hilbert space. Tlle following properties of a function cp: G ---+ B(H) are equivaJent: (i) Tlle linear mapping defined on span[A(t) It E G] by M",(A(t)) extends to a c.b. map
= A(t) 0 cp(t)
M",:C~(G) ---+
B(£2(G) 02 H) witllIIM",llcb::; 1.
(ii) Tllere is a Hilbert space ii and bounded fUllctions x: G ---+ B(H, ii) and y: G ---+ B(H,ii) witll SUPtEC Ilx(t)1I ::; 1 and SUPsEG Ily(s)11 ::; 1 sllcll tlwt \;f s,t E G cp(sC I ) = y(s)*x(t). Proof. Assume (i). Then, by Theorem 1.6, there are a Hilbert space ii, a representation 7r: C~(G) ---+ B(ii), and operators V;: £2(G) 02 H ---+ ii (i = 1,2) with IIVIIII1V211 ::; 1 such that (8.7) We define x(t) E B(H, ii) and yes) E B(H, ii) by x(t)h = 7r(A(t- I )) VI (8 t 0h) and y(s)k = 7r(A(s-I))V2(8 s 0 k). Note that (A( 0) 0 cp(O))( 8t 0 h), 8s 0 k) = 1 {6=st-l} (
= (y(s)*x(t)h, k),
and we obtain (ii). Conversely, assume (ii). Define 7r: C~(G) ---+ B(£2(G)0 2ii) by 7r(x) = X0Iii' Let V;: £2(G)02H ---+ f.2(G) 02 ii be defined by VI (8 t 0h) = 8t 0 x(t)h and V2(8 s 0 k) = 8s 0 y(s)k. Then for any 0, t, s, h, k we have W2*7r(A(O))VI(8t 0 h), 8s 0 k)
= (A(O)8t, 8s )(y(s)*x(t)h, k) = (A(O) 0
152
Introduction to Operator Space Theory
hence M
Remark 8.4. Consider a complex-valued function cp: G ~ C. Then the preceding result applied with C = B(H) yields: IIM:(G) ~ C>:(G)llcb ~ 1 iff there are Hilbert space valued functions x, Y with SUPt IIx(t) II ~ 1 and sUPslIy(s)1I ~ 1 such that cp(st- 1) = (y(s),x(t)).
V s,t E G
Proposition 8.5. Let G be a discrete group and let f c G be a subgroup. Then the correspondence )..r(t) ~ )..a(t), (t E f) extends to an isometric (C*algebraic) embedding J: C>:(f) ~ C>:(G). Moreover, there is a completely contractive and completely positive projection P from C>: (G) onto the range of this embedding. Proof. Let Q = G/f and let G = UqEQ fgq be the decomposition of G into disjoint right cosets. For convenience, let us denote by 1 the equivalence class of the unit element of G. Since G ~ f x Q, we have an identification
such that )..a(t)
"It E f
= )..r(t) ® I.
This shows of course that J is an isometric embedding. Moreover, we have a natural (linear) isometric embedding V: £2(f) ~ f2(G) (note that the range of V coincides with £2(f) ® 81 in the above identification) such that
VtEf Let u(x)
)..r(t)
= V*xV. Clearly we have V t
u()..a(t))
= )..r(t)
if
tEf
E
= V* )..a(t)V. G
and
u()..a(t))
=0
if
t ¢. f.
Therefore P = Ju is a completely positive and completely contractive projection from C>:(G) onto JC>:(f). • In the case of the full C* -algebras, the analogs of the preceding results are as follows.
8. Group C* -Algebras. Universal Algebras and Unitization
153
Proposition 8.6. Let G be a discrete group. The following properties of a function cp: G ----> B(H) are equivalent: (i) Tlle linear mapping taking Ua(t) to cp(t) extends to a c.b. map Tep: C*(G) ----> B(H) witll IITeplicb ~ 1. (ii) The "multiplier" taking Ua(t) to Ua(t) 0 cp(t) extends to a c.b. map Afep: C*(G) ----> C*(G) 0 m in B(H) with IIMeplicb ~ 1. (iii) There is a unitary representa.tion 71": G ----> B(H7r) and contractions V, It': H ----> H7r SUell tlwt cp(t) = V*71"(t)W.
Vt E G
Proof. We will show (i) ~ (iii) ~ (ii) ~ (i). Assume (i). Then (iii) follows by an immediate application of the factorization theorem for c.b. maps, namely, Theorem 1.6. Assume (iii). By maximality of C*(G), the mapping taking Ua(t) to Ua(t) 071"(t) extends to a representation on C*(G); hence, since Jllep = (I0V*)[Ua(t)071"(t))(10W), we have IIMeplicb ~ 1110V*II·III0WII ~ 1, whence (ii). Assume (ii). Since the mapping Tl taking Ua(t) to 1 (the unit in q is clearly a representation on C* (G) (associated to the trivial representation on G), and since we have obviously Tep = (Tl 0 J)Mep, we obtain IITeplicb
~
•
1, whence (i).
Corollary 8.7. A function cp: G ----> C*(G) iff tllere are a unitary representation 71": G ----> B(H7r) and ~,'f/ in H7r such that cp(t) = (71"(t)~, 'f/}. AJoreover, we lwve
where the infimum runs over all possible if ~ = 'f/, Afep is completely positive.
71", ~,
'f/ for which this llOlds. Finally,
Proof. The first assertion for c.b. maps and the equality IIMeplicb = inf{II~IIII'f/II} follow from the equivalence (ii) {::} (iii) in the preceding statement, using V, W: H7r defined by VI = 'f/, WI =~. Now, if Mep is bounded, then (with the above notation) Tep: C*(G) ---->
154
Introduction to Operator Space Theory
Proposition 8.8. Let G be a discrete group and let reG be a subgroup. Then the correspondence Ur(t) ~ UG(t), (t E r) extends to an isometric (C*algebraic) embedding J ofC*(r) into C*(G). Moreover, there is a completely contractive and completely positive projection P from C* (G) onto the range of this embedding. Proof. We will need the classical fact that, for any unitary representation 11": r ~ B(H), there is a Hilbert space fl with H efland a unitary representation 1i': G ~ B(fl) such that \i t E
r
1i'(t)H CHand 1I"(t) = 1i'(t)IH.
Here is a quick justification for this fact (see [Riel for a more general framework). Let {Sj I j E /} be a set of elements of G chosen so that G = UjEI Sjr is the decomposition of G into left cosets. For convenience, we assume that / contains a distinguished element, which we denote by 1, such that Sl = e (unit element of G). Let 11"0: G ~ B(H) be the trivial extension of 11" defined for all x in G by 1I"0(x) = {1I"(X) if x E ~ o otherw1se. We let fl
= €ajEI Hj , where each Hj is a copy of H. Then, for any x in G, we
define 1i'(x): fl ~ fl as the operator associated to the matrix (1I"0(S;lXSj»ij with entries in B(H). We then leave as an exercise for the reader to check that 1i' has the' announced properties. Following l\'lackey's classical terminology, 1i' is the representation "induced" by 11". Using (8.1) and the preceding extension property, it is immediate that J is an isometric embedding. Now, let 11"1: r ~ B(C) ~ C be the trivial representation (11"1 == 1) and let 1i'1: G ~ B(H) be the induced representation as above, with C C fl. Let E fl be the unit in C viewed as sitting in fl. Then it is easy to check that the function cp defined by
e
\i t E G
is nothing but the indicator function of r. Therefore, by Corollary 8.7, the multiplier Mcp defines a completely contractive and completely positive projection from C*(G) onto JC*(r). • Let F be a free group with generators (gi)iEI. The first part of the next result is based on the classical observation that a unitary representation 11": F ~ B(H) is entirely determined by its values Ui = 1I"(gi) on the generators; and if we let 11" run over all possible unitary representations, then we obtain all possible families (Ui) of unitary operators. The second part is also well known (see, e.g., [Pa5]).
8. Group C*-Algebras. Universal Algebras and Unitization
155
Lemma 8.9. Let F be a free group. Let (Ui)iEI be the family composed of the unitary generators of C*(F) (i.e., these are the unitaries corresponding to the free generators of F in the universal unitary representation of F). Let (Xi)iEI be a finitely supported family in B(H). Consider the linear map T: €oo{l) --+ B(H) defined by T«O:i)iEI) = LiEf O:iXi. Tllen we have
LUi. ® Xi iEI
= IITllcb = sup{11 LUi ®
xill
rnill },
(8.8)
C, (F)0minB(H)
where tIle sup runs over all possible families of unitaries (Ui). Actually, the latter supremum remains the same if we let it run only over all possible families of finite-dimensional unitaries (Ui). lIJoreover, if dim( H) = 00, tllen
IILUi®Xill tEl
=inf{IILYiy;111/21ILZ;Zilll/2},
(8.9)
C*(F)0min B (H)
where the infimum that runs over all possible factorizations Xi = YiZi with Yi, Zi in B(H) is actually attained. Moreover, all tllis remains true if we enlarge the family.(Ui)iEI by adding tIle unit element ofC*(F). Proof. It is easy to check by going back to the definitions that, on one hand
where the sup runs over all possible families of unitaries (ud, and on the other hand that
where the sup runs over all possible families of contractions (td. By the Russo-Dye Theorem, any contraction is a norm limit of convex combinations of unitaries, so (8.8) follows by convexity. Actually, the preceding sup obviously remains unchanged (see Exercise 2.1.1) if we let it run only over all possible families of contractions (t i ) on a finite-dimensional Hilbert space. Thus it remains unchanged when restricted to families of finite-dimensional unit aries (Ui). Now assume IITllcb = 1. By the factorization of c.b. maps we can write T(o:) = V*n(o:)W, where n: €oo{l) --+ B(H) is a representation and \vhere V, Ware in B(H, H) with 1IVIIIIWil = IITllcb. We can assume I finite and, since dim(H) = 00, H = H. Let (ei)iEI be the canonical basis of €oo(I). We set
156
Introduction to Operator Space Theory
It is then easy to check IILYiYilil/21ILZizilil/2 ::; IIVIIIIWII = IITllcb. Thus we obtain one direction of (8.9). The converse follows from the inequality
(already proved in Remark 1.13) applied to bi = Ui 0 Yi and ai Finally, the last assertion follows from the next remark.
=
1 0 Zi. •
Remark. Let {O} be a singleton disjoint from the set I and let J = I U {O}. Then, for any finitely supported family {Xj I j E J} in B(H) (H arbitrary), we have (8.10)
where the supremum runs over all possible families (Uj )jEJ of unitaries. Indeed, since
L Uj 0 x j jEJ
=
III
0 xo +
LU01~i 0Xil , iEI 1110
Xi
the right side of (8.10) is the same as the supremum of Xo + LiEI Ui 0 II over all possible families of unitaries (Ui)iEI, but (recalling U(gi) = Ui ) the latter supremum coincides with
I I0 XO+ LUi0Xill. tEl mm I
•
Remark 8.10. Let F be a free group with free generators (gi). Then, for any finitely supported sequence of scalars (ai)' for any H, and for any family (Ui) of unitary operators in B(H), we have
Indeed, this follows from the absorption principle (Proposition 8.1) applied to the function a defined by a(gi) = ai and = 0 elsewhere, and to the unique unitary representation1l" on F such that 1l"(gi) = Ui· We now return to the unital operator algebra of an operator space E, which we introduced in Chapter 6. Recall that OAu(E) is defined as the completion of the unital tensor algebra Tu(E) for the operator space structure determined by (6.3). Moreover,
8. Group C* -Algebras. Universal Algebras and Unitization
157
OA(E) is the subalgebra of OAu(E) formed by all elements with vanishing constant term.
By construction, the operator algebra OA(E) (resp. OAu(E)) is characterized by the following universal property (recall that "completely contractive" is abbreviated by c.c.): For any c.c. map a: E ---> B(H), there is a unique c.c. morphism &: OA(E) ---> B(H) (resp. unital morphism &: OA,,(E) ---> B(H)) extending a. This characterizes OA(E) (resp. OA,,(E)) lip to a completely isometric isomorphism of algebras (resp. unital algebras) among all (resp. unital) operator algebras containing E completely isometrically. The next result is a reformulation of von Neumann's classical inequality. Theorem 8.11. If E = C, then OA,,(E) can be identified completely iso-
metrically with the disc algebra A(D)
= span{zn In 2:: O}
C C(fJD).
Here D denotes the open unit disc in C. Proof. Assume E = C. Let liS denote by z E E c OAu(E) the element corresponding to the unit of C. Then any element in T" (E) can be written as P = ajz0j (aj E q, which we identify with the polynomial P(z) d . 2::0 ajzl. By (6.3) we have
2::g
where the supremum runs over all contractive T in B(H) (H arbitrary). Let peT) = aoIH + 2::~ a j Tj. We now claim
IIP(T)II :::; sup{IP(w)11 w E fJD}.
(8.11)
This is nothing but von Neumann's classical inequality. Here is a quick proof: As already noted, for any contractive T in B(H) there is a unitary V in lIh(B(H)) c::: B(H EB H) such that T = Vll . Indeed, we simply set
V = ( -(1 _
~*T)I/2
(1-TT*)1/2) T* .
Let j and p denote, respectively, the inclusion and the projection corresponding to the first coordinate in H EB H. Then T = Vll can be rewrit.t.en as T
=p
(1 0) (1 0). 0
0
V
0
0
),
Introduction to Operator Space Theory
158
and more generally (since pj = 1), for any polynomial P, we have trivially
P(T) = pP(U(O))j, where we have set
U(w) =
(~ ~) V (~ ~).
Hence IIP(T)II
= IlpP(U(O))j1l ::;
sup IIP(U(w))II, wED
but since w yields
hence
---+
P(U(w)) is obviously analytic in w, the maximum principle IIP(T)II ::; sup IIP(U(w))II; Iwl=l ::; sup{IIP(U)111 U unitary},
and, by the spectral theorem for unitary (or normal) operators, we obtain our claim: IIP(T)II ::; sup IP(z)l· zE8D
Thus we conclude that 1IPIIOAu(C)
=
IIPIIA(D),
or, equivalently, OAu(lC) ~ A(D) isometrically. Using operator coefficients instead of scalars, it is easy to check that this identification actually is completely isometric. • More generally, we have (cf. [Bol]):
Theorem 8.12. Let I be any set. Let FI be the free group with generators (gi)iEI and let Ui E C*(FI) be the unitary element associated to gi. We denote byei the canonical basis of £1(1). Then the correspondence ei ---+ Ui extends to a unital homomorphism that is a completely isometric embedding ofOAu (max(£l(I))) into C*(FI). Proof. Let E = max(£l(I)). We may as well assume by density that I is finite. Let P E K ® Tu(E). We have P = Ed Pd with Pd E K ® E®d of the form Pd = Ail ... id ® (eil ® ... ® eid)
L
with Ait ... id E K. Then, by (6.3), we have
8. Group C* -Algebras. Universal Algebras and Unitization
159
where the supremum runs over all H and over all families of contractions (Ti)iEI in B(H). By the same t.rick as in the preceding proof, the supremum remains the same if we restrict it to families of unitaries (d. [Bol]). Therefore we obtain, by the observation preceding Lemma 8.9, IIPIIK0".;nOA,,(E) =
IlL .L d
' \ I ... id
® Uil U;2 ...
'·t···l,/
which shows t.hat the correspondence pletely isometric homomorphism.
eit ei2 ... ei,/ --->
ui ,/
I
K0".;n C '(Fil
Uil Ui2
...
'
Ui,J is a com•
Corollary 8.13. Fix d 2': 1. \Vit11 the same notation as before, we lmve a completely isometric isomorphism
Proof. \Ve simply combine the preceding statement with Proposition G.G . • We will now introduce (following [Pes]) the universal C* -algebra of an operator space E, which will be denoted by C*(E). The definition stems from the next statement. Theorem 8.14. Let E be an operator space. There is a C*-algebra A and a completely isometric embedding j: E ---> A wit11 the following properties: (i) For any C*-algebra B and any completely contractive map u: E ---> B there is a representation 11": A ---> B extending u, tl18t is, such that 1I"j = u. (ii) A is the smallest C*-algebra containing j(E).
Moreover, (ii) ensures that the representation
11"
in (i) is unique.
Proof. The proof is immediate. Let I be the "collection" of all u: E with lIullcb : : : 1. Let
We define j: E
--+
--->
BlI
B by
j(x)
= EBu(x). uEI
It is easy to check that j is completely isometric. Then, if we define A to be the C*-algebra generated by j(E) in B, it is very easy to check the announceo • universal property of A.
160
Introduction to Operator Space Theory
Notation. We will denote by C*(E) the C*-algebra A appearing in the preceding statement, and we denote by C~(E) its unitization. Note that C*(E) is essentially unique. Indeed, if jl: E -> Al is another completely isometric embedding into a C* -algebra Al with the property in Theorem 8.14, then the universal property of A (resp. AI) implies the existence of a representation 11": A -> Al (resp. 11"1: Al -> A) such that 1I"j = jl (resp. 1I"dl = j). Since C* -representations are automatically contractive, we have 1111"11 ::; 1, 111I"tII ::; 1, and 11"1 = 11"-1 on Jt(E) and hence on the *-algebra generated by jl(E), which is dense in At, by assumption. This implies that 11" is an isometric isomorphism from A onto AI. • Similarly, C~ (E) is characterized as the unique unital C* -algebra C containing E completely isometrically in such a way that, for any unital C*algebra B (actually we may restrict to B = B(H) with H arbitrary), any c.c. map u: E -> B uniquely extends to a unital representation (Le. *-homomorphism) from C to B. Remark. If two operator spaces E, F are completely isometrically isomorphic, then E and F can be realized as "concrete" operator subspaces E c A and FeB of two C* -algebras A and B for which there is an isometric *-homomorphism 11": A -> B such that 1I"(E) = F. Indeed, let A = C*(E) and B = C*(F), let u: E -> F be a completely isometric isomorphism, let 11": C* (E) -> C* (F) be the ( unique) extension of u (as in Theorem 8.10), and let 0': C*(F) -> C*(E) be the (unique) extension of u- 1 • Then clearly we must have (by unicity again) 0'11" = I and 11"0' = I: hence 0' = 11"-1. • To illustrate the use of this notion, we can state a nice relationship between direct sums of operator spaces and free products. Recall that we have defined the direct sum El 871 E2 of two operator spaces (in the iI-sense) in §2.6 and the free products in Chapter 5 (before Theorem 5.13). Using formula (5.7) together with Theorem 5.1 and comparing with (5.17), the following result is easy to check (we leave the details to the reader). Theorem 8.15. Let Et, E2 be arbitrary operator spaces. canonical (C*-algebraic) isomorphisms
Then we ha've
and
Proof. It suffices to check that the free product C*-algebras C*(E1 )*C*(E2) and C~(El) * C~(E2) have the universal property characteristic, respectively,
8. Group C* -Algebra.''!. Universal Algebras and Unitization
161
of C*(E) and C~(E) when E = El EEll E2 is embedded into the free product in the natural way. • A similar result holds for N-tuples E l , ... , EN (or actually for arbitrary families) of operator spaces. Remark 8.16. It is easy to see that OA(E) (resp. OA,,(E)) can be identified with the closed (resp. unital) subalgebra generated by E in C* (E) (resp. C~(E)). Of course, C*(E) can also be identified with a C*-subalgebra of C~(E).
For any c.c. map 0": E - t B(Ha), we denote by (j : OAu(E) c.c. morphism extending 0" and by
-t
B(Ha) the
the unital representation extending 0". Note that (j can be identified with the restriction of tra to OAu(E) viewed as a subalgebra of C~(E). The following simple fact is sometimes useful. It shows the "residllal finiteness" (see [Pes]) of the universal algebras of E. Proposition 8.17. Let E be any operator space. Let
C = {O": E
-t
B(H) I dim(H) <
00,
1I001Icb ::;
I}.
Then the embedding J: C~ (E)
-t
EElaEcB(Ha)
defined by J(x) = EEl aEC7r a (X) is a completely isometric unital representation. A fortiori, when we restrict to either C*(E) (resp. OAu(E) or OA(E)), we obtain a completely isometric representation (resp. morphism). Proof. The fact that J is isometric is an easy consequence of the following elementary fact. Let Tl' ... ' Tn E B(H). Let P( {Ti , Tn) be a polynomial in Tl, ... ,Tn , Tt, ... ,T~. Then
where the supremum runs over all finite-dimensional subspaces J( c H. This implies that, for any x in the *-algebra generated by E in C~(E), we have
Ilxll ::; IIJ(x)ll, and hence that J is isometric. The completely isometric case is proved similarly using polynomials with operator coefficients instead of scalar ones. •
Introduction to Operator Space Theory
162
Let (Eo, Ed be a compatible couple of operator spaces and let Eo = (Eo, E1)o, with 0 < () < 1. We have c.c. injective maps Eo ---+ Eo + El and El ---+ Eo + El that extend to C.c. morphisms OAu(Eo) ---+ OAu(Eo + Ed and OAu(Ed ---+ OAu(Eo+Et). We claim that these morphisms are injective. Indeed, using Proposition 6.6, this can be reduced to the fact that, if a c. b. map j: E ---+ F is injective, then, for any N ~ 1, j ® ... ® j extends to an injective map from E ®h'" ®h E to F ®h'" ®h F (N times). The latter fact follows from Proposition 5.16. Using this claim, we can view (OAu(Eo),OAu(E 1)) as a compatible couple. The next result shows that the interpolation functor commutes with the functor E ---+ 0 Au (E) .
Theorem 8.18. With the above notation we have completely isometric identities and OA(Eo)
~
(OA(Eo ), OA(E1))o.
Proof. We will prove only the second identity. The proof of the first one is exactly the same. By Propositions 6.6 and 2.7.6 we have a completely isometric embedding v: Eo C (OA(Eo), OA(Et))o
obtained by interpolating between Eo
---+
OA(Eo)
and
El
---+
OA(Ed.
Let OAo = (OA(Eo),OA(Ed)o. By the universal property of OA(Eo), v extends to a c.c. morphism OA(Eo) ---+ ~Ao. To show that is completely isometric, it suffices to show by Proposition 8.17 (applied to OA(Eo)) the following claim: For any n and any c.c. map a: Eo ---+ A1n there is a c.c. OAo ---+ Mn extending a. Indeed, this claim allows us to extend morphism the embedding Eo ---+ OA(Eo) to a c.c. morphism from OAo to OA(Eo), which must be the inverse of v. To prove the preceding claim, we first assume that Eo n El is dense both in Eo and El and we make crucial use of the following identity:
v:
v
a:
liaIiCB(Ee,M,,) = li a li(CB(Eo,Mn ),CB(E lo M n
»9'
which follows from Theorem 5.22 and the discussion of the duality in §2.7 (recall that CB(E, Mn) can be identified with Cn®hE*®hRn or, equivalently, with (Rn ®h E ®h Cn)*)' Finally, the restriction that Eo n El be dense both in Eo and El can be removed a posteriori using Lemma 2.7.2. .
8. Group C*-Algebras. Universal Algebras and Unitization
163
An alternate proof, perhaps more direct (not using duality), can be given using only Propositions 5.22 and 26.14.
-
.
Let E be an operator space. We define its unitization E as the linear span of the unit and E in OA1t(E). It is easy to check that E can be characterized as the unique unital operator space containing E completely isometrically and such that any complete contraction a: E --+ B(H) admits a unique unital completely contractive extension 0:: E --+ B(H). But actually, the next result shows that this notion of unitization is essentially trivial. Proposition 8.19. TIle unitization E of an operator space E can be identified completely isometrically with tIle direct sum C EElI E, with the unit corresponding to (1,0). Proof. Let j: CEElI E --+ E be the mapping taking (A,x) to AU+x, where we have denoted by U the unit in C c OA1t(E). We claim that j is completely isometric. Let eo = (1,0) E C EEl E, let XI,"" Xn be elements of E, let ei = (0, Xi), (1 ~ i ~ n) and let ao, ... , an be in K. We have by (6.3)
Iltai0j(ei)l. t-O
=sup{llao0I+Lai0a(XitniJ,
mill
where the sup is over all complete contractions a: E hand, we have, by definition (see §2.6),
Itai 0 eill. i
t-O
= sup
{ilao 0 T + t
, __ I
IIl1l1
--+
ai 0
B(H). On the other
a(Xi)II},
where the supremum runs over all c.c. maps a: E --+ B(H) and all contractions Tin B(H). By the Russo-Dye Theorem, the sup is the same if we restrict T to be unitary, and then after multiplication by 10 T- I (on the right say) we obtain exactly the same supremum as above. Hence we conclude that
which completes the proof by (2.1.8).
•
Exercises Exercise 8.1. Prove that any unital C*-algebra (resp. separable C*-algebra) C is isomorphic to a quotient of C*(IF) for some free group IF (resp. for IF IFoo), so that C:::: C*(IF)jI for some (closed two-sided) ideal I c C*(IF).
=
164
Introduction to Operator Space Theory
Exercise 8.2. Consider an operator space El and a closed subspace E2 eEl. Let j: E2 ---; El be the inclusion map. Then j extends to completely isometric embeddings C*(E2) ---; C*(El) and OA(E2) ---; OA(Ed (and similarly in the unital case). Exercise 8.3. Let E 1 , E2 be as above. Let q: El ---; Ed E2 be the quotient map. Then the unique representation 11": C*(E1) ---; C*(EdE2) and the unique morphism u: OA(E 1 ) ---; OA(EdE2) associated to q are complete metric surjections (and similarly in the unital case). (Thus the functors E ---; OA(E) and E ---; C* (E) are "projective"; they are also "injective" by the preceding exercise.) Exercise 8.4. A discrete group G is called amenable if it admits an invariant mean, that is, a functional cp in £00 (G)+ with cp( 1) = 1 such that cp( St * f) = cp(f) for any f in £oo(G) and any tin G. It is known (and the reader should use this as an alternate definition) that G is amenable iff there is a net (fo) in the unit sphere of £2(G) such that II.\(t)fo - foll2 -,...; 0 for any tin G. Show that the following are equivalent:
(i) G is amenable. (ii) C* (G)
= C A(G).
(iii) There is a generating subset S subset E c S, we have
c
lEI =
G with e E S such thet, for any finite·
II~ .\(t)ll·
Chapter 9 Examples and Comments In Banach space theory, ever since Banach and l\Iazur's early work, several examples have played a privileged role, such as fp, Co, L p , and C(K) (with K compact). These are usually called the classical Banach spaces. In the light of more recent developments, it is tempting to extend the list to the Orlicz, Sobolev, and Hardy spaces as well as the disc algebra and the Schatten pclasses, although in some sense all these examples are derived from those of the first generation. Analogously, one could make a list of all the classical C*algebras or von Neumann algebras (see e.g. [Da2]). In the present chapter, our aim is to describe the spaces that, in our opinion, are the best candidates to appear on a list of the classical operator spaces (we have already met some examples, sllch as R,C, min(£2), max(£2), and OH).
9.1. A concrete quotient: Hankel matrices \Ve start with a natural example of a quotient operator space, namely, the space
(here Loo is equipped with its natural o.s.s., i.e., the minimal one), which can be identified with the subspace of B(£2) formed of all the Hankel matrices, that is, all the matrices (aij) in B (£2) such that aij depends only on i + j. To explain this identification, we need some specific notation and background from classical harmonic analysis. \Ve denote by Lp the space Lp on the torus 11' equipped with the normalized Haar measure m. \Ve denote by H p the su bspace of Lp formed of all functions cp with Fourier transform CP vanishing on the negative integers and by H~ c Hp the subspace of those cp such that cp(O) = o. l\Iore generally, given a Banach space X, we denote by Lp(X), Hp(X) and H~(X) the analogous spaces of X-valued functions on 11' (1 ::; p < 00). To any cp in L oo , one classically associates the Hankel matrix (aij) with entries aij = cp( -i - j) \Ii, j ~ O.
It is easy to see (identifying £2 to H 2 ) that this matrix defines a bounded linear operator u
166
Introduction to Operator Space Theory
is an isometry from L:xAMn)/H!(Aln) into Mn(B(£2)) = B(£2(£2))' Hence, in the language of operator spaces, we can reformulate the "vectorial" Nehari Theorem as follows: Theorem 9.1.1. The correspondence 1jJ embedding of L(XJ/H! into B(£2)'
->
1-l[1jJ] is a completely isometric
This is a special case of Theorem 9.1.2, which is proved below. Thus, Hankel matrices provide us with a very natural and concrete "realization" of the quotient Loo/ H! as an operator space (the existence of which is guaranteed a priori by Ruan's Theorem). It would be interesting to have an analogous description in the general case for the quotient E1 / E2 of two operator spaces with E2 c E 1. Even in the particular case when E1 = L(XJ and E2 is a translation invariant subspace, this does not seem to be known (see, however, Exercise 8.2). We will prove in the following a generalization to Hankel matrices with entries in a von Neumann algebra. For that purpose, we need more notation. Let M c B(H) be a von Neumann algebra on a Hilbert space H. For simplicity, we assume H separable. We say that a function rp: 11' -> AI is a-measurable if for any x, y in H the function (rp(·)x,y) is measurable. We will denote by Loo(A1) the tensor product (von Neumann sense) L(XJ(1I')0M. Equivalently, L(XJ(AI) is formed of all equivalence classes of bounded a-measurable functions rp: 11' -> AI equipped with the norm Ilrpll(XJ = ess sup IIrp(t)IIJI,f· tEl'
Note that, for such a function rp, the Fourier transform ip: Z defined by the formula:
v x,y E H
(ip(n)x, y) =
->
Al can be
J
(rp(t)x, y)e-intdm(t).
We denote by H!CM) the subspace formed of all rp in L(XJ(M) such that ip(n)=OVn~O.
For any rp in L(XJ (M) let
be the operator of multiplication by the (operator-valued) function rp (taking x E L 2 (H) to the function t -> rp(t)x(t)). Clearly
(9.1.1) We will work with the subspaces H 2 (H)
Hg(H).
c
L2(H) and Hg(H).l = L 2(H)
e
9. Examples and Comments
167
The Hankel operator H[
The associated matrix is given by V i,j
~
0
(H[ifJ);j
=
( - (i + j)).
Note that IIH[ifJII ~ II if II "". Here again H[
Theorem 9.1.2. The correspondence
Proposition 9.1.3. Consider if in Loo(M).
(i) Let Tep: £2(Z; H)
~
£2(Z; H) be tl1e operator defined by the matrix
(Tep)ij
= (i - j),
(i,j E Z).
Then IITepil = II
Tben IITepil = IIifiloo. (Note that Tep may be viewed as the restriction of Tep to £2(Z+; H).) (iii) Consider a function b: Z ~ M. Let T(b): £2(Z+; H) ~ £2(Z; H) be the operator defined by the matrix T(b)ij = b(i - j), (i E Z, j E Z+). Then T(b) is bounded iff there is a (necessarily unique)
II
{If
(
Introduction to Operator Space Theory
168
x
where the supremum runs over all x, y in the unit ball of L 2 (H) such that and fj are finitely supported. In particular, we may assume x = E;'-N x(j)e ijt with x(j) E H. We then have
(Al",x, y) =
L
($(i - j)x(j), fj(i)) =
L
($(i - j)x(j - N), fj(i - N));
i,jEZ
hence which establishes (ii). (iii) Assume first that there is cP in Loo(1\J) with $ = b. Then Ilr(b)11 = IIr",lI, and hence in this case (iii) follows from (ii). In general, for any 0 < r < 1, let br : Z -> M be defined by br(n) = r1n1b(n). Assume that r(b) is bounded. Then clearly sUPn IIb(n)1I < 00, which implies that the series CPr = EnEZ br(n)e int is absolutely convergent. Hence by the first part we have IICPriloo = IIr(br)1I ~ IIr(b)lI. Let cP be a weak-* cluster point • in Loo (M) of {CPr} where r -> 1. It is easy to see that $( n) = b( n). Our main result will follow from the following beautiful lemma proved by several authors, including Parrott [ParI].
Lemma 9.1.4. Consider a decomposition K
K. We view an operator on K as a matrix
= Kl EB K2 of a Hilbert space
(~ ~)
with operator entries.
Then
and this infimum is attained. Moreover, if Kl = K2 = H and if b, c, d are all in a von Neumann algebra M C B(H), thef!. this infimum is attained 01] an element a of M. Proof. It suffices to show that, if the right side is that
II (~ ~) II ~
1. So assume the right side is
~
~
1, then there is a so 1. Equivalently, we
have both IIb*b + d*dll ~ 1 and llce* + dd*1I ~ 1 (see Remark 1.13), which implies b*b ~ 1 - d*d and cc* ~ 1 - dd*. The latter ensure that we can find operators V, W with I!VII ~ 1, IIWII ~ 1 such that b = V(1 - d*d)1/2 and c = (1- dd*)1/2W. We then set a = -Vd*W. Let U
(1 - dd*)1/2) d = ( -(1 _ d*d)1/2 d* .
It is easy to check that U is unitary and, moreover,
9. Examples and Comments hence we conclude
I (~ ~) I : :
169
1. If b, c, d all belong to AI, it is easy to see
that, in the above argument, we can find V, W in l1f so that a = - V d* lV also lies in M. • Theorem 9.1.5. Let (a(n)k:::o be a sequence in M. Let r = (a(i + j))i,j'20 be the associated Hankel matrix. Then IIfIlB(l'2(H» ::::: 1 iff tllere is cP in Loo(M) with IIcplloo ::::: 1 such that \fn~O
a(n)
= ij5( -n).
Proof. If a( n) = ij5( -n) for all n ~ 0, then r is the matrix associated to 1t[cp]; hence Ilfll ::::: Ilcplloo and the "if" part is obvious. Conversely, assume Ilfll ::::: 1. We claim that we can find operators a(-I), a( -2), a( -3), ... in M such that the function b: Z ---+ AI defined by b(n) = a( -n) satisfies IIr(b)11 ::::: 1. We can first find a( -1) in M so that the matrix a( -1) ( a(O) a(l)
a(O) a(l) ...
a(l) ... ...
... ) . .. . ..
has norm ::::: 1. This is possible: Indeed, this matrix can be written as ( a( ~1)
with
II(~ ~)II
=
II(~ ~)II
= Ilfll;
~)
he~cetheexistenceofa(-l) follows
from the preceding lemma, since we assume lin ::::: 1. Repeating the same argument, we can find a( -2) in l1f so that a(-2) ( a(-l) a(O)
a(-l) a(O) ...
a(O) ... ...
... ) . .. . ..
has norm < 1. Continuing exactly in the same way, we find a( -3), a( -4), and so on. This gives us an extension of a to the whole of Z such that, if we set b(n) = a( -n), we have Ilr(b)1I ::::: 1. Then. by Proposition 9.1.3(iii), there is cp in Loo(M) with IIcplloo ::::: 1 such that b = 'P. Hence a(n) = 'P( -n) for all n~O.
•
Proof of Theorem 9.1.2. By Theorem 9.1.5, for any 1/J in Loo(M) we have
111i[1/J]1I = inf{llcplioo 11t[cp] = 1i[1/J]} = inf{lI1/J + hll oo I hE H!(M)}.
170
Introduction to Operator Space Theory
This shows that 1/J ---+ 1i[1/J] defines an isometric embedding of Loo(A1)/ H!(M) into B(H2(H), H8(H).L). Replacing M by Mn(M) (n ;::: 1), we obtain that this embedding is actually completely isometric. • Corollary 9.1.6. Let Aft C B(HI ) and M2 C B(H2) be two von Neumann algebras. Let A = (Aij) and B = (Bij) be Hankel matrices wit11 entries, respectively, in MI and M 2. Assume that there is a bounded linear map v: MI ---+ M2 with IIvll ::; 1 such that V(Aij) = Bij for all i,j. Then
Proof. By the preceding result, for any bounded Hankel matrix (Aij) with entries in MI there is cp in Loo(Md with 11'1'1100 = II (Ai}) II such that Aij = $( -i - j). We claim that
II(v($( -i - j)))11 ::; 11($( -i - j))II· The following, slightly incorrect, argument is easier to understand on a first reading: Since Ilv: MI ---+ M211 ::; 1, we have IIJ®v: Loo(Md ---+ L oo (M2)1I ::; 1; hence, passing to the quotient v defines a contraction from Loo(Jtld/ H!(Md to L oo (M2 )/H!(M2 ), and the above claim now follows from the preceding result. The incorrect point in the preceding argument lies in the fact that, since cp(t) is only a-measurable, the a-measurability oft ---+ v(cp(t)) is questionable when v is not weak-* continuous, so that J ® v is not a well-defined map from LooCMd to LooCM2)' However, this is easy to repair: Fix 0 < r < 1. Then CPr(t) = LnEZ rlnl$(n)e int is strongly measurable, so the preceding objections do not apply and we have t
---+
Letting r
i 1, we obtain the announced result lI(v(Aij))11 ::; II(Aij)lI.
•
Remark. Actually Lemma 9.1.4 itself reflects a "concrete" realization of a quotient operator space, namely, fl.f2/ S, where S C Af2 is the one-dimensional subspace spanned by ell. Indeed, Lemma 9.1.4 implies that the mapping
Af2
---+
. (a b) (0 b) tB (0 0)
Af2 tBM2 takmg
c
d
to
0
d
c
d
.
defines (after passmg
to the quotient modulo its kernel) a completely isometric embedding of AI2 /S into M2 tB M 2 • By a rearrangement of the basis, this also shows that the mapping
9. Examples and Comments
171
defines a completely isometric embedding of Ah/Cel2 into 1Il2 EB A12. In that
= (~ ~) to the set of strictly upper triangular matrices (Le., such that a = c = d = 0). case, infb
II (~ ~) II
appears as the distance of the matrix x
More generally, for any n 2 1, let Tn C AJn be the subspace of upper triangular matrices (a;j), that is, such that a;j = 0 for all j < i. Let Pk be the orthogonal projection onto span(el, ... ,ek). Then, for any matrix x = (aij), we have inf
Ilx - y/l
yET"
=
sup /1(1 - Pk)xPkll·
l~k
l\Joreover, the mapping
x
--+
E9
(1 - Pk )xPk
l~k
defines a completely isometric embedding of lIln lTn into the (n - I)-fold direct sum Mn EB ... EB Un. This can be proved by repeated applications of Lemma 9.1.3. There is an analogous result for more general "nest algebras" (for which Tn is a prototypical case), called "Arveson's distance formula." See either [DaI] or [Pow] for more on all this. Remark 9.1.7. Let us denote simply by C the subspace formed in Loo by the continuous functions, and let A = Hoo n C be the subspace of C formed by the boundary values of analytic functions. By the vectorial version of Hartman's Theorem (cf. [NiD, the restriction to the continuous functions of the correspondence appearing in Theorem 9.1.1 is a complete isometry between the quotient operator space CIA and the space of all compact Hankel operators on £2' In particular, this implies that the operator space CIA is exact in the sense of §I7 below. On the other hand (this remark is due to Junge), the results of [HPI, Theorem 2.2] suggest that the space Lool Hoo might be exact, but this does not seem clear at the moment. Until recently, we knew of no example of a quotient of a commutative C*algebra (by a closed subspace) that is not exact as an operator space, but an example of this kind has just been exhibited by N. Ozawa [OzI]. An operator space that is (completely isometrically) a quotient of a minimal operator space is called a Q-space; equivalently, t.hese are subspaces of quotients of commutative C* -algebras. The class of Q-spaces seems to be quite interesting to study. See [Ri2] for an investigation of an analog of the Haagerup tensor product for Q-spaces.
172
Introduction to Operator Space Theory
9.2. Homogeneous operator spaces
We say that an operator space is homogeneous (resp . .A-homogeneous) if, for all u: E ~ E, we have Ilulicb
= lIuli (resp. Ilulicb ::;: .Allull)·
In this case, every surjective isometry on E is a complete isometry. For example, it is easy to check that, for any Banach space E, the spaces min(E) and max(E) (defined in Chapter 3) are homogeneous. Actually, the most interesting case seems to be the Hilbertian one: We will say that E is Hilbertian if it is isometric (as a normed space) to a Hilbert space. When E is only .A-isomorphic to a Hilbert space for some .A > 1, we will say that E is .A-Hilbertian. The spaces Rand C (introduced in (0.1) and (0.2» as well as min(t'2), max(t'2), and OH are examples of homogeneous Hilbertian operator spaces. The homogeneous Hilbertian operator spaces reproduce the same operator space structure inside themselves; when a space E is a mixture of two distinct structures, for instance, E = REEl C, then it is not .A-homogeneous for any .A. More precisely, we have: Proposition 9.2.1. Let E c B(H) be a Hilbertian operator space. Then E is homogeneous iff, for any unitary operator U: E ~ E (here "unitary" refers to the Hilbert space structure of E), we have IlUlicb ::;: 1. In tlIat case, whenever two subspaces E1 c E and E2 C E and llave the same Hilbertian dimension, they are completely isometric. Proof. The first assertion follows from the Russo-Dye Theorem: The unit ball of B(E) is the closed convex hull of its unitaries ([Ped, p. 4]). Then the second assertion follows: Indeed, let u: E1 -~ E2 be an isometric isomorphism. After comparing with the orthogonal projection onto E 1 , we can extend u to u: E ~ E such that Ilull = 1 and, if E is homogeneous, we have Ilulicb = lIull, whence Ilulicb ::;: 1. Similarly (reversing the roles) lIu- 11lcb ::;: 1. Thus we conclude that u is completely isometric. • Proposition 9.2.2. Every .A-homogeneous operator space is completely isomorphic to a homogeneous (i.e., l-homogeneous) one E with dcb(E, E) ::;: .A. Proof. Let (3 be the family of all the operators u: E ~ E with Ilull ::;: 1. Let Bu = B(H) for all u in (3. Consider the direct sum (f7uE~Bu (equipped with the operator space structure defined in §2.6) and ~et J : E ~ EEluE~Bu be the isometric map defined by J(x) = (U(X»UE~· Let E be the range of J. Then J is an isometry from E onto E. Moreover, IlJllcb :::; .A and 11J;~1I1cb ::;: 1. Finally, by its very definition, the space
E is homogeneous (i.e., I-homogeneous).
•
9. Examples and Comments
173
Remark. The consideration of a few simple examples like R, C, or OH gives the impression that a homogeneous Hilbertian operator space might be detenllined by its two-dimensional subspaces. But it is not so: Using [r\'IP], C. Zhang [Z2J exhibited for each integer N a pair of homogeneous Hilbertian (infinite-dimensional) operator spaces that have the same (Le. completely isometric) N-dimensional subspaces bllt which are not even completely isomorphic. The examples that we will review in this chapter demonstrate the rich diversity of the homogeneous Hilbertian operator spaces appearing in the literature. Nevertheless, it seems reasonable to expect that a elassification or a parametrization of this class of operator spaces will be available soon.
9.3. Fermions. Antisymmetric Fock space. Spin systems Let I be any set. Consider a family of operators in B(H) satisfying the following relations, called the canonical anticommutation relations (in short, CAR): (CAR) {
+ Vj*Vi
'Vi,j E I
ViVj*
'Vi,jE1
ViVj+VjVi=O.
=
6;)1
Here, of course, we have set 6 jj = 1 if i = j and 6;j = 0 if i i- j. The closed span in B(H) of the family [Vi liE 1J gives us a very interesting example of an operator space. We will denote it by (I). We set = (N) and n = <1>( {I, 2, .. , n}). Let us first justify the existence of such families. Let 1i = £2(I) equipped with an orthonormal basis (ei)iEI, let Ho = te, and let Hn = 1i An (antisymmetric Hilbertian tensor product) for n ~ 1. \Ve define the antisymmetric Fock space associated to 1i as follows:
For each h in 1i, it is classical to denote by c(h) (resp. a(h» the so-called operator of "creation (resp. annihilation) of particle" defined by 'Vx E H
c(h)x = h /\ x
(resp. a(h) = c(h)*).
Then we have the relations 'Vh, k E 1i
a(h)a(k)* +a(k)*a(h)
=< h, k > I
et
a(h)a(k) +a(k)a(h)
= o.
If we let Vi = a(ei) (or if we set Vi = c(e;) for all i), we obtain a family with the announced properties. We refer the reader to [Gu2, Gu3, EvL, BRJ for more information.
The next result is classical.
174
Introduction to Operator Space Theory
Theorem 9.3.1. The operator space (I) is (isometrically) Hilbertian and homogeneous. Up to a complete isometry, it only depends on the cardinality of I (and not on the particular family (Vi)iEf chosen to define it). Proof. Let (Oi)iEf be a finitely supported family of scalars and let T = LiEf 0i Vi· By the CAR we have
Then Hence whence (9.3.1) Therefore, the space (1) = span[Vi liE I] is isometric to £2(1). Let H' be another Hilbert space and let {Wi liE I} be a system in B(H') satisfying the CAR. Let A (resp. B) be the C* -algebra with unit generated by {Vi liE I} (resp,{Wi liE I}). By a classical argument (see [BR, p. 15J), one shows that there is a C* -algebraic isomorphism 7r: A -+ B such that 7r(Vi) = Wi for all i. A fortiori (since, by Proposition 1.5, 7r is completely isometric), the spaces span[Vi] and span[Wi ] are completely isometric. This justifies the last assertion in Theorem 9.3.1. Finally, let us show that (I) = span[Vi] ~ £2(I) is homogeneous: According to Proposition 9.2.1, it suffices to show that every unitary U: (I) -+ (I) is completely contractive. But, as shown by a simple computation, if U is unitary, then the operators Wi = U(V;) still satisfy the CAR; hence the mapping U extends to an isometric representation from A onto B. A fortiori, U is a complete isometry and 1lUlicb = 1. Remark. We should mention that some variants of the space (I) appear in the theory of Clifford algebras. On the other hand, a space very similar to (I) also appears in the study of certain domains of holomorphy, called Cartan factors of type IV (cf. [HaJ). In the latter theory, the term "spin system" is used to designate a family of Hermitian and unitary operators {Ui liE I} in B(H) such that 'Vi
i= j
UiUj + UjUi = O.
For instance, the classical Pauli matrices
form a spin system.
9. Examples and Comments
175
There is (essentially) a one-to-one correspondence between spin systems and systems satisfying the CAR. Indeed, if (Uj)jEI and (Uj')jEI are two systems such that their disjoint union forms a spin system, then the system (ltj)jEI, defined by ltj = (Uj + iUj')/2, satisfies the CAR. In the converse direction, if {Vi 1 i E 1} is a family satisfying the CAR, then t.he operators Ui = Vi + ~* form a spin system. The subspace of B(H) generated by a spin system (Ui);EI clearly is isomorphic to £2(1). To be more explicit, we must distinguish the real case from the complex one. We have
V(o:;) E 1R(I)
V(O:i)
E C(1)
(2: 10:;12)1/2 =
(2: IO:iI 2)1/2 ~
II 2: O:iU;.II,
II 2: O:iUili
~ 21/2(2:
IO:iI 2)1/2.
More precisely (see [Ha]), we have
Remarks. It is known that, if 1 = {l, ... , n}, t.he unital C* -algebra generated by <1>(1) = <1>n is isomorphic to the algebra 11[2" of 2n x 2n matrices (one can obtain an entirely explicit realization of <1>n by using tensor products of the Pauli matrices). This remark shows that <1>(1) is an "exact" operator space with constant = 1 in the sense of §l7. It is possible to show that, if 1 is infinite, for any completely isometric embedding <1>(1) c B(H), there is no c.b. projection from B(H) onto <1>(1) (but there is a bounded one). Except for these few facts, there is very little available informat.ion on the operator space structure of <1>(1). It would be interesting to compute its dual <1>(1)* as well as the iterated tensor products <1>(1) ®min ... ®min <1>(1) (n times) in the style of [HP2] (note, however, that <1>(1) can be easily distinguished from all the examples reviewed in this chapter; see Chapter 10). The field of fermionic analysis is a very active one at the moment. See, for example, [M] for the probabilistic viewpoint and also [BoSl, BoS2, CL, BCL, CK, PX]. It is only natural to expect that there will be further points of contacts with operator space theory.
9.4. The Cuntz algebra On Let S1, ... , Sn E B(H) be isometries such that
The C*-algebra generated by St ... , Sn is called the Cuntz algebm (cf. [Cu]) and is denoted by On. One can show that it does not depend on the particular
176
Introduction to Operator Space Theory
choice of the sequence {Si} but only on n. It is tempting to study the operator space En = span[sl' .. , sn], but this is disappointing: A very simple calculation shows that for all al, ... ,an in B(H), we have II Lai 0 sill = II Lai'ailll/2. It follows that En is completely isometric to Cn. Similarly, span[si, .. , s~] is completely isometric to Rn. Of course, similar remarks hold for the C*algebra 0 00 generated by a sequence of isometries (Si) such that L~ SiSi' ::; I (this time the equality is not required for the unicity of 0 00 ), Another very important family of isometries is provided by the creation operators on the full Fock space. To define this we need to introduce some notation. Let H be a Hilbert space. We denote by :F(H) (or simply by :F) the full Fock space associated to H, that is to say, we set 'Ho = C, 'Hn = H~!m (Hilbertian tensor product) and finally
:F = EBn~o 'Hn· We consider from now on 'Hn as a subspace of :F. For every h E :F, we denote by C(h): :F ~:F the operator defined by:
C(h)x=h0x. More precisely, if x = >'1 E 'Ho = Cl, we have C(h)x = >'h and if x = Xl 0 X2 ... 0 Xn E 'H n , we have C(h)x = h 0 Xl 0 X2 ... 0 x n . We will denote by n the unit element in 'Ho = Cl. Now assume H = C~ with its canonical basis (ei). We denote :Fn = :F(C~) and we set (i=l, ... ,n). Then {C i } is a family of isometries on :Fn with orthogonal ranges, and we have n
LCiC: = I - Po, I
where Po is the orthogonal projection onto cn. Now let {xd be any n-tuple in B(H) (H arbitrary). We will denote by £ {Xi} the closed linear span of I and all the products of the form XiI Xi2 ... Xip XiI' .. Xiq (p 2:: 0, q 2:: 0). In particular, taking q = 0, we see that
span[I,xi] C £{Xi}' The space £{Xi} can also be viewed as the span of all products P{Xi}Q{Xi}, where P{Xi } and Q{Xi } are polynomials in noncommutative variables {Xi}' Note that our polynomials are allowed to have a "constant term" equal to a multiple of the unit in the free unital semi-group generated by {X~, ... , X n }. In a series of papers ([Pul-3]) G. Popescu stud;ed various extensions of von Neumann's inequality for n-tuples of operators; in particular, he proved the following.
9. Examples and Comments
177
Theorem 9.4.1. ((Pu3i) Assume IIL~ x~x;11 S 1. For any finite family Pk(Xl, ... ,Xn ), Qk(Xl, ... ,xn ) of polynomials in noncommuting variables {Xl, ... , Xn} we have
More generally, tile linear mapping
defined by .
is completely contractive. This shows that {£ 1, ... , £n} is in some sense "maximal" among all the n-tuples {xd with IILxixil1 S 1.
Remark. In [Pu3], Popescu also observed that for "analyt.ic" polynomials, the Cuntz isometries {sd can be substitut.ed t.o {£d in t.he preceding st.at.ement.. Namely, for any polynomial P we have
Moreover, the correspondence £i -> Si extends to a complet.ely isometric unit.al homomorphism on the (nonself-adjoint) algebra generated by {£i}'
Corollary 9.4.2. We llave completely isometrically
and similarly with {Si} substituted to {£i}. Proof. Let us denote by eo the unit in C. \Ve will show that the correspondence u: C EIh C n ~ span[l, el , ... , en] defined by u(eo) = land u(eil) = (i is completely isometric. Note that, by Remark 1.13, the restriction of u t.o en is completely contractive. Hence, by the definition of 611, we have lIullcb S 1. To show that Ilu-llicb S 1, it suffices to establish the following claim: For any family of matrices {ao, al, ... ,an} in !lIN we have (9.4.1 )
Introduction to Operator Space Theory
178
where the norm is the norm in MN(IC EB1 Cn) and MN(B(Fn)). By Theorem 8.19, the left side of (9.4.1) is equal to the supremum of
where v: C n ---- B(H) runs over all possible maps with IIvll cb ::; 1. But, if we let Xi = v(eil), we clearly have (see Remark 1.13) IIExixill1/2 ::; Ilvllcb ::; 1. Hence we find
Ilao ® eo + L and since
1IFlicb =
ai ® ei111 = sup lIao ® I
+ La; ® Xi I '
1, we obtain
Thus we conclude lIu -111 cb ::; 1 and the last assertion concerning {Si} follows • from the preceding remark. Remark. It would be interesting to study more generally the relation between . an operator space and the C* -algebra that it generates. For some information in this direction, see [KiV, Zl, KiW, Pll].
9.5. The operator space structure of the classical Lp-spaces In this section, we describe the operator space structure of the usual (Le. commutative) Lp-spaces. We will see that these spaces can be equipped with a specific structure, which we call their natural structure. We refer to [P2] for more details. Let (0, A, JL) be a measure space. We will denote by Lp(JL) the associated Lp-space of complex-valued functions. Given a complex Banach space X, we will denote by Lp(JL; X) the Lp-space of X-valued measurable functions in Bochner's sense. It is well known that
with () = lip. A priori, this formula is an isometric identity that is only valid in the category of Banach spaces, but the complex interpolation of operator spaces will allow us to extend this formula to the operator space setting. The case p = 00. First, if A is a C*-algebra, it possesses a privileged operator space structure associated to any realization of A as a C*-subalgebra of B(H). Indeed, by Proposition 1.5, iftwo C*-algebras are isomorphic (as C*-algebras), then they are completely isometric, and hence they are identical as operator spaces. We will say that this particular structure is the natural operator space structure on a C* -algebra.
9. Examples and Comments
179
In particular, if p = 00, we have a natural operator space structure on the space Loo(Jl) or on the space C(T) of all continuous functions on a compact set T. It is easy to check that this natural structure is determined by the isometric identity C(T) Q9min K = C(T; K), where the space on the right-hand side is the space of K-valued continuous functions equipped with its usual norm. This space coincides with the injectiye tensor product (in Grothendieck's sense) of C(T) and K. In other words, the natural operator space structure on E = £'XJ(/L) or E = C(T) makes it identical to the space min(E), as defined in Chapter 3. The case p = 1. If p = 1, again the choice is clear: The natural structure is defined as the Olle induced on Ll (/L) by the dual space Loo(/L)* equipped with its dual operator space structure. Explicitly, this means that the norm of lIIn (L 1 (Jl)) is by definition the norm induced by CB(Loo(Jl),lIfn ). (Note that lIfn (L I (Jl)) can be identified with the a(Loo(p), Ll (It ))-continuous linear maps from Loo(Jl) to lIfn .) For the resulting operator space structure Oll L 1 (Jl), we have L 1 (Jl)* = Loo(/L) completely isometrically. Proposition 9.5.1. The operator space obtained by equipping Ll (/1) with its natural o.s.s. (as defined above) coincides with max(L l (p)). Proof. Indeed, let B = L 1(/1) and let E = max(B). By (2..5.1) (or Corollary 5.11), the inclusion E ---> E** is a complete isometry, and, by Exercise .3.2, we have E** = (max(B))** = ma..x(B**) = (min(B*))*. Therefore the o.s.s. induced by max(B**) on B coincides with that ofmax(B). In tIle present case, we have max(B**) = (min(B*))* = Loo(/l)*, so we conclude that tIle opera,tor space structure of max(L l (Jl)) is tIle same as tlJat induced by Loo (Jl)*, where L oo (/l) is equipped with its natural (i.e. minimal) operator space structllre. • In particular, in the case n = N, we have a natural o.s.s. on fl. It is not hard to verify that this natural o.s.s. on (1 also coincides wit.h the one obtained by considering £1 as the dual of c{). vVe can describe the associated norm II Ilmin on K Q9 fl in the following manner: Let. (en) be the canonical basis of C1. For any finite sequence (an) in K (or in B (£2)) we have (9.5.1) where the infimum runs over all possible decomposit.ions an = bncn in K (resp. B(f 2 ))· We will give two more descriptions of the same structure, in (9.5.3) for p = 1 and in (9.6.1). Analogously, we can describe the natural o.s.s. of Ll (f.l) as follows. Let f E K ®min L 1 (Jl). We may consider f as a K-valued function on n. We
Introduction to Operator Space Theory
180 then have
IlfllK:®mlnLl(/L) = inf
{III
I III
g(t}g(t}*dt.t(t} 1/2
I
h(t}*h(t}dt.t(t} 1/2} ,
where the infimum runs over all possible decompositions of f (measurable) K-valued functions.
(9.5.2) as a product of
The formula (9.5.1) (resp. (9.5.2}) is the quantum version of the fact that the unit ball of £1 (resp. L 1 (t.t)} coincides with the set of all products of two elements in the unit ball of £2 (resp. L 2 (t.t)). These two formulas (9.5.1) and (9.5.2) can be deduced from the fundamental factorization Theorem 1.6.
The case 1 < p < 00. For the general case, we use interpolation. Consider the operator space structure on Lp(t.t} corresponding to (L oo (t.t},L 1 (t.t}}o as defined in §2.7. By definition, the norm on Mn(Lp(t.t)} is the one induced by the space (Mn (L oo (t.t}},Mn (L 1 (t.t}}o with () = lip. Once again, we will say that this is the natural operator space structure on Lp(t.t). We can describe more explicitly this structure using the Schatten classes Sp already introduced (see (7.19}). Indeed, let f E K @ Lp(t.t) (viewed as a K:-valued measurable function). Then we have (9.5.3) where the supremum runs over all a, b in the unit ball of S2p. Since Lp(t.t} has been equipped with an o.s.s., we may now unambiguously discuss completely bounded maps u: Lp(t.t} ---4 Lp(t.t}. These turn out to be easy to describe. Proposition 9.5.2. A linear map u: Lp(t.t}
---4 Lp(t.t} is completely bounded (on Lp(t.t) equipped with its natural o.s.s.} iff the mapping u@Is" is bounded on Lp(t.t; Sp}. Moreover, we have
It has been known for a long time that the Hilbert transform is bounded on the (vector-valued) Lp-space of Sp-valued functions for any 1 < p < 00 (see also in [Bou1] the analogous result for martingale transforms). This implies that, if 1 < p < 00, the Hilbert transform is completely bounded on Lp (on the torus or on JR.). The same is valid for the Riesz transforms and their Gaussian analogs on the n-dimensional torus or on JR.n; see, for example, the discussion in [P15]. It is then tempting to compare boundedness and complete boundedness for maps on Lp. By Proposition 9.5.2, it is easy to check that, in the cases· p = 1,2 or 00, every bounded map on Lp is automatically completely bounded.
9. Examples and Comments
181
However, if 1 < p 1= 2 < 00, there are examples of Fourier multipliers on Lp of the torus that are not c.h. Such examples (implicit in [P4, pp. 110--113]) can be obtained easily by observing that the norm in 8 p is invariant under all transformations (aij) ----> (E;aijE'j) for any choice of signs E; = ±1, E'j = ±1, but that there exist, if 1 < p 1= 2 < 00, mappings that are unbounded on 8 p and of the form (aij) ----> (E;ja;j) with Eij = ±1 for i :::; j and (say) Eij = 0 otherwise. Fix (E;j) such that the latter mapping is unbounded on 8p- Then let u: Lp ----> Lp be the Fourier multiplier defined as follows: If n is of the form n = 3 k + 31 with k :::; l, then u{e;nt) = Ekleint and u{e int ) = 0 otherwise. By a classical variant of the Khintchine inequalities (cf., e.g., [LR, p. 65]), it is known that u is bounded on L p , but by the choice of (cij) (and by the invariance property ofthe 8 p-norm recalled above) u is unbounded on Lp{8p), and hence it is not c.b. Similarly, it follows from [HP1] that there are Fourier multipliers bounded on Hi (of the torus) but not c.b. The preceding definitions can be extended in the case of noncommutative Lp-spaces. Consider, for instance, the Schatten class 8 p • It is well known that, if () = lip, we have an isometric identity (9.5.4) Here, it would be more appropriate to denote by 8= the space K! The space K can be equipped with its natural o.s.s. as a C* -algebra and 8 1 with the dual o.s.s. associated to the identity 8 1 = K*, relative to the duality defined for x E 8 1 , Y E K by (x, y) = tr(t xy). Then, the formulas (2.7.3) and (9.5.4) allow us to define an o.s. structure (again called natural) on 8p- By Theorem 7.10, the space 8 2 is then completely isometric to OH(fir x fir) or, equivalently, to OH (see Remark 7.13 and the discussion preceding it). As we already mentioned, given an operator space E c B(H), exactly the same idea leads to the definition of a natural o.s.s. on the E-valued Lp-space that remains valid in the noncommutative case. When p = 1, we define 8 1 [E] as the "projective operator space tensor product" of 8 1 with E, introduced in §2.8 and denoted by 8 1 ®" E. By [ER8], we have (9.5.5) 8dE*] ~ (K ®min E)* completely isometrically. When 1 < p < 8 p [E] = (K
00
®min
we define
E, 8t[E])
°,
(9.5.6)
where () = lip· Then the following result holds (cf. [P2, Lemma 1.7])
Proposition 9.5.3. Fix p with 1 :::; p < 00. Let u: E ---+ F be a linear map between two operator spaces. Then u is c.b. (resp. completely isometric) iff
182
Introduction to Operator Spa,ce Theory
the map Isp ®u defines a bounded (resp. an isometric) linear map from Sp[E] to Sp[F]. Moreover, we have
Ilulicb = IIIsp ® UIlS,,[Ej--+S,,[Fj'
:0
:1 '
Furthermore (see [P2, Corollary 1.4), if 1 :::; Po, P1 :::; 00 and = 1;;:,0 + then for any compatible couple of operator spaces (Eo, Ed we have completely isometrically (9.5.7) Spo[(Eo,E1 )0] = (Spo[EO],Spl[E 1])o. We refer the reader to [P2] for more details.
9.6. The C"-algebra of the free group with n generators Let Fn (resp. Foo) be the free group with n (resp. countably many) generators, and let {gll g2, ". }) be the generators. Let 11": F 00 -4 B (H) be a unitary representation of the free group. We will see that, in several instances, the operator space E(1I") spanned in B(H) by {1I"(gi) Ii = 1,2,,,.} has interesting properties. We will first illustrate this with the universal representation U: Foo -4 B(1t), which generates the full C" -algebra C" (F00), as introduced in Chapter 8. The other classical choice for 11" is the left regular representation A , which generates the "reduced" C"-algebra C~(Foo). This case is discussed in the next section, §9.7. We let E[j = span[U(gi) I i = 1,2, '" n] and Eu
= span[U(gi) Ii:::: 1].
By Lemma 8.9, for any finite sequence (ai) in B(£2)' we have (9.6.1)
where the supremum runs over all sequences (Ui) of unitary operators in B (H) and over all possible Hilbert spaces H. Actually, the supremum remains unchanged if we restrict ourselves to H = £2 or to H finite-dimensional with arbitrary dimension. But then formula (2.11.2) shows that, if we denote by (en the dual basis to the canonical basis of Co (equipped with its natural 0.S.8.), we also have
II L ai ® U(gi)lImin = II Eai ® e;IIB(i2)®mlnCo' Therefore, the mapping u: Co - 4 Eu that takes ei to U(gi) is a complete isometry. Hence, we have proved:
9. Examples and Comments
183
Theorem 9.6.1. The operator space Eu spanned by the generators in C* (Feo) is completely isometric to fl equipped with its natural operator space structure (or, equivalently, its o.s.s. as the dual of co). Similarly, Efj is completely isometric to
fr.
The formula (9.6.1) can be viewed as the "quantum" analog ofthe classical formula
\I(..\i) E
C(I)
11(..\i)II£1 = L I..\il = sup{1 L..\i zill
Zi
E C,
Iz;1 =
I}.
(9.6.2) The space Eu gives us a concrete realization of the space fl as an operator space. More generally, for any measure space (0, It), one can describe t.he natural operator space structure of Ll(O,lt) (induced by Leo(O"t)*) as follows. For all f in L 1 (0, It) ® B(f2)' we have
IIfIILd!1'JL)0I11iIlB(~2) = sup
{II!
f(w) ® g(w)dJL(w)
I
B(f2)0I11iIl B (f2)
},
(9.6.3)
where the supremum runs over all functions 9 in the unit ball of the space of Leo-functions with values in B(f2)'
9.7. Reduced C*-algebra of the free group with n generators
Let G be a discrete group. We have defined in Chapter 8 the left regular representation..\: G -+ B(f2(G)). Recall that ..\(t) is the unitary operator of left translation by t on f2(G). We denote by C~(G) the C*-algebra generated in B(f2(G)) by {..\(t) I t E G} or, equivalently, CHG) = span{..\(t) I t E G}. Clearly, we have a C* -algebra morphism Q: C*(G)
-+ C~(G)
that takes U (t) to ..\( t). By elementary properties of C* -algebras, it is onto and we have C~(G) ~ C*(G)j ker(Q). In general, ker(Q) =I- {O}, but one can show that C~(G) = C*(G) (Le., ker(Q) = {O}) iff G is amenable. The free groups are typical examples of nonamenable groups. The fact that the algebras C~ (G) and C* (G) are distinct in this case is manifestly visible on the generators. Indeed, if we let
E>: =
span[..\(gi)
Ii
= 1, .. , n]
184
Introduction to Operator Space Theory
we can see that E>. is a very different space from its analog in the full C*algebra, namely, the space Eu studied in §9.6. Indeed, as Banach spaces, we have Eu ~ €l and E>. ~ €2. The first isomorphism is elementary (see Theorem 9.6.1), while the second one is due to Leinert [Le]. Using Haagerup's ideas from [H2], one can describe the operator space structure of E>. as follows (see [HP2] for more details). Consider the space B(€2) EB B(€2)' equipped with the norm II(x EB y)1I = max{llxll,llyll}. In the subspace REB C c B(€2) EB B(€2)' we consider the vectors 8i (i = 1,2, ... ) defined by setting
We will denote by RnC the closed subspace spanned in REBC by the sequence
{8d· This notation is compatible with the notion of "intersection" defined in §2.7, provided we view the pair (R, C) as a compatible pair using the transposition mapping x - 4 tx as a way to embed R into C. This means that we let X = C, we use x -4 tx to inject R into X, and we use the identity map of C to inject C into X. Similarly, we will denote by Rn n C n the subspace of R n C spanned by {8 i I i = 1,2, ... , n}. It is easy to verify that, for any Hilbert space H and for any finite sequence (ai) in B(H), we have
Then, we can state (see [HP2]):
Theorem 9.7.1. The space E>. is completely isomorphic to R n C. l\fore precisely, for any finite sequence (ai) in B(H), we have (9.7.1) so that tIle mapping u: R n C -4 E>. defined by u(8i ) = .>.(gd is a complete isomorphism satisfying lIulicb = 2 and Ilu-11l cb = 1. Moreover, the map P: C~(Foo) -4 E>., defined by P.>.(t) = .>.(t) ift is a generator and P.>.(t) = 0 otherwise, is a c.b. projection from C~(Foo) onto E>. with norm IlPlicb : : ; 2. (Similar results hold for E'): and Rn n C n .)
ct
Proof. Let G = Foo. Let c G (resp. Ci- C G) be the subset formed by all the reduced words that start by gi (resp. gil). Note: Except for the empty word e, every element of G can be written as a reduced word in the generators admitting a well-defined "first" and "last" letter (where we read from left to right). Let p/ (resp. Pi-) be the orthogonal projection on €2(G) .
9. Examples and Comments with range span[8t check that
It
E
ct]
(resp. span[(8 t
E
Ci-D. Then it is easy to
+ )..(g;)(1 - Pi-) + p/ )..(g;)(1 - P;-), + p/ )..(gi),
)..(gi) = )..(gi)Pi= )..(gi)Pi-
= )..(gi)Pjso that )..(gi)
It
185
= Xi + y; with (note xixi = Pi- and
y;y;
= p/)
Therefore, for any finite sequence (ai) in B(H) we have by (1.11) (note a; 0 X; = (ai 0 1)(1 0 x;) and similarly for ai 0 y;) Ill: ai 0 )..(gi) II
~
Ill: ai 0 Xiii
~ IIl:a aill i
i 2 /
+ Ill: ai 0
Yill
+ IIl:a: ai ll i / 2
~ 211l:a; 08ill· The converse follows from a more general inequality valid for any discrete group G: For any finitely supported function a: G -+ B(H) we have max {11l:a(t)*a(t)ll
i 2 / ,
IIl:a(t)a(t)*ll
i 2 / }
~ IIl:a(t) 0
)..(t)llmill· (9.7.2)
To check this, let T = 2::a(t) 0 )..(t). For any h in BH we have T(h 0 8e ) 2:: a(t)h 0 8t . Hence IIT(h 0 8e )11 = (2::t Ila(t)hI1 2 ) 1/2, and hence Ill: a(t)*a(t)ll i / 2
= sup (l: Ila(t)hI1 2) 1/2 hEB/I
=
~ IITII.
Similarly, since T* = 2::a(t- 1 )* 0)..(t), we find
IIl:a(t)a(t)*ll
i 2 /
~ IIT*II = IITII,
and we obtain (9.7.2). If we now return to the case G = IF",,, we find that (9.7.2) implies the left side of (9.7.1). Moreover, if P is as in Theorem 9.7.1, we have
(I0 P) (l: a(t) 0 )"(t») = l : a(gn) 0 )..(gn). Hence, if T
= 2:: a(t) 0 )..(t), we have
11(10 P)(T) I
~ 2 max {liE a(gn)*a(gn)11 1/ 2 , liE a(gn)a(gn}* 111/2} ::; 2 max
{liE a(t)*a(t) 111/2 , liE a(t)a(t)* 111/2} ,
186
Introduction to Operator Space Theory
and by (9.7.2) this is:::; 211TII. Thus we conclude, as announced, that 1IPlicb :::;
2.
•
We will now describe the operator space generated by the free unitary generators {.\(gi) I i = 1,2, ... ,} in the noncommutative Lp-space (1 :::; p < 00) of the free group Foo. Let us denote by r the standard trace on CHFoo) defined by r()..(f» = f(e). Let Lp(r) denote for 1 :::; p < 00 the associated noncommutative Lp-space. The space Ll (r) is the predual of the von Neumann algebra generated by )..(Foo), which we will denote by Loo(r). When 1 < p < 00 and () = lip, the space Lp(r) can be identified with the complex interpolation space (C! (F00), Ll (r»o or, equivalently, with (Loo (r), Ll (r»o . Using §2.7, we may view Lp(r) as an operator space. Let us denote by Ep the closed subspace of Lp( r) generated by the free generators {.\(gi) I i = 1,2, ... } (note that Eoo = EA)' We may view Ep as an operator space with the o.s.s. induced by Lp(r). Clearly, the orthogonal projection P from L2(r) to E2 is completely contractive (since, by Theorem 7.10, L2(r) can be identified with OH(I) for a suitable set I). On the other hand, by Theorem 9.7.1, that same projection P is completely bounded from C!(Foo) onto EA' Actually, the proof of Theorem 9.7.1 shows that P extends to a weak-* continuous projection from Loo(r) onto EA' By transposition, P also defines a c.b. projection from Ll(r) onto E 1 . Therefore, by interpolation, P defines a completely bounded projection from Lp(r) onto Ep for any 1 < p < 00. Moreover, by Proposition 2.7.6, the existence of this simultaneous c.b. projection ensures that the space Ep can be identified completely isomorphic ally with (EA' Edo with () = lip. In addition, El ~ (EA)*' By Theorem 9.7.1, we have EA ~ R n C and by duality El ~ R + C; hence we have (completely isomorphically) Ep ~ (RnC, R+C)o. We will compute the latter space more explicitly in Theorem 9.8.7, but let us state what we just proved. Corollary 9.7.2. Let Lp(r) denote the noncommutative Lp-space of the free group and let Ep be the closed subspace generated by tIle free generators {.\(gi) I i = 1,2, ... }. Then we have, completely isomorphically (with () =
lip), Ep
~
(RnC,R+C)o,
(9.7.3)
where, as before, we use the transposition mapping as the continuous injection from R to C, which allows us to view (R, C) as a compatible couple. Moreover, the orthogonal projection from L2(r) onto E2 defines a c.b. projection from Lp(r) onto Ep for all 1 :::; p :::; 00. Moreover, the equivalence constants in (9.7.3) as well as liP: Lp(r) ---+ Epllcb remain bounded when p runs aver the whole range 1 :::; p:::; 00. We will see later (Theorem 10.4) that the space RnC is very different from the Fermionic space ~ described in §9.3. Nevertheless, the following simple statement holds.
9. Examples and Comments
187
Proposition 9.7.3. Let (V;) be a sequence satisfying tlle CAR.. For any finite sequence (ai) in B(H), we llave
II Lai QSH5i li min ::; hll Lai 0 V;llmin, so that the mapping v: q,
--->
RnC defined by v(V;) = 8i has norm IIvllcb ::;
/2.
Proof. It suffices to show the same result for q,n' \Ve may assume that VI, ... , Vn are in a von Neumann algebra equipped with a normalized trace r (for example, ]1.[2" equipped with its usual normalized trace). Then let T =
L ai 0 Vi. The CAR and the trace property imply r(\tj* V;) = r(V; \tj*) = O. Hence, if we set T = Lai 0 V;, we obtain II Laiaill = 211(I 0 r)(T*T) II < 2I1TII~in' and similarly II Laiaill ::; 21IT11~in' We conclude that
•
(9.7.4)
\Ve will now generalize Theorem 9.7.1: Instead of the generators {9d we consider all the reduced words of length d, d being a fixed integer. Recall that (after all possible cancellations have been made) any element t of a free group can be viewed as a reduced word in the generators and their inverses. We denote by It I the length (i.e., the number of letters) of this reduced word. For instance, for the empty word, we have lei = 0, and also 19i1 = Igi 1 1= 1, 19i9; 11 = 2 Vi #- j, and so on. Let G be a free group, freely generated by {9j lie f}. Let Wd = {t E G Iitl = d}. We will study the operator space spanned inside C~ (G) by {,X( t) I t E Wd}. \\'hen d > 1, the analog of RnC is a bit more difficult to describe, because there are now (d + 1) ways (and not only two) to write an element t in H'd as a reduced product L = (3-y with 0 ::; 1(31 ::; d and 1(31 + l-yl = d. Indeed, we can take for (3 the word formed by the first j letters of t and for -y the one formed of its last d - j letters. Actually, we will look at the d + 1 ways to decompose t as t = (3-y-l instead of t = (3-y, but this does not really matter here. This idea leads to a decomposition that we now describe precisely. Let H be any Hilbert space. Let a = {a(t) I t E G} be a finitely supported family in B (H). For any fixed 0 ::; j ::; d, we consider the matrix {a ~~] I (3 E Wj, -y E Wd-j} defined by [d,j]
a{3,'"'(
= a ((3-y -1)1 {1{3,",(-II=d}'
Note that here 1(3-y- 1 1 = d expresses that there is no cancellation in the product (3-y-l. We denote by Iiall [d,j] the norm of this matrix as an operator acting from l2(Wd _ j ; H) to l2(Wj ; H). The following extension of Theorem 9.7.1 is due to Buchholz [Buel] and Haagerup (unpublished).
Introduction to Operator Space Theory
188
Theorem 9.7.4. Let G be a free group and let d 2: 1. With the above notation we have (9.7.5) !v[oreover, the projection Pd: C~(G) ---+ span[A(t) I t E WdJ defined by PdA(t) = A(t)l{tE W d} is c.b. and satisfies IIPdl1 :s: d + 1 and IIPdli cb :s: 2d.
Proof. Let T =
2:
a(t) 0 A(t). We claim that there is a decomposition
tEWd
T = To
+ Tl + ... + Td
with each Tj of the form
"~
Tj =
[d,jj
a{3,"'(
0
u{3v"'(,
1{3I=j,l"Yl=d-j
where
u{3, v"'(
E B(f2(G))
are such that
As we will see, this implies
and then the triangle inequality yields the right side of (9.7.5). We now proceed to check this claim. Note that if we denote by es,t(s, t E G) the matrix units in B(f 2 (G)), we have T
=
L
a(be- 1 )1{lbc- 1 1=d} 0 eb,c·
b,cEG
,X
Note that Ibe- 11= d iff we can write (uniquely) b = (3x (reduced) and e= (reduced) with 1(3,-11 = 1(31 + 1,1 = d. Note that if 1(31 = j, after reduction the word be- 1 consists of the first j letters of b followed by the last (d - j) letters of e- 1 • Moreover, the equalities l(3xl = 1(31 + Ixl and h'xl = III + Ixl are a convenient way to express that the words (3x and 'YX are reduced. A simple calculation then shows that, if 1(31 = j, we have Ibl = j + Ixl and lei = d - j + Ixl; hence j = (Ibl- lei + d)/2. Thus the set of all pairs (b, e) in G x G such that Ibe- 11= d can be decomposed as a disjoint union Co U C 1 U··· U Cd, where Cj
= {(b, e) Ilbe-il = d, Ibl - lei + d = 2j}.
9. Examples and Comments Let Tj
Lg T
Tj =
j
189
L(b,c)ECj a(bc- 1) 0 eb,c' By the preceding discussion we have T
=
=
and moreover
L
aUh,-l)l{lth,-ll=d} 0
liJl=j,hl=d-j
Tj
L
e!3x,")'x1{liJxl=I!3I+lxl}1 {I")'xl=hHlxl}'
xEG
=
a(,B')'-1)1{iJ")'-11=d} 0UiJ V ")"
L 1!3I=j,hl=d-j
Moreover, a simple calculation shows L
UiJU~
II3I=j
= L L eiJx,iJx1{liJxl=liJl+lxl} = 1!31=j x
Hence IlL UiJU~11 ::; 1 and similarly of the above claim.
L
eb,b·
b:lbl~j
IlL v~v")'l1 ::; 1. This completes the proof
Now, an elementary verification shows that, if 9 is any fixed element in G (for instance 9 = e),
and hence
IITjll = IITj 0eg,gll::; La~~)010eiJ,")' = Ilall[d.j). 13,")'
Then, by the triangle inequality, we have d
IITII ::; L
lIall[d,j) ::; (d
+ 1) max{lIall[d,j) 10::; j
::; d},
o
which establishes the right side of (9.7.5). To establish the left side, we will first study the projection Pd. Let a: G ~ B(H) be a finitely supported function. Let Y =
L a(t) 0 A(t) tEG
and
T
= (I 0
Pd)(Y)
=
L
a(t) 0 A(t).
Introduction to Operator Space Theory
190
Then, with the same notation as before (recallllTjll :::; Ilalhd,j]), we will show that for any 0 < j < d we have Ilall[d,j] :::; 211Y11 and that, if either j = 0 or j = d, the factor 2 can be removed. Let H t = H ® 8t . We have an orthogonal decomposition H ®2 f2(G) = EeHt, tEG
relative to which Y is represented by a matrix (Y(s, t)) with coefficients in B (H) defined by Y(s,t) = a(sC 1). Now fix j with 0 :::; j :::; d. If we restrict the above matrix to act from EB1tl=d-j Ht to EB1tl=j H t , we find
(9.7.6) 1,BI=j,bl=d-j Now assume first that a is supported on Wd so that Y = T and the sum in (9.7.6) can be restricted to 1.8),-11 = d. Then, the last estimate yields Ilall[d,j] :::; IIYII = IITII, whence the left side of (9.7.5). \Ve will now estimate the norm of Pd' We no longer assume that the coefficients (a(t)) are supported on W d , but we assume that they are scalar. Then it is easy to check that lIall[d,j] :::;
(L la[d,j](t3),-1W)1/2 :::; (L la(t)12)1/2 :::; ,B,,,(
tEG
IlL
a(t)A(t)11
= IIYII·
tEG
Thus, we obtain IITII :::; L:~ IITjl1 :::; L:~ Ilall[d,j] :::; (d + 1)11Y11 and hence IIPdll :::; d + 1. We will now estimate the c.b. norm of Pd' We return to the general case of a finitely supported function a: G -+ B(H). Note that if either j = 0 or j = d, (9.7.6) implies lIall[d,j] :::; IIYII. Now assume 0 < j < d. For any tin G of positive length we denote by get) the last letter of t (equal to a generator or the inverse of one), and we define x(t) = 8g (t), so that (assuming again lsi> 0) (x(s), x(t)) = 0 iff the product st- 1 is reduced (in othe~ words, iff Ist- 11 = lsi + Itl) and otherwise (x(s),x(t)) = 1. By Exercise 1.5 and the triangle inequality, we deduce from (9.7.6)
L li3I=j,bl=d-j
a(t3),-1)(t3),-1)®e,B,,,(1-(x(t3),xb))) :::;2I1YII,
9. Examples and Comments
191
or, equivalently, (9.7.7) This gives us
and hence
IIPdlicb
~
•
2d.
Remark 9.7.5. Fix an integer D 2: 1. Let a: G -+ B(H) be a finitely supported function with support in lVo U ... U WD . Let ad: G -+ B(H) be the function equal to a on l,vd and equal to zero elsewhere, so that
L
aCt)
Q9
A(t) =
Itl:SD
L
ad(t)
Q9
A(t).
O:Sd:SD
We then have
2- 1 max max {IIadll[d 'J} < O
-
_1-
.J
-
"a(t)
~ Itl:SD
~ 2D(D
Q9
A(t)
+ 1) O:Sd:SD max ma..x {IIadll[djJ}' O:Sj:Sd •
(9.7.8)
Indeed, the left side is an immediate consequence of (9.7.7) while the right side follows from the triangle inequality and (9.7.5).
9.8. Operator space generated in the usual Lp-space by Gaussian random variables or by the Rademacher functions Let (n, A, P) be a probability space. We will say that a real-valued Gaussian random variable (in short r.v.) is standard if E"( = 0 and E"(2 = 1. We will say that a complex-valued Gaussian r.v. 1 is Gaussian standard if we can write 1 = 2- 1/ 2 h' + h") with ,,(', "(" real-valued, independent, standard Gaussian r.v.'s. Let hn In = 1,2,oo.} (resp. 1n In = 1,2,oo.}) be a sequence of real(resp. complex-) valued independent standard Gaussian r.v.'s on (n, A, P). As is well known, for any finitely supported sequence of real (resp. complex) scalars (ank.!he r.v. S = L: ani (resp. L: a;1j) has the same distribution as the variable S = (L: laiI 2 )1/2"(1 (resp. (L: laiI 2 )1/2 1t). In particular, we have for any finitely supported sequence of complex scalars (9.8.1)
Introduction to Operator Space Theory
192
Let Qp be the subspace of Lp (O, A, P) generated by FYn I n = 1,2, ... }. Then, as a Banach space, Qp is isometric to £2 for all 1 ~ p < 00. (To simplify, we will discuss mostly the complex case in the sequel, although the real case is entirely similar provided we restrict ourselves to lR-linear transformations.) Moreover, for any isometry U : Qp ~ Qp the sequence {U{1i) I i = 1,2, .. } has the same distribution as the sequence {1i I i = 1,2, .. }. If we equip Qp with the o.s.s. induced by Lp{O, A, P), it follows (see Proposition 9.5.2) that U is a complete isometry from Qp to Qpo Therefore, by Proposition 9.2.1, we have Proposition 9.8.1. For any 1 Hilbertian operator space.
~
p <
00,
the space Qp is a homogeneous
Let {en I n = 1,2, ... } be a sequence of independent, identically distributed (in short LLd.) r.v.'s on (O, A, P) with ±1 values and such that P{en = +1} = P{en = -I} = 1/2. The reader who so wishes can replace {en I n = 1,2, ... } by the classical Rademacher functions (rn) on the Lebesgue interval; this does' not make any difference in the sequel. Let Rp be the subspace generated in Lp{O, A, P) by the sequence (en). The analog of (9.8.1) for the variables (en) (or, equivalently, for the Rademacher functions) is given by the classical Khintchine inequalities (cf., e.g., [LTl, p. 66] or [DJT p. 10]), which say that, for 1 ~ p < 00, there are positive constants Ap and Bp such that, for any scalar sequence of coefficients (an) in £2, we have (9.8.2) (Note that we have trivially Bp = 1 if p ~ 2 and Ap = 1 if p :::: 2.) This implies that, as a Banach space, Rp is isomorphic to £2 for any 1 ~ P < 00. A fortiori, Rp and Qp are isomorphic Banach spaces if 1 ~ p < 00. We now wish to describe the operator space structure induced by Lp on Qp (resp. Rp). By (9.5.3), this can be reduced to the knowledge of the norm
Il L 'YnXnll L,,(fl,P;Sp) II
(resp. LenxnIiL,,(fl,p;S,,» when (xn) is an arbitrary finite sequence of elements of Sp. In other words, to describe the o.s.s. of Qp (resp. Rp) up to complete isomorphism, it suffices to produce two-sided inequalities analogous to (9.8.1) and (9.8.2) but with coefficients in Sp instead of scalar ones. The noncommutative versions of Khintchine's inequalities proved in [LuP] and [LPP] are exactly what is needed here. The case 1 < p < 00 is a remarkable result due to F. Lust-Piquard ([LuP]). The case p = 1 comes from the later paper [LPP], which also contains an alternate proof of the other cases.
193
9. Examples and Comments Theorem 9.8.2. (i) Assume 2 ~ p < 00. Tllen tllere is a constant tllat, for any tinite sequence (xn) in Sp, we llave
B~
sucll
max{II(l:x~xnf/21Is,., 11(l:xnx~)1/2tJ
(9.8.3)
~ IIl:cnXnll L,.(O,P:s,.)
~B~ max {I (l: x~xn ) 1/211 s,.' I (l: XnX~) 1/211 sJ . (ii) Assume 1 ~ p ~ 2. Tllen tllere is a positive constant for any tinite sequence (xn) in Sp, we llave
A~
stlcll tl1at,
(9.8.4) wllere we llave set
AJoreover, similar inequalities are valid witll a real or complex Gaussian U.d. sequence h'n) or (1n) in tIle place of (cn). Finally, tIle same inequalities are valid wIlen Sp is replaced by allY nOllcommutative Lp space associated to a semi-tinite faitIlfulnormal trace 011 a von Neumann algebra.
Remark. (Observed independently by l\Jarius Junge). We claim that there is a numerical constant C such that, for all 2 ~ p < 00,
Since this is only implicitly contained in [LPP] and it might be of independent interest, we will give the details explicitly. Let us denote by PI: L2 -+ R2 the orthogonal projection. Recall that the K-convexity constant of a Banach space X is defined as follows:
By a standard averaging technique, one easily verifies that
When X K(Lp(Sp»
=
Sp, since fp(Sp) embeds isometrically into Sp, we have
In [LPP] it is proved that the constants A~ are uniformly bounded when 1 :::; p :::; 2. By duality, it follows from the preceding = K(Sp).
194
Introduction to Operator Space Theory
estimate of 'IPI ®JXIILp(X)-+Lp(X) for X = Sp that there is a numerical constant C' such that, for all 2 ~ p < 00, B~ ~ C' K(Lp(Sp)) = C' K(Sp). By [MaP, Remark 2.1OJ, the latter constant is dominated by the type 2 constant, and by [T J2J the type 2 constant of Sp is equal to the best constant in the classical (scalar) Khintchine inequalities Bp (at least when p is an even integer) that is of order pl/2 when p -+ 00. Thus we obtain our claim that there is a numerical constant C such that, for all 2 ~ p < 00,
Remark. By Proposition 9.5.2, we can deduce from the known general results on Gaussian and Rademacher series in Banach spaces (cf. [MaP, Corollaire 1.3]) that the spaces 9p and 'Rp are completely isomorphic for any 1 ~ p < 00. Of course, in the case p = 2, 92 and 'R2 are completely isometric to 0 H since the space L 2 (n, A, P) itself is completely isometric to OH(J), where the cardinal J is its Hilbertian dimension. Let us now identify 9p and 'Rp as operator spaces. We start with the case p = 1, which is particularly interesting. Consider the direct sum R EElI C (as defined in §2.6) and its subspace ~ c R EElI C defined by ~
=
({x, _tx) , x E R}.
In accordance with the definitions in §2.7, we will denote by R+C the quotient operator space (R EElI C)!~. Since R EElI C is equipped with a natural o.s.s. (see §2.6), the space R + C itself is thus equipped with a natural o.s.s. as a quotient space (see §2.4). It is easy to see (cf. §2.7) that (R n C)* = R + C
completely isometrically.
(9.8.5)
In particular, R + C is isomorphic to £2 as a Banach space. We will denote by (O'i) the natural basis of R + C that is biorthogonal to the basis (8i ) of ~l. ~ R n C. Equivalently, if we denote by q: R EElI C -+ R + C the canonical surjection, then we have O'n = q(el n EEl enl). Similarly, we will denote by Rn + C n the quotient operator space (Rn EEll Cn)! ~n' with ~n = ({x, _tx) , x ERn}. We also can identify Rn + C n with the subspace spanned in R + C by (0'1. ... , O'n). Moreover, we have (Rn n C n )* = Rn + Cn completely isometrically. Now we can reformulate the main result of [LPP) in the language of operator spaces as follows:
Theorem 9.8.3. The space 'R 1 (resp. 91) is completely isomorphic to R+C . via the isomorphism that takes Ci (resp. ::Yi) to O'i·
"'·"It
Proof. It is easy to verify that the norm appearing in Theorem 9.8.2 is the dual norm to the natural norm of the space K®min (RnC). Equivalently,
9. Examples and Comments
195
in the notation of §9.5 (using (9.5.5» it coincides with the norm in the space SI [(R n C)*] = SI [R + C]. Therefore, the linear mapping that takes Ci to O'j defines an isomorphism from SdRd onto SdR + C]. Thus, the announced result for Rl follows from Proposition 9.5.3 applied with p = 1. The case of 91 is analogous. • By combining (9.8.5) with Theorem 9.7.1 we obtain a surprising connection between the standard Gaussian (or ±l-valued) independent random variables and the generators of C~(Foo):
Corollary 9.8.4. We lmve (E)..)* ~ 91 ~ Rl completely isomorphically. More precisely, let us denote by ("\*(9j)) the system in (E>.)* that is biortllOgona1 to (..\(9i». Then the mapping u : Rl ---> (E>.)* (resp. u: 91 ---> (E>.)*) defined by U(Ci) = "\*(9i) (resp. U(1i) = "\*(9j») is a complete isomorpbism. IVlore generally, for all 1 ::; p ::; 00, let us denote by R[p] (resp. C[p]) the operator space generated by the sequence {e1j I j = 1,2, ... } (resp. {ei1 I i = 1,2, ... }) in the operator space Sp equipped with its natural o.s.s. defined by interpolation (see (9.5.4». Here again we set Soo = JC. Note that R[oo] (resp. C[oo]) obviously coincides with the row (resp. column) space R (resp. C). A moment of thought (recall Exercise 2.3.5) shows that R[l] ~ R* and C[l] ~ C* completely isometrically; therefore we may identify R[l] with C on one hand, and C[l] with R on the other. IVloreover, we obviously have a natural projection simultaneously completely contractive from Sl to R[l] and Soo to R[oo] (and similarly for columns). This implies, by Proposition 2.7.6, that if we make the couple (R, C) into a compatible one (as we did in Chapter 7) by using transposition to inject R into C, then we have completely isometric identifications (with () = l/p)
R[p]
=
(R, C)o
C[P]
= (C, R)o.
Now, if we use (9.5.7), we find
(9.8.6) and Sp[C[P]] = (Soo[C], SdR])o. If we view Soo[R] (resp. SdC]) as a space of sequences of elements of Soo (resp. S1), then the norm in Soo [R] (resp. S1[C]) is
I
1/21Is=
I
1/211sJ
easily seen to be (Xj) ---> (L XjXi) (resp. (Xj) ---> (L XjX;) Therefore, since we have simultaneously contractive projections (see the discussion before Proposition 2.7.6) onto the corresponding subspaces of Soo [Soo] and SdStl, we find that the norm in (Soo[R], SIlC])o coincides with (Xj) ---> p • In other words, for any sequence (Xj) in Sp we have, by (9.8.6),
II(Lxjxj)1/21Is
(9.8.7)
Introduction to Operator Space Theory
196 Similarly, we have
IILXi 0 eilIISp[C[pJ] = II (LX;Xi) 1/211s"
(9.8.8)
We denote by R[p]nC[p] the subspace of R[P]EBC[p] formed of all couples of the form (x,t x). On the other hand, we denote by R[P] + C[p] the operator space that is the quotient of R[P] 671 C[P] modulo the subspace formed of all couples of the form (x, _tx). Then, by (9.8.7) and (9.8.8), the norm appearing on the left in (9.8.3) is equivalent to the natural norm ofthe space Sp[C[P]] n Sp[R[p]] or, equivalently, Sp[R[P] n C[P]]. Similarly, the norm III in Theorem 9.8.2 (case p ~ 2) is equivalent to the natural norm of the space Sp[C[P]] + Sp[R[P]] or, equivalently, Sp[R[P] + C[P]]. This allows us to state:
",P
Theorem 9.S.5. Let 1 < p < 00. The space 9p (or the space 'Rp)) is completely isomorphic to R[P] + C[P] if p ~ 2 and to R[P] n C[P] if p 2: 2. Proof. First observe that the natural norm in the space Sp['Rp] is equal to the norm induced by Lp(O, A, P; Sp), by the isometric case in proposition 9.5.2. Then, by (9.8.3) and the preceding discussion, the latter norm is equivalent to the natural norm of either the space Sp[R[p] n C[P]] if p 2: 2 or the space Sp[R[P] + C[P]] if p ~ 2, whence the announced complete isomorphisms by • Proposition 9.5.3. Remark. Note that Theorem 9.8.3 is nothing but the natural extension of Theorem 9.8.5 to the case p = 1. Remark 9.S.6. In the Banach space setting, it is well known that the orthogonal projection P2: L2 --4 92 (resp. Q2: L2 --4 'R 2) extends to a bounded linear projection Pp: Lp --4 9p (resp. Qp: Lp --4 'Rp), provided 1 < p < 00, and this fails if p = 1 or p = 00. (Warning: It is customary in harmonic analysis to consider that Pp and P2 are the "same" operator, since they coincide on simple functions.) In the operator space setting, the situation is analogous: For any 1 < p < 00, Pp (resp. Qp) is a c.b. projection from Lp onto 9p (resp. onto 'Rp). This can be seen easily using Proposition 9.5.2 and the fact that, when 1 < p < 00, Sp is a K-convex Banach space in the sense of [P16]. (See also [TJ1, p. 86] or [DJT, p. 258].) This can also be viewed as a corollary of Theorem 9.8.4, since the latter result implies that (Qp)* ~ 9p' and ('Rp)* ~ 'Rp' (completely isomorphically) when 1 < p,p' < 00 with ~ + -;, = 1. Indeed, the complete boundedness of the natural mapping
(9p )* =
Lp' /G;
--4
9p '
is clearly equivalent to the complete boundedness of Pp, and similarly for Qp. Using the last assertion in Theorem 9.8.2, we obtain another striking isomorphism.
9. Examples and Comments
197
Theorem 9.S.7. Let 1 :s; p < 00. Let Ep be as in Corollary 9.7.2. Then the correspondence Ci -+ )..(gi) (resp. 1i -+ )..(gi») is a complete isomorphism between the spaces Rp (resp. Qp) and Ep. Proof. It suffices to prove this for p ;:::: 2. Let (Xi) be a finitely supported sequence of elements of Spo Let
and
II(Xi)ll,\ = IILxi ® )..(g;) II S,,[E,.j . By Proposition 9.5.3, it suffices to show that these two norms are equivalent for each 2 :s; p < 00. By Proposition 9.5.3 again (isometric case), the norm II . lie coincides with the norm induced by Sp[Lp(n, A, P)], which, by "Fubini's Theorem" (cf. [P2, (5.6)]), coincides with that of Lp(n, A, P; Sp), or, equivalently, with the norm in the middle of (9.8.3). Similarly, the norm II 11,\ coincides with the norm induced by the space Sp[Lp(r)]. The latter space is identical (by Fubini again; cf. [P2, (5.6)]) to the noncommutative Lp-space associated to B(€2)®M equipped with the trace tr ® r, which we will denote simply by Lp(tr ® r). Let (~;) be any choice of signs ~i = ±l. By a well-known result, the linear map that takes )..(g;) to ~i)..(gi) (i = 1,2, ... ) extends to an isometric C* -representation from Al to Al and (by transposition) also defines a complete isometry from Af* to M*. By interpolation, the same map defines a complete isometry on Lp(r) for all 1 < p < 00. In particular, this implies that
and, therefore, after averaging over
II (Xi) 11,\ =
~i =
±1, we have
J
II (CiXi)lI,\dP.
n
(Note: A slightly weaker result, namely, the equivalence of the last two expressions, follows from Corollary 9.7.2, and this suffices for the present argument.) If we set Yi = Xi ® )..(gi) E Sp ® Lp(r), this yields
II (Xi) 11,\
=
J
IILC;Yill L,.(tr
<8Ir)
dP;
hence, by (9.8.3),
II(Xi)ILx :s; B~ max
{II (LY;Yi) 1/211
L,,(tr <8Ir)
'
I (LYiY;) 1/211
L,.(tr <8Ir)
}
Introduction to Operator Space Theory
198 But since
Yiyi = XiX; Q9 1 and yiYi = XiXi Q9 1, we obtain finally
lI(xi)IIA
~ B~ max {II (L X;Xif/21Is" ,II (Lxix:) 1/2 I sp } ~ B~II(xi)IIA.
In other words, we conclude, that the norms
II·IIA and II· liE
are equivalent .•
Curiously, the preceding result combined with the earlier Corollary 9.7.2 implies a result of independent interest on the couple (R, C), as follows.
Corollary 9.S.S. For any 1 < p < isomorpllic identities
00
and ()
=
lip we have completely
(R n C, R + C)o ~ RlP] n ClP]
if p ? 2
(RnC,R+C)o~RlP]+ClP]
ifp~2.
Proof. Indeed, by Corollary 9.7.2, the left side can be identified with E p , and, by Theorem 9.8.5, the right side can be identified with np. Thus the result follows from Theorem 9.8.7. • Remark. By a different argument (see [P2, p. 109-110]), one can show that the equivalence constants appearing in the preceding corollary remain bounded uniformly when p runs over the whole interval) 1, 00(. Thus, together with (9.7.3), this implies completely isomorphic identities Ep ~ RlP] n ClP] Ep ~ R[p]
+ C[p]
if p ? 2 if p
~
2,
with equivalence constants independent of 1 < p < 00. 'When p is an even integer, the best possible constants are computed in the remarkable paper [Buc2].
Remark 9.S.9. Let k ? 1 be a fixed integer. Let (e;)n~1. (e;)n~1. ... ' (e~)n~1 be independent copies of the original sequence (en) as above, on a suitable probability space (0, A, P). Let us denote by n~ the subspace of Lp(O, A, P) spanned by the functions of the form e~k (nl ? 1, n2 ? 1, ... ). Then, modulo a simple reformulation, the results of the paper [HP2] describe the operator space structure of the space nt for any k = 1,2, ... and its dual. (The Gaussian case is similar by general arguments.) The paper [HP2j also describes the space EA Q9min ••• Q9~in EA (k times) and proves that nt is completely isomorphic to (EA Q9min •.• Q9 m in EA) *. Here, of course, the isomorphism constants depend on k.
e;l e;2 ...
9. Examples and Comments
199
Concerning n~ for 1 < P < 00, it is easy to iterate the inequalities appearing in Theorem 9.8.2 to obtain (after successive integrations) two-sided inequalities describing the operator space structure of n~. To describe these iterated inequalities, assume for simplicity that k = 2. Let (Xij) be a matrix with entries in Sp, with only finitely many of them nonzero. Then both x = (Xij) and the transposed matrix t x = (x ji) can be viewed as elements of Sp on the Hilbert space £2 ED £2 ED .. " and we denote the corresponding norms simply by lix II s,' and Iltxlls". Then, after iteration, (9.8.3) becomes, when 2 ~ p < 00,
}
(9.8.9)
s,'
i,j
';(B;l' max
L,,(O.P;S,,)
{IIXIIS"
lI'xlis.,
Moreover, the orthogonal projection induces a c.b. projection from Lp(0., P) onto n~, for any k = 1,2, ... , so that (9.8.9) can be dualized to treat the case 1 < p < 2. We do not spell out the corresponding inequality. Again, when 2 ~ p < 00, these inequalities can be interpreted as describing n~ as completely isomorphic to the intersection of four operator spaces, as follows. First recall that, for any Hilbert space H, we denote by He (resp. Hr) the operator space obtained by equipping H with the O.S.s. of the column (resp. row) Hilbert space. Let us define Hr[P] = (Hr' Heh/ p and Hc[p] = (He, Hrh/ p ~ H,.[p'] for 1 < p < 00. (Also set, by convention, Hr[oo] = H r , He [00] = He,
Hr[l] = He, Hc[l] = Hr.) Then (9.8.9) can be interpreted as saying that n~ is completely isomorphic to the intersection SpnS~pn(S2)e[P]n(S2)r[P]. Thus we can extend essentially all the preceding discussion of np to the spaces n~. In particular, here is what becomes of Corollary 9.8.8 in the case k = 2: Let () = lip, 1 < p < 00, and let us denote K by Soo. Then the interpolation space
is completely isomorphic to the intersection
200
Introduction to Operator Space Tl1eory
and to the sum
In the case p = 1, the results of [HP2] show that n~ is completely isomorphic to the sum (S2)r+(S2)c+Sl +Srp • The case of a general k > 2 can be handled similarly, and we obtain for p 2: 2 (resp. p :::; 2) the intersection (resp. the sum) of a family of 2k operator spaces. We leave the details to the reader (see [HP2] for the cases p = 1 and p = 00). Remark. By a well-known symmetrization procedure, one can deduce from the Khintchine inequalities that, for any sequence (Zn)n::~:l of independent mean zero random variables in Lp (1 :::; p < 00), we have (for any n)
Note that the partial sums Sn L~ Zi form a very special class of martingales. The preceding inequalities were extended to the case of general martingales by Burkholder, Davis, and Gundy (see [Bur]). For a noncommutative version of the Burkholder-Gundy inequalities, with an application to Clifford martingales and stochastic integrals, see [PX].
9.9. Semi-circular systems in Voiculescu's sense In his recent and very beautiful theory of "free probability," Voiculescu discovered a "free" analog of Gaussian random variables; see [VDN] (see also [HiP]). This discovery gives a new insight into a remarkable limit theorem for random matrices, due to Wigner (1955). In Wigner's result, a particular probability distribution plays a crucial role, namely, the probability measure on lR. (actually supported by [-2,2]) defined as follows: /.lw(dt) = 1[-2,2J
J4=t2 dt/27r.
We will call it the standard Wigner distribution. We have
In classical probability theory, Gaussian random variables playa prominent role. They usually can be discussed in the framework attached to a family (-Yi)iEI (resp. (7i)iEI) of independent identically distributed (LLd. in short) real- (resp. complex-) valued Gaussian variables with mean zero and
9. Examples and Comments
201
L 2 -norm equal to 1. When (say) I = {I, 2, ... ,n} the distribution of CYi)iEl (resp. CYi)iEl) is invariant under the orthogonal (resp. unitary) group O(n) (resp. U(n)).
In Voiculescll's theory, stochastic independence of random variables is replaced by freeness of C* -random variables. We will review the basic definitions below. After that, we will introduce a free family (lFi)iEl of C* -random variables, each distributed according to the standard \Vigner distribution. Tlwse are called free semi-circular variables. The family (Wi)iEI is the free analog of CYdiEI in classical probability; it satisfies a similar distributional invariance under the orthogonal group. But actually, since we work mostly with complex coefficients, we will also introduce a free family (lV;)iEI that is the free analog of CYi)iEI; their "joint distribution" satisfies an analogous unitary invariance. Such variables are called free circular variables. \Ve now start reviewing the precise definitions of the basic concepts of "free probability," following [VDNJ. Definitions. A C* -probability space is a unital C* -algebra A equipped ,dtll a state 'P (a state is a positive linear form of norm 1). lYe will say that an element x of A is a C*-random variable (in short, C*-r.v.). Ifx is self-adjoint, we will say tl1at it is a real C* -r. By definition, the distribution of a real C* -r. v. x is tlle probability measure ILx on JR such that
'T.
It follows that, for any continuous function f: JR 'P(f(x)) =
J
---->
f(t)/Lx(dt),
JR, we haye (9.9.1)
Indeed, we can approximate f by a sequence of polynomials uniformly on every compact subset. Hence, ill particular, for all 0 < p < 00, (9.9.2) l\,Ioreover, if 'P is "faithful" on the C* -algebra Ax generated by x (meaning that 'P(y) = 0 for y ~ 0 implies y = 0), then the support of /Lx is exactly the spectrum of x, denoted by a(x). Therefore, we can record here the following fact: Let (A, 'P) and (B, 1/J) be two C* -probability spaces with 'P and 1/' faithful. Let x E A and y E B be two real C* -r. v. with the same distribution, that is, such that /Lx = /Ly. Then we necessarily have IIxll = lIyll (where IIxll is the norm. in A and lIyll the norm. in B).
202
Introduction to Operator Space Theory
This property is immediate, since
Ilxil = SUp{IAII AE a(x)}.
(9.9.3)
It can also be obtained by letting p tend to infinity in (9.9.2). Note that it suffices that
=
J~tr(a(w))dP(w).
Assume moreover that a(w) = a(w)* almost surely. Let (Al(W), ... , An(W)) be the eigenvalues of the matrix a(w). Then the distribution J-ta of the real C*-r.v. a is nothing but
JL 1
J-ta
=
~
n
8)";(w)dP(w).
1
Definitions. Let (An,
More generally, we can define the joint distribution of a family x = (Xi)iEI of real C*-r.v.'s, but it is no longer a measure: We consider the set of all polynomials in a family of noncommuting variables (Xi)iEI. First we define
then we extend F linearly to a linear form on P(J). We will say that F is the "joint distribution" of the family x = (Xi)iEI. If we give ourselves for each n such a family (Xf)iEI with distribution Fn, we say that (Xf)iEI converges in distribution to (Xi)iEI if F n converges pointwise to F.
9. Examples and Comments
203
Let (A,cp) be a C*-probability space and let (Ai)iEI be a family of subalgebras of A. We say that (Ai)iEI is free if CP(ala2 ... an) = 0 every time we have aj E Ai.i,il -I- i2 -1- ... -I- in and cp(aj) = 0 Vj. Let (Xi)iEI be a family of C*-r.v.'s in A. Let Ai be the unital algebra (resp. C* -algebra) generated by Xi inside A. We say that the family (Xi)iEI is free (resp. *-free) if (Ai)iEI is free. The preceding definitions are restricted to C*-r.v.'s, which correspond to bounded r.v.'s in the commutative case, but, by convention, we will use the same definition whenever we are dealing with a sequence in a noncommutative Ll-space. So, from now on, convergence in distribution or weak convergence means the convergence of all moments just like in the preceding definitions (of course, for this to make sense, we implicitly assume that all moments exist, like in the Gaussian case, for instance). When the limit distribution is determined by its moments (in particular, if it is compactly supported, or in the Gaussian case) this coincides with the usual notion of convergence in distribution. We can now reformulate Wigner's Theorem in Voiclliescu's language. Fix n ~ 1. We introduce the random (real symmetric) n x n matrix
an
=
(9ii h:S;ij:S;n
with entries defined as follows: {9ij 1 i ::; j} is a collection of independent Gaussian real-valued r.v.'s with distribution N(O, l/n) (Le., E(9ij) = 0 and EI9ijl2 = l/n) and 9ij = 9ji Vi > j. We assume these (classical sense) random variables defined on a sufficiently rich probability space (for instance, the Lebesgue interval). Let An = Loo(O, A, P; lIIn) and let CPn be the state defined on An by setting Vx E An
CPn(X)
=
J~tr(x(w))dP(w).
Then, Voiculescu's reformulation of Wigner's Theorem is Theorem 9.9.1. If we consider an as a real C*-r.v. relath;e to (An' CPn), tllen we have tlle weak convergence of probability measures: I-lG"
-4
I-lw
when n
-4
00.
More generally, Voiculescu showed: Theorem 9.9.2. Let (ai )iEI be a family of independent copies (in the llsual sense) of the random variable an. Then, when n - 4 00, the family (ai)iEI converges in distribution to a free family (Wi)iEI of real C*-r.v.'s eac11 with the same distribution equal to I-lw.
We will say that a real C*-r.v. X is semi-circular if there exists A > 0 such that the distribution of AX is equal to I-lw. If A = 1, and if X admits exactly
Introduction to Operator Space Theory
204
J.lw for its distribution, then we will say that x is semi-circular standard. We then have cp(x) = 0, cp(X2) = 1. (We should warn the reader that our standard normalization differs from that of [VDN].)
Actually we even have an almost sure result as follows:
Theorem 9.9.3. Let (Gf )iEI be as above. Then, for almost all w in 0, we have sup IIGf(w)IIMn <
00
n
for each i and, moreover, the distribution of (Gf(W))iEI (on Mn equipped with its normalized trace) tends to that of (Wi)iEI when n --+ 00.
Proof. This is based on a (so-called) "concentration of measure" argument. Actually, the most basic form of Sobolev's inequality in the Gaussian setting will suffice for our purposes. Let {gm} be any collection of independent standard (Le., N(0,1)) real-valued Gaussian random variables. Then it is well known (cf., e.g., [Chel) that, for any polynomial F = F(gbg2," .), we have (9.9.4) where
Now let P and let
= P({Xd)
be a polynomial in noncommutative variables {Xi},
As we already mentioned, we have almost surely lim sup IIGfllM" n-+oo
and also for any p <
< 00
00
limsuplEIIGfll~n n-+oo
<
00.
•
Using this, it is rather easy to deduce from (9.9.4) that there is a constant Cp depending only on P such that (9.9.5) Indeed, it suffices to majorize IIV' FII2 accordingly when P is a monomial, that is a finite product Xit X i2 ••• Xi", and in that case the preceding observations
9. Examples and Comments
205
and the specific normalizations ofTn and G? yield (9.9.5). Then (9.9.5) implies for any E > 0 P(lFn -lEFnl > E) < 00,
L
and hence for almost all w lim sup IFn -lEFnl = O. n ....... oo
On the other hand, by Voiculescu's central limit theorem (i.e., Theorem 9.9.2) we know that lEFn ---> r(P({Wd))· Hence, we conclude that for almost all w we have
Tn(P({G?(w)}))
--->
(9.9.6)
T(P({Wd))·
Finally, taking a suitable countable intersection, we find a subset n' c n with p(n') = 1 such that, for any win n', (9.9.6) is satisfied by all polynomials P with (say) rational coefficients and moreover sUPn IIG?(w) II < 00. This yields the announced result. •
Remark. In the preceding proof the passage from convergence in distribution to the almost sure one is based on a very general principle called the concentration of measure phenomenon. which has many important applications in geometry and analysis (see [l'dS]). As pointed out by Voiculescu in [V02, pp. 216-217], the random matrix situation (either in the Gaussian or unitary case, for instance) is usually so "extremely concentrated" that results such as the preceding one are then easy to derive using this principle. Note in passing that we chose to use only the elementary Sobolev bound (9.9.4), but more refined estimates are actually available (see, e.g., [1\1S] or [P8, pp. 4448]). For other methods to obtain similar almost sure results see [HiP, T]. In Voiculescu's theory, the analog of an independent family of standard real Gaussian variables is a free family of standard semi-circular C* -r.v. 'so Such a family can be realized on the full Fock space, as follows. Let H = £2(1). Recall we denote by F(H)(or simply by F) the full Fock space associated to H; that is to say, we set 'Ho =
We consider from now on 'Hn as a subspace of F. For eVf'ry h by €(h): F ---> F the operator defined by: €(h)x
=
h@x.
E
F, we denote
Introduction to Operator Space Theory
206
More precisely, if x = A1 E ?to = C1, we have £(h)x = Ah, and if x = 0 X2'" 0 Xn E ?tn, we have £(h)x = h 0 Xl 0 X2'" 0 X n • We will denote by n the unit element in ?to = C1. The C* -algebra B(:F) is equipped with the state c.p defined by cp(T) =< Tn, n > . Xl
Let (ei)iEI be an orthonormal basis of H. The pair (B(:F),cp) is an example of a C*-probability space. Moreover, cp is tracial on the C*-algebra generated by the operators £(ei) + £(ei)* (i E I), that is, we have cp(xy) = cp(yx) for all x, y in this subalgebra. (Note however that cp(£(h)*£(h)) = (h, h) and cp(€(h)f(h)*) = 0, so that cp is not tracial on the whole of B(:F)!) In this subalgebra, let
Then the family (Wi)iEI is an example of a free family of standard semicircular C* -r.v.'s, or, in short, a standard semi-circular free family. This family enjoys properties very much analogous to those of a standard independent Gaussian family (gdiEI. Indeed, for every family (O:i)iEI E lR(I) with E O:~ = 1 the real C* -r.v. S = EiEI O:i Wi admits /-lw as its distribution. This is analogous to the rotational invariance of the usual Gaussian distributions. More explicitly, this means that, for every continuous function f: lR ---+ lR, we have
cp(f(S» =
!
f(t)/-l\V(dt).
In particular, by the fact preceding (9.9.3), for all finitely supported families of real scalars we have
II LO:iWili = 2(Lo:~)1/2. iEI Thus, the operator space lR-linearly generated by (Wi)iEI is isometric to a real Hilbert space. We now pass to the complex case. Let (Zi)iEI be a family of (not necessarily self-adjoint) C*-r.v.'s. We can then consider the distribution F of the family of real C*-r.v.'s obtained by forming the disjoint union of the family of real parts and that of imaginary parts of (Zi)iEI. We will say that F is the joint *-distribution of the family (Zi)iEI. Of course, if the family is reduced to one variable Z, we will say that F is the *-distribution of Z. Note that the data of the *-distribution of (Zi)iEI are equivalent to those of all possible moments of the form
cp(Xit X i2 where Xi
= either Zi
••
.Xin ),
or Z; and where il, i2 •...• in are arbitrary in I.
9. Examples and Comments
207
We now come to the analog of complex Gaussian random variables. Let (W', W") be a standard semi-circular free family (with two elements). We set W = ~(W' + iW"). Every C*-r.v. having the same *-distribution as TV (resp. as AlV for some A > 0) will be called "standard circular" (resp. "circular"). Suppose we are given a (partitioned) orthonormal basis {ei liE I} u {Ii liE I} of H. Then, one can show that Wi = £(ei) + £Ui)* is a *-free family of standard circular C* -r.v.'s (in short, a standard circular *-free family). Now, let (TVi )iEI be any *-free family formed of standard circular variables. Then, for any finitely supported family (O:i)iEI of complex scalars with L IO:il 2 = 1, the variable S = LiEI O:i Wi has the same *-distribution as W. As above, we have (9.9.7) II LO:iWil1 = 2(L IO:iI 2)1/2 (one can verify that IIWII
= 2).
Let VI be the operator space spanned by this family {lVi 1 i E I}. By (9.9.7), VI is isometrically Hilbertian and (Tl'i)iEI is an orthonormal basis. l\Ioreover (see [VDN, p. 56]) for any isometric transformation U: VI -+ VI the family (U(Wi»iEI has the same *-distribution as (Wi)iEI. In particular, by the next lemma, this implies that the operator space VI is homogeneous.
Lemma 9.9.4. Let (A, if) and (B, f/!) be two C*-probability spaces with if and'l/J faitllful. Let (Zi)iEI and (Yi)iEI be two families ofC*-r.v. 's ill A and in B, respectively, admitting the same joint *-distribution. Let A z (resp. By) be the C*-algebra generated by (Zi)iEI (resp. (Y;)iEI), and let Ez C Az (resp. E y C By) be tIle operator space spanned by the families. Then the linear mapping U defined by U(Zi) = Yi extends to a complete isometry from E z onto E y and actually to an isometric representation from Az onto By. Proof. Without restricting the generality, we may replace the family (Zi )iEI by the disjoint union of the families (Z;)iEI and (Z;)iEI , and similarly for the family (Yi)iEI. Then let P = LO:iti2 ... i~.Zil ... Zh. be a polynomial with complex coefficients in the noncommutative variables (Zi )iE I. We set
Then, since (Zi) and (Y;) have the same joint *-distribution, P* P and 7r(P)*7r(P) have the same distribution; hence, by the fact stated before (9.9.3), since if and 'l/J are faithful, IIPII = 117r(P)II. In particular, 7r extends to an isometric representation from Az onto By. A fortiori (cf. Proposition 1.5), t.he restriction U of 7r to Ez is completely isometric. •
208
Introduction to Operator Space Theory
Actually, it is very easy to identify the operator space VI (up to complete isomorphism), as the next result shows (see [HP2] for some refinements).
Theorem 9.9.5. The operator space VI gellerated by a standard circular *free family (Wi)iEI is Hilbertian and homogeneous. Similarly, the closed span of a free semi-circular family {(Wi)iEI} is 2-Hilbertian and 2-homogeneous. Moreover, if (say) I = N, each of these spaces is completely isomorphic to R n C or, equivalently, to E>.. Proof. We already saw that VI is Hilbertian and homogeneous. Let (ai)iEI be a finitely supported family in B(H). The identity Wi together with L £( ei)£( ei)* ::; I yields
=
£( ei) + £( ei)*
II I>i ® Willmin ::; II I>i ® £(ei)IImin + II I>i ® £(ei)*IImin ::; II L ai a ill 1/ 2 + II L ai ai'II 1/ 2 whence (9.9.8) Conversely, it is easy to check that ip(Wt W j ) = ip(Wj Wt) 0 if i :I j and = 1 otherwise. Hence, letting T = L ai ® Wi, we have II L ai a; II = 11(1 ® ip)(T*T)II ::; IITII~in' and similarly we have II Laiai'li ::; IITII~in' It follows that (9.9.9) The inequalities (9.9.8) and (9.9.9) imply that span[Wi liE I] is 2-Hilbertian and 2-homogeneous. For simplicity, we assume I = N in the rest of the proof. By (9.7.3) and Theorem 9.7.1, the last two inequalities imply that the closed span of (Wi)iEI is completely isomorphic to E>. or, equivalently, to R n C. Finally, as the variables lVj = (lV; + iW;')2- 1 / 2 appear as a sequence of "blocks" (normalized in £2) on a standard semi-circular system, the same inequalities (9.9.8) and (9.9.9) remain valid if we replace (W;);EJ by (W; )iEI. Therefore, we conclude that VI itself is completely isomorphic to E>. or to R n C. This· last point can also be deduced from the concrete realization Wi = £(ei)+£(Ji)* already mentioned for a standard circular *-free system. •
Remark 9.9.6. Let M be the von Neumann algebra generated by a free semicircular family (Wi)iEI. We assume I = N for simplicity. Recall a classical notation: For any x in M, we define Xip E M .. by Xip(Y) = ip(Yx) for all y in M. Thus we obtain a continuous injection M - M .. that allows us to
9. Examples and Comments
209
consider the int.erpolation spaces (1I1,lIf*)o for 0 < () < 1. Let us denot.e for simplicity L'X!(
T(x) =
L8
i
Let V: R n C -+ 1If be the mapping defined by V(8 i ) = Wi. By (9.9.8), the composition VT: M -+ M satisfies IWTllcb ~ IWlicb ~ 2. Moreover, VT is the adjoint of an operator on l\{* and VT "coincides" with P on the *-algebra generated by (Wi)iEI. Therefore, VT is a completely bounded projection from III onto Woo, which naturally extends P. By transposition, we obtain a c.h. projection from LI (
Theorem 9.9.7. For simplicity, let I = N. Let (Wi)iEI (resp. (Wi)iEI) be a standard free semi-circular (reps. *-free circular) family. For 1 ~ P ~ 00, let Wp (resp. Wp) be the closed span of (Wi)iEI (resp. (Wi)iEI) in Lp(
Introduction to Operator Space Tl1eory
210
Note that there is a Fermionic variant in which a free semi-circular family appears as the limit or suitably normalized matrices with entries satisfying the CAR. On the other hand, the reader will find in [Sk] a description of the applications of Voiculescu's theory to von Neumann algebras.
9.10. Embeddings of von Neumann algebras into ultraproducts In this section, we discuss ultraproducts in the von Neumann sense. This is a slightly different notion from the usual one considered in §2.8. We will prove in detail that the von Neumann algebra of a free group embeds into a (von Neumann sense) ultraproduct of finite-dimensional matrix algebras. We will also describe several important related open questions for which operator space theory might be useful. Let {M(n) I n ~ I} be a sequence of von Neumann algebras equipped with normal faithful traces {Tn I n ~ I} with Tn(1) = 1. Let B = foo({l\f(n) I n ~ I}), and let U be a free ultrafilter on N. We define a functional f E B* by setting for all t = (tn)"l~l in B
Clearly fu is a tracial state on B. Let Hn be the Hilbert space associated to the GNS construction for (M(n), Tn). Note that Hn can be viewed as the noncommutative version of the L 2-space over a probability space. (Indeed, if M (n) is commutative, Tn can be identified with a probability on its spectrum and Hn ::= L2(Tn». Let ~n be the (cyclic) unit vector of Hn corresponding to the identity of M(n). We then have 'V
Xn
E
M(n)
Let Hu = llHn/U. We will denote by (h,;,) the equivalence class in Hu of a bounded sequence (h n ) with hn E Hn for all n. Clearly, M(n) acts by left and right multiplication on Hn. Therefore, passing to the ultraproduct we obtain representations
L:
B
-+
B(Hu)
and
R:
BOP
-+
B(Hu)
defined by
But actually, this Hilbert space Hu is ''too large." We need to reduce it and to consider the restriction of Land R to a smaller space H u , which we now describe. The space Hu can be defined as the Hilbert space associated to the tracial state fu in the GNS construction applied to B. We denote by L: B
-+
B(Hu)
and
R: BOP
-+
B(Hu)
9. Examples and Comments
211
the representations of B corresponding to left and right multiplication by an element of B. More precisely, let p(x) = limu Tn(X~xn)1/2. Clearly p is a Hilbertian semi-norm on B. Let
Iu = ker(p). Then Iu is a closed two-sided ideal, and Hu is defined as the completion of B / Iu equipped with the Hilbertian norm associated to p.
= (tn)n in B we denote by. i the equivalence . class of tin Bllu. Then Land R are defined by L(x)t = ti and R(t)t = &. Clearly, since lu is For any t
tracial, these are contractive representations of B on Hu. Note that we have a natural isometric embedding (9.10.1)
Hu'---->Hu,
which takes t to t. Clearly Land R are nothing but the restrictions of Land R to the invariant subspace Hu. The next result seems to go back to Mac Duff's early work (see [Sa]). Note that the kernel Iu is not weak-* closed in B, so the fact that the quotient B IIu is a von Neumann algebra is a priori somewhat surprising. We call B IIu the (von Neumann) ultraproduct of the family (M(n), Tn) with respect to U. Note that a priori, this is a quotient of the (Banach space) ultraproduct described in §2.8. Theorem 9.10.1. The kernels of Land R coincide with the set'
After passing to tlle quotient, Land R define isometric representations Lu: Bllu
~
B(Hu)
and Ru: BOPllu
~
B(Hu)
with commuting ranges. Moreover, lu defines a faithful trace such that, if q: B ~ B I Iu denotes the quotient map, we have ru(q(t))
=
lu(t)
[Lu(Bllu)l' = Ru(BOPllu)
and
TU
on B I Iu
"It E B.
Finally, tlle commutants satisfy [Ru(BOPllu )]' = Lu(Bllu).
In particular, the ranges Lu (B I Iu) and Ru (BOP I Iu) are von Neumallll subalgebras of B(Hu ), and tlley are factors (i.e., their center is reduced to the scalars) if all the M (n) are themselves factors. Proof. Let In be the unit of M(n) and let fu(t)
=
(L(t)~,~)
=
~ =
(In)n
(R(t)~,~).
E
B. We have
212
Introduction to Operator Space Theory
If L(t) = 0, then L(t*t) = 0, which, by the preceding line, implies fu(t*t) = OJ hence t E Iu. Conversely, if t E I u , then x*t*tx E Iu for any x in Band
hence (Cauchy-Schwarz) fu(x*t*tx) = 0, which means tx = 0 for all x in B or, equivalently, L(t) = O. A similar argument applies for R, so we obtain that ker( L) = ker( R) = Iu. Then, after passing to the quotient by Iu, Land R define the isometric representations Lu and Ru with the same respective ranges. Therefore, Lu and Ru still have commuting ranges. Finally, let T E B(Hu) be an operator commuting with Lu(B/lu), that is, T E Lu(B/lu)'. We will show that T must be in the range of Ru. Let
/3 = T(~)
E
Hu·
We claim that there is b = (b n ) in B such that /3
T
= b and that
= R(b) = Ru(b).
Indeed, we have for any t = (tn) in B TL(t)~
= L(t)T~ = L(t)/3j
(9.10.2)
hence (9.10.3) We now use the embedding (9.12..1). Let (/3n) with /3n E Hn and sup II/3nIIH.. < 00 be a representative of /3 in Hu. Then (9.10.3) implies for any tin B (9.10.4) Let /3n = hnvn be the polar decomposition of /3n in Hn (see the next remark) with hn E H n , hn ~ 0, Vn partial isometry in .M(n), and hn = (/3n/3~)1/2. Fix c > O. Let Pn be the spectral projection of h n associated to )IITII + c, 00(. Note that /3n/3~Pn = h;Pn ~ (IITII + c)2Pn. A priori (Pn) defines an element of flu, but actually, since /3 E Hu, it is rather easy to show that (Pn) also corresponds to an element in Hu. We leave this point to the reader (see the next remark for some clarification). Hence (9.10.4) implies (with tn = Pn)
This forces limu Tn(Pn) = 0, and hence limu Tn(/3n/3~Pn) = 0., Therefore, if we set finally bn = (1 - Pn)hnvn, we find IIbnll :::; 11(1 Pn)hnll :::; IITII + c and lI/3n - bnllkn :::; IIPnhnvnllkn :::; Tn(/3n/3~Pn)j hence limu II/3n - bnllHn = O. Let b = (bn). Note that bE B with IIbllB :::; IITII + c. Then, going back to (9.10.2), we obtain finally
TL(t)e = L(t)/3 = L(t)b = (tnbn) = Ru(b)L(t)e.
9. Exa.mples a.nd Comments
213
This shows that T = Ru(b), which completes the proof that Lu(Bj1u)' = Ru(BOPj1u). The same argument clearly yields Ru(BOPj1u)' = Lu(Bj1u). and hence Lu (B j Iu)" = Lu (B j Iu ), which proves (von Neumann's bicomnl\ltant theorem) that Lu(Bj1u) is a von Neumann algebra. •
Remark. In the above, we invoked the polar decomposition in H n , viewing Hn as the noncommutative L 2 -space relative to Tn and using its structure as a bimodule over M(n). But actually, we will apply the preceding result only in the case when A/(n) is a matrix algebra. In that case, all the points left. to the reader in the preceding proof are especially easy to verify and the polar decomposition is then done in M(n) itself. Remark 9.10.2. By construction, Hu appears as the closure in Hu of the subspace of all elements of the form (b;,) with sUPn IlbnIlAI(n) < 00. Alternatively, Hu c flu can also be described as the subspace corresponding to the "uniformly square integrable" sequences. More precisely, let fi = ((3n )" be an element of flu, with sup" II.BnIIH.. < 00. Then fi belongs to Hu iff
or iff e~~ limu Tn(.B~.Bn1{i3~i3n>e})
= 0,
where we have denoted (abusively) by l{be} the spectral projection of the Hermitian operator h for the interval )c, 00(. A group G is called residually finite if there exists a collection of finite groups (G;) and homomorphisms '-P;: G --+ G; separating the points of G: that is, for any finite subset S c G there is an i for which the restriction of '-P; to S is injective. Without loss of generality, we may assume that G i = Gjr;, where each C eGis a normal subgroup with finite index and '-Pi is the canonical quotient map. Thus, G is residually finite iff it admits a family of normal subgroups with finite index (r i ), directed by (downward) inclusion and such that niEI G i = {e}.
Corollary 9.10.2. ([Wal]) Let G be any countable residually finite discrete group. Then V N( G) embeds into an ultraproduct of the form B j Iu as above with all the algebras AI (n) finite-dimensional. Proof. Let r 11 be a decreasing sequence of normal subgroups of finite index with intersection reduced to the unit e, and let G n = G jr n' Let '-Pn: G --+ G n denote the quotient morphism and let JI[(n) = VN(G n ) equipped with its normalized trace T 11' We consider B = £00 ( {1If (n) In;::: I}) and its quotient Bj Iu as above. For any tin G, we denote by yet) the equivalence class modulo Iu of (),G .. (t))n~l E B.
Introduction to Operator Space Theory
214
Let
7
denote the normalized trace on VN(G). We then have
IW 7 {.XG .. (t)) = 7(AG(t)).
"It E G
n
This implies that the family {y(t) I t E G} has the same *-distribution with respect to 7U as the family {AG(t) I t E G} with respect to 7. Therefore, by Lemma 9.9.4, the von Neumann algebras that they generate are isomorphic, • via the isomorphism taking AG(t) to y(t). The following fact is classical. Lemma 9.10.3. Free groups are residually finite. Proof. Let G = Fl. Let {gi liE J} be the (free) generators. Let C c G be a finite subset. It suffices to produce a (group) homomorphism h: G --+ r into a finite group r such that, for any e in C, we have h(e) '" er if e '" e, where er denotes the unit in rand e the unit in G. We may assume that C c G', where G' is the subgroup generated by a finite subset {gi liE J} of the generators. Let k = max{lel leE C} (recall lei is the length of e). We then set S
= {t E G' Iitl :::;
k}.
We will take for r the (finite) group of all permutations of the (finite) set S. For any i in J, we introduce
Then clearly Si C Sand giSi C S. Hence (since ISil = IgiSil and S is finite) there is a permutation O'i: S --+ S such that O'i(S) = giS for any S in Si' Thus if s,t E S and if git = s (or, equivalently, t = g;ls), we have O'i(t) = S (or, equivalently, t = O';I(S)). Thus it is easy to check that if a reduced word t = g:11 g:; ... g:::: (m :::; k ei = ±1) lies in S (note that, by definition of S, e and all the subwords of t also lie in S), we have
Therefore, if we define h: G --+ r as the unique homomorphism such that h(gi) = O'i Vi E J and O'(gi) = er Vi 1. J, we find, for t as before, h(t) = 0':110':; .. . 0':::: and h(t)(e) = t in particular, we have h(t) '" er whenever t E S and t '" e. Since C C S, we obtain the announced result. • Consequently:
9. Examples and Comments
215
Theorem 9.10.4. ([lVal]) Tile \'011 NeunuI11ll algebra of the free groups lF n or lFoo embeds into a von Neumann uitraprodlzct of matrix a.lgebraB. We wish to give a second proof of the preceding result using Voiculescu's "matrix model" for free semi-circular systems. This approach is well known to specialists. See also [T] for more along this line. Theorem 9.10.5. Let !If be the von Neumann algebra generated by a free semi-circular system (VVi)iEI (as in Theorem 9.9.2) witll I coulltable. Then !If embeds illto a VOll Neumanll ultraproduct of matrix algebras. Proof. Let Tn (resp. T) be the normalized trace 011 AIn (resp. on AI). Let B = £00 ( {lIIn I n 2:: I}). Let U be any free ultrafilter on N and let Iu be as before. By Theorem 9.9.3 (using the same notation), we know that, for almost all w, (Gi(w)k::l is in B. I\,Ioreover, if we denote by Gi(w) the equivalence class modulo Iu associated to (Gi(W))n;?:l, then the family {Gi(w) liE I} has the same distribution with respect. to (BjIu,Tu) as the system {Wi liE I} with respect to T. Thus, by Lemma 9.9.4, the von Neumann algebra.,> they generate must be isomorphic. • In [Col], Alain Connes observes in pa.'>sing the above Theorem 9.10.4 and casually asks whether the same embedding is valid for any III factor. This has become one of the main open problems in von Neumann algebra theory. Actually, since, by desintegration, any finite von Neumann algebra (on a separable Hilbert space) can be seen as the "direct integral" of a field of factors, one may reformulate Connes's question as follows: Problem. Let (!If, T) be a von Neumann algebra (on a separable Hilbert space) equipped with a faithful, norma.l, and normalized trace T. Is tllere a.lways a (trace presenTing) embedding of Af (as a von Neumanll suba.lgebra) in an ultraproduct of matrix algebras?
9.11. Dvoretzky's Theorem
It is natural to wonder whether there is an analog for operator spaces of the famous Dvoretzky Theorem on the spherical sections of convex bodies. This fundamental result can be formulated as follows: Every infinite-dimensional Banach space X contains, for each c > 0, a sequence of subspaces En C X (n = 1,2, ... ) such that En is (1 + c)-isomorphic to e~ for every n. Vitali Milman (see [MS] and [FLM]) gave a remarkable proof based on Paul Levy's isoperimetric inequality on the unit sphere of lRn. It is then tempting to try and find in every (infinite-dimensional) operator space X a sequence of subspaces (En) uniformly completely isomorphic to OHn . Unfortunately,
216
Introduction to Operator Space Tlleory
this is absurd: For a counterexample it suffices to consider a homogeneous Hilbertian X different from OH (for instance, R or C). Then, by Proposition 9.2.1, every n-dimensional subspace En C X is completely isometric to the ndimensional version of X (for instance, Rn if X = R). Therefore, one cannot find inside X anything else but "copies" of X. The best that one can hope to find in an arbitrary (infinite-dimensional) operator space seems to be a sequence of almost Hilbertian and almost homogeneous subspaces (En) (here "almost" means up to €) with dim En = n for aU n. In this form, it is rather easy to adapt Milman's ideas (the concentration of measure phenomenon; see [MS]) to the quantum situation. There is a slight difficulty due to the noncompactness of the unit ball of the quantum scalar coefficients, that is, the unit ball of IC (see the discussion in §2.11). To circumvent this difficulty, we choose the language of ultraproducts, which is convenient in the present situation. See §2.8 for precise definitions. With the ultraproduct terminology, we can reformulate Dvoretzky's Theorem as follows: Theorem 9.11.1. Let (Xi)iEI be a family of infinite-dimensional Banacll spaces (or merely sucll tllat limu dim(Xi ) = 00). Tllen tlleir ultraproduct lliEIXi/U contains a subspace isometric to £2. The "quantum" version of this statement is then (cf. [P9]): Theorem 9.11.2. Let (Xi);E! be a family of infinite-dimensional operator spaces (or merely sucll tllat lirnu dim(X;) = 00). Tllen tlleir ultraproduct lliEI Xi/U contains a subspace completely isometric to an infinite-dimensional Hilbertian llomogeneous operator space. This statement explains perhaps why Hilbertian homogeneous operator spaces appear so often in "nature."
Chapter 10. Comparisons In the previous chapters, we have met the following examples of homogeneous (or A-homogeneous) Hilbertian operator spaces:
OH,
<1>,
R,
C,
R n C,
R
+ C,
min(€2),
max(€2)'
We will now show that all the spaces in that list are mutually completely nonisomorphic. More precisely, let E, F be any two spaces in the preceding list, and let En, Fn be their n-dimensional version. We will show that dcb(En , Fn) --+ 00, and we will give an asymptotic estimate of dcb(En' Fn) when n --+ 00. Our aim is to illustrate what can be done with the ideas of operator space theory, by elementary tools. In the very early days of Banach space theory, methods were discovered to distinguish, up to isomorphism, the so-called classical Banach spaces, such as the spaces €P or LP (see [Ba, Chapter 12]) and also to estimate the Banach-Mazur distances d(€~, €~) with 1 :=:; p =I- q :=:; 00. Our aim in this chapter is to initiate an analogous study for operator spaces. Many estimates in this chapter come from C. Zhang's PhD thesis ([Z3]) or can be easily derived by his methods. (The only exception may be the fact that and max(R 2) are not isomorphic, which was only established in [JP].) The following result, from [Z3], will be very convenient throughout this chapter (see also [ZI, Z2]). Proposition 10.1. Let E, F be two Hilbertian homogeneolls operator spaces. Assume for simplicity that E and F are isometric to €2. Let (en) (resp. (in)) be an orthonormal basis of E (resp. F). Let En = span[eJ, ... , en] (resp. Fn = span[ft, ... , In])·
(i) Let Un: En
--+ Fn (resp. u: E --+ F) be the (isometric) linear map defined by un(ei) = Ii (resp. u(ei) = Ii). Then we llal'e
(10.1) and dcb(E, F) = Ilullcbllu-11lcb = sup Ilunllcbllu~lllcb.
(10.2)
n
(ii) We also have (10.3)
Proof. Using the polar decomposition of an n x n matrix, it is easy to see that dcb(En, Fn) = inf{lIvllcbllv-1Ilcb}, where the infimum runs over all maps v: En --+ Fn of the form v(ei) = Adi with Ai > O. Let a be a permutation of [1,2, .. , n]. Let Uu (resp. Vu ) be the unitary transformation defined on En by
Introduction to Operator Space Theory
218
Uu(ei) write:
= eU(i) (resp. Vu(fi) = !u(i»). Since En, Fn are homogeneous, we can
Since ~ L:u Vu-l V Uu = (l/n)(L: Ai)Un , we deduce Ilunllcb(l/n)(L: Ai) ::; IIvllcb. Similarly, exchanging the roles of En and Fn , we find Ilu;;lllcb(l/n) (L: Ail) ::; IIv- 11Icb. By Cauchy-Schwarz, we have 1 ::; n- 1 (L: Ai)n- 1 (L: Ai 1), whence Taking the infimum over all possible v, we find
Since the inverse inequality is obvious, we obtain (10.1). Let w: E ---4 F be an arbitrary isomorphism. By Proposition 9.2.1, w(En) is completely isometric to Fn. Therefore we have dcb(En' Fn) ::; deb (En , w(En)) ::; Ilwllebllw-11Ieb, whence deb(En , Fn) ::; deb(E, F). This implies (10.2)' sup Ilunllebllu;;llleb ::; dcb(E, F). n
Note then that lIulicb (resp. lIu- 1I1eb) is the nondecreasing limit of lIunlleb (resp. lIu;;ll1eb) when n ---4 00. From the obvious inequality dcb(E, F) ::; lIullcbllu-11Ieb, we deduce the converse of (10.2)' and obtain (10.2). To show (ii), observe that if u: E~ ---4 En denotes the map defined by the canonical isometry from E~ onto En, and if v: OHn ---4 En is an arbitrary isometry, we have u = vtv hence lIulleb = Ilvll~b by (7.9). • Remark. In [Z2], Zhang made the interesting observation that, for any normed space E,
deb(min(E), max(E» = II min (E)
---4
inax(E)lIeb,
or, equivalently, deb(min(E),max(E» is equal to Paulsen's constant a(E) described in Chapter 3. The argument is quite simple: Let iE: min(E) ---4 max(E) be the identity. Then, for any complete isomorphism u: min(E) ---4 max(E) we have iE = (iEU- 1 )u, but since max(E) is homogeneous, we have lIiEU-Illeb = IliEU-111 = lIu- 111; hence
IliEllcb::; which implies
lIiEU-Illcbllulieb :$
lIu-Ililiulieb :$ Ilu-Illebllulleb,
IliElieb :$ deb(min(E),max(E».
10. Comparisons
219
•
The converse is obvious.
Convention. Throughout the sequel, we will write an ~ bn if there exist positive constants C1 and C2 such that C1 an :S bn :S C2an for all n. Notation. Let En and Fn be two (isometrically) Hilbertian operator spaces of dimension n. Assume Fn homogeneous. If u: En --+ Fn and v: En --+ Fn are two isometries; then Ilulicb = Ilvllcb. Indeed, uv- 1: Fn --+ Fn and vu- 1: Fn --+ Fn are isometric; hence by the homogeneity of Fn we have Iluv- 1 1lcb = Ilvu- 1 1lcb = 1, which obviously implies lIulicb :S Iluv-1l1cbllvllcb = Ilvllcb and conversely IIvllcb :S Ilulicb. We will simply denote by
the c.b. norm of any isometry from En into Fn. By the triangle inequality, this c.b. norm is :S
11.,
whence (10.4)
Therefore we can rewrite (10.1) as follows: (10.1)' We start by estimating the distance dcb(OHn , En) when E is anyone of the spaces in the above list. Recall that we know (Corollary 7.7) that, if En is an arbitrary n-dimensional operator space, we have dcb(OHn , En) :S yn. Proposition 10.2. The following estimates hold: (10.5)
dcb(OHn , min(e~)) = dcb(OHn , max(e~)) ~ yn, deb (OHn , n) = dcb(OHn , ~) ~ dcb(OHn , Rn n C n ) = dcb(OHn , Rn
Vii,
+ Cn) =
(10.6) (10.7)
n 1/ 4 •
(10.8)
Proof. \Ve have
IIOHn
--+
Rnllcb =
lit eli Q9 elil1112 1
IDltl
= n
1/ 4 ,
(10.9)'
Introduction to Operator Space Theory
220
and similarly with C n instead of Rn; whence (10.5). We could also deduce this from (10.3) and the identity dcb(Cn , Rn) = n already mentioned above. By Theorem 3.8
II min(e~) ~ max(e~) Ilcb := n,
(10.10)
whence (10.6) by (10.3) and (10.1)'. Proof. To show (10.7), let {Vi} be as in §9.3. Note that IIQHn ~ nllcb = II 2:: Vi ® Vill:fi~' Then (with the same notation as in the proof of Proposition' 9.7.3) II 2:: Vi ® Villmin ~ 2::7 T(ViY:*) = n/2, and hence IIOHn ~ nllcb ~ (n/2)1/2. Conversely, we have obviously (by the triangle inequality) 112:: Vi ® -
1/2 Vi II min
~
y'n, and hence (10.11)
By Proposition 9.7.3 and by (2.11.4), we have IIn ~ OHnllcb ~ J2. This establishes (10.7), taking into account (10.1), and Theorem 9.3.1. Finally, to check (10.8), we observe
In the converse direction, the inequality (2.11.4) implies IIRnnCn ~ OHnllcb = 1. Then by (10.1) we have (10.8). • Note in passing that, since IIRn n C n ~ OHnllcb 9.7.3) IIn ~ Rn n Cnllcb := 1,
=
1 and (see Proposition (10.12)
we have (taking (10.4) into account) (10.13) and hence lIn
~ ~lIcb :=
1. By (10.11) and (10.3) we therefore have II~ ~ nllcb := n.
Moreover, since we have trivially IIOHn ~ min(e~)lIcb =
(10.14)
II max(e~)
~
OHnllcb = 1, (10.6) implies
II min(e~)
~ OHnllcb
= IIOHn ~ max(~)lIcb := Vn.
(10.15)
By Proposition 1O.1(ii) we immediately deduce from the preceding statement:
10. Comparisons
221
Corollary 10.3. The following estimates hold:
In the next statement we compare
cf>n
to the other spaces in the list.
Theorem 10.4. The following estimates hold: (10.16)
deb(cf>n,Rn nCn)::o.;n,
(10.17)
+ Cn ) ::0 .;n, deb(cf>n' min(£~)) ::0 n, debe cf>n, max( £~)) ::0 .;n.
( 10.18)
debe cf>n, Rn
Proof. It is easy to check that that
(10.10) (10.20)
II L \!;*Vill ::0 II L Vi\!;* 11::0 n
(in fact
= 11), so (10.21)
Conversely, we have (10.12); hence a fortiori (recalling (10.4)) (10.22) whence (10.16). To check (10.17), let c = IIRn n C n
-+
cf>nllcb.
vVe have clearly
Conversely, we can show (cf. [HP2, Remark 1.2]) that there exists a decomposition Vi = a; + b; with II L aia; 111/2 S; c and II L bib; 1 1 / 2 S; c. This yields (with the notation in the proof of Proposition 22.3, using r(b;bi) = r(bib;))
Hence finally c 2': vn(2v2)-1. Thus we have c = By (10.12) and (10.1)' we then obtain (10.17).
IIRn nCn
-+
cf>nllcb::O
vn·
To check (10.18), note that by (10.21) and (10.22) we have (a fortiori)
whence (10.18).
222
Introduction to Operator Space Theory
To check (10.19), note that (trivially) can write by (10.14)
II~n ~
min(£2')llcb = 1. Then we
n ~ II~~ ~ ~nllcb ::; II~~ ~ min(£2')llcb II min(£2') ~ ~nllcb
=11 min(£2') ~ ~nllcb,
whence (by (10.4»
II min(£2') ~ ~nllcb ~ n.
Thus we obtain (10.19).
To check (10.20), note that (trivially) II max(£2') (10.13) and (10.15»
~ ~nllcb =
1. Then (by
whence dcb(~n, max(£2'» ::; C..;n with C independant of n. In the converse direction, the results of [JP] show that max(£2» is not exact, contrary to~. The corresponding estimate from [JP] (see Exercise 19.2 below) is d s dmax(£2')) ~ 4- 1 ..;n, whence dcb(~n' max(£2')) ~ 4- 1 ..;n. Thus we obtain (10.20). • It is now easy to complete these results. First we can compare Rand C to the other spaces in the list (recall dcb(R n , Cn) = n): Theorem 10.5. The following estimates hold:
dcb(Rn ,min(£2'»
= dcb(Cn ,min(£2'» = dcb(Rn ,max(£2'» = dcb(Cn, max(£2'»
= ..;n.
Proof. By (1.5) we have IIRn ~ Cnllcb = IICn ~ Rnllcb = ..;n, from which it immediately follows that IIRn ~ Rn n Cn Ilcb = IICn ~ Rn n C n Ilcb = ..;n. On the other hand, we can verify easily (using (1.11) and (3.1»
II min(£2') ~ Rnllcb =
II min(£2') ~ Cnllcb =
¥n.
(10.23)
It is then an easy task to prove the missing estimates using the duality. Let us now compare Rn n Cn and Rn + Cn to the remaining spaces.
•
10. Comparisons
223
Theorem 10.6. The following estimates llOld:
Proof. From (10.23) we deduce that
II min(f2) --+ Rn n Cnllcb A fortiori, we have
IIRn
II min(f~)
+ Cn --+ max(f~)llcb = n ~II min(f~) --+
--+
--+
Rn
+ Cnllcb
Vii·
= ~
vn, whence, by duality,
vn. Therefore, we can write by (10.10)
max(£~)llcb ~
II min(f2)
--+
Rn
+ CnllcbllRn + Cn
max(f2 )llcb ~
Viii I min(f2) --+ Rn + Cnllcb,
and we obtain
II min(£2) --+
Rn
+ Cnllcb ~ Vii·
Since we have trivially IIRnnCn --+ min(f~)llcb = IIRn+Cn --+ min(e~)llcb = 1, we obtain the estimates on the first line and those on the second one follow ~~ili~
•
We can now state, as a recapitulation: Theorem 10.7. Tile operator spaces
OH,
~,
R,
C,
R n C,
R
are mutually completely nonisomorphic.
+ C,
min(f 2),
max(£2)
PART II OPERATOR SPACES AND C*-TENSOR PRODUCTS
Chapter 11. CO-Norms on Tensor Products. Decomposable Maps. Nuclearity Let AI, A2 be two C· -algebras. Their algebraic tensor product Al ® A2 is an involutive algebra for the natural operations defined by
and
Then a norm II lion Al ® A2 is called a C·-norm if it satisfies Ilxll = Ilx·ll,
IIxYIl :::; Ilxllllyll
and
Ilx·xll
= IIxl1 2
for any x, y in Al ® A 2. It can be shown that we then automatically have (11.0) This subject was initiated in the 1950s by Turumaru in Japan. Later work by Takesaki [Tal] and Guichardet [Gu1] lead to the following striking result. Theorem 11.1. There is a minimal CO-norm II Ilmin and a maximal one II Ilmax, so that any CO-norm II ·11 on Al ® A2 must satisfy Ilxlimin :::; Ilxll :::; IIxll max . ~Ve denote by Al ®min A2 (resp. Al ®max A 2) tlle completion of Al 129 A2 for the norm II Ilmin (resp. II Ilmax).
The maximal C* -norm is easy to describe. We simply write
where the supremum runs over all possible Hilbert spaces H and all possible *-homomorphisms 7r: AIQ9A2 --+ B(H). It is eaRY to see that, for any such 7r. there is a pair of (necessarily contractive) *-homomorphisms 7ri: Ai --+ B(H) (i = 1,2) with commuting ranges such that
Conversely, any such pair 7ri: Ai --+ B(H) (i = 1,2) of *-homomorphismR with commuting ranges determines uniquely a *-homomorphism 7r: Al 129 A2 --+ B(H) by setting 7r(al ® a2) = 7rI(aI)7r2(a2). Thus, we can write for any x = Eal ®a% in Al ®A 2
where the supremum runs over all possible such pairs.
Introduction to Operator Space Theory
228
Then the inequality II II:::; II IImax follows by considering the GelfandNeumark embedding of the completion of (AI ® A 2 , II II) into B(H) for some H. All this goes back to [Gu1]. The lower bound II IImin:::; II II is due to Takesaki [Tal] and is much more delicate. For a proof, see either [Ta3] or [KaR]. The minimal norm can be described as follows: Embed Al and A2 as C*subalgebras of B(H1) and B(H2)' respectively. Then, for any x = E a}®a~ in Al ® A 2, Ilxllmin coincides with the norm induced by the space B(HI ®2 H 2), that is, we have an embedding (Le. an isometric *-homomorphism) of the completion, denoted by Al ®min A 2, into B(Ht ®2 H 2). In other words, the minimal tensor product of operator spaces, considered in §2.1, when restricted to two C* -algebras coincides with the minimal C*tensor product. Let (B I ,B2 ) be another pair of C*-algebras and consider c.b. maps Ui: Ai --+ Bi (i = 1,2). Then (as emphasized already in §2.1) Ut ®U2 defines a c.b. map from At ®min A2 to Bl ®min B2 with lIut ® u211cb = Il u tllcbll u2l1cb. In sharp contrast, the analogous property does not hold for the max-tensor products. However, it does hold if we moreover assume that Ut and U2 are completely positive, and then (see e.g. [Ta3, p. 218], [Pal, p. 164], or [Wa2, p. 11]) the resulting map Ul ®U2 is also completely positive (on the max-tensor prod uct), and we have
This follows from the next two statements.
Theorem 11.2. Let cp: At ® A2 --+ C be a linear form and let u",: Al be the corresponding linear map. The following are equivalent:
--+
A2
(i) cp extends to a positive linear form in the unit ball of (AI ®max A2)*' (ii) u",: At
--+
.42
is a contractive c.p. map, that is, lIu",1I :::; 1, and
L(U",(Xij), Yij) ;::: 0 "In "Ix i,j
E
Mn(At}+ Vy
E
Mn (A 2)+.
Proof. Assume (i). Recall (11.0). Then clearly Ilu",1I :::; 1. Moreover (by the GNS construction) there are a representation 71': Al '8lmaxA2 --+ B(H) and in the unit ball of H such that cp(.) = (7I'(·)e, e). We may assume that 71' = 71't . 71'2 as above. Let X,Y be as in (ii). Let Z = yl/2, so that Yij = Lk ZkiZkj (note Zik = zki)' We claim that the matrix (7I'1(Xij)7I'2(Yij)) is positive. Indeed, for each fixed k the matrix
e
11. C*-Norms on Tensor Products. Decomposable ~l\faps. Nuc1earity
229
is positive and, since 7r1, 7r2 have commuting ranges, we have 7r1(Xij)7r2(Yij) = Lk 7r2(Zkd*7rl(Xij)7r2(Zkj)'
This proves our claim. Let [ ~ EEl··· EEl~. Then we have
E
L(U
H E9 ... EEl H (n times) be defined by [ =
= ([ 7r I(Xij) 7r2(Yij)][,[) ~
0,
which shows that (i) =} (ii). Conversely, assume (ii). Consider t = L aj 0 bj in Al 0 A 2· Then t*t = Li,j aiaj 0 bibj, and hence by (ii)
)
~ 0,
which shows that
IImax).
Let tEAl 0 max A2 and () = (UI 0 U2)(t). Since () E (BI 0 max B 2 )+ iff
In the nonunital case, we obtain the same conclusion using approximate units .
•
230
Introduction to Operator Space Theory
Let A, B be C*-algebras. We will denote by CP(A,B) the set of aU completely positive maps u: A --+ B. Recall (see Exercise 11.5 (iii» that, for any c.p. map u: A --+ B, we have IIullcb = Ilull and in the unital case IIuli = Ilu(1)11. Thus, by Corollary 1.8, a unital map u: A --+ B is completely positive iff it is completely contractive. More generally, we denote by D(A, B) the set of all decomposable maps u: A --+ B, that is, maps that can be written as u = Ul - U2 + i(U3 - U4) with Ul, •.. , U4 E CP(A, B). This space can be normed in a fairly straightforward way by setting 4
IIull[d] = inf{L IIUill}, 1
where the infimum runs over all the above possible decompositions of u with Ui E CP(A, B). It is easy to see that D(A, B) equipped with this norm is a Banach space. However, we will use an equivalent norm, the "decomposable norm" IIulidec, which was introduced by Haagerup in [HI]. This norm allows us to suppress various unnecessary numerical factors (such as 2 or 4) that would appear were we to use the norm IIull[d] instead. The decomposable norm Ilulidec is defined as follows. Consider all possible mappings 8 1 ,82 in CP(A,B) such that the map v: A --+ l\12(B) defined by
vex) = (
8 1(X) u(x)
is completely positive. Then we set
where the infimum runs over all possible such mappings. See Chapter 14 for more on D(A, B). We will use the following list of basic results (from [H3, Proposition 1.3 and Theorem 1.6]): VuE D(A,B)
(11.2)
If u and v are as above, and if, say, B C B(H), we can write u(x) = P2v(x)pi, where Pi: H EB H --+ H is the projection to the i-th coordinate:
VuE CP(A,B)
(11.3)
If C is any C*-algebra, u E D(A, B), and v E D(B, C), then vu E D(A, C) and !Ivulldec ~ IIvlidecllulidec, (11.4)
VuE CB(A, B(rt»
lIulicb = lIulidec.
(11.5)
11. C*-Norms on Tensor Products. Decomposable J\-laps. Nuclearity
231
Now let A l , A 2, B l , B2 be arbitrary C*-algebras, and let Ul E D(Al. B l ), U2 E D(A2' B2). We claim that Haagerup's results imply that Ul Q9 U2 extends to a decomposable map, denoted by Ul Q9max U2, from Al Q9max A2 into Bl Q9max B2 satisfying (11.6) Ilul Q9max u211dec S II u llidecll u 211dec. By (11.2) we have a fortiori (11.7) All this extends to tensor products of n-tuples A l , A 2, ... ,An of C*algebras, but the resulting tensor products Al Q9min ... Q9min An and Al Q9max .•. Q9max An are associative, so that we have, for instance, when a = either min or max,
Therefore the theory of multiple products reduces, by iteration, to that of products of pairs. \Ve have obviously a bounded *-homomorphism q: Al Q9maxA2 -+ Al Q9min A 2, which (as all C*-representations) has a closed range; hence Al Q9min A2 is C*-isomorphic to the quotient (Al Q9max A 2)/ker(q). The observation that in general q is not injective is at the basis of the theory of nuclear C* -algebras:
Definition 11.4. A C*-algebra A is called nuclear if, for any C*-algebra B, we have II II min = II II max on A Q9 B or, in short, if A Q9min B = A Q9max B. In tllat case, by Tlleorem 11.1, tll€re is only one C*-norm on A Q9 B. This notion was introduced (under a different name) by Takesaki and was especially investigated by Lance [La1]. For example, if dim(A) < 00, A is nuclear, because AQ9B (there is no need to complete it!) is already a C* -algebra; hence it admits a unique C* -norm. Moreover, all commutative C*-algebras are nuclear and J( = K(£2) is nuclear (see Exercise 11.7), but B(£2) is not nuclear (cf. [Wa1]; we will reprove this later at least twice in Chapter 17 and in Chapter 22). Also, the Cuntz algebra (see §9.4) is nuclear. Other examples or counterexamples can be given among group C* -algebras. For any discrete group G, one can define the full C* -algebra C* (G) and the reduced one C!(G) (see §§9.6 and 9.7 for precise definitions). Then C*(G) (or C! (G)) is nuclear iff G is amenable. So, for instance, if G is the free group on two generators, C*(G) and C!(G) are not nuclear. (Note that for continuous groups the situation is quite different: Connes [Col] proved that, for any separable connected locally compact group G, C*(G) and CnG) are nuclear.)
Introduction to Operator Space Theory
232
Moreover, as already mentioned in Chapter 8, if G ll G 2 are two discrete groups, we have
The basic characterizations of nuclear C* -algebras are incorporated in the next statements.
Theorem 11.5. Let A be a C* -algebra. The following are equivalent. (i) A is nuclear.
(ii) There is a net (Ui) of finite rank completely positive maps on A, converging pointwise to the identity. (We then say that A has the "completely positive approximation property," in short CPAP.) (iii) There is a net of maps (Ui) admitting a contractive completely positive factorization through matrix algebras of the form
A
_ _U....:i'---+l
A
(that is, Ui = WiVi and Vi, Wi are contractive completely positive) and converging pointwise to the identity. (iv) There is a net of maps
(Ui) admitting a completely contractive factorization tllrough matrix algebras of tIle form
A
_ _U....:i_-+l
A
(i.e., Ui = WiVi and Vi, Wi are completely contractive) and converging pointwise to the identity. The equivalence of (i) and (ii) was obtained independently in [Ki3] and [CElIo (i) ¢:} (iii) was obtained in [CE2] and (i) ¢:} (iv) in [Sml]. Moreover, using Connes's deep work [Col], the following striking result was shown in [CE2j. (See §2.5 for a description of the C* -algebra structure of the bidual AU.)
Theorem 11.6. A C*-algebra A is nuclear iffits bidual A** is injective, that is, there is a completely contractive projection P from B(H) onto A**. By Corollary 1.8, P is automatically completely positive. The next notion was introduced in [ELl·
11. C*-Norms on Tensor Products. Decomposable Maps. Nuclearity
233
Definition 11.7. A von Neumann algebra At is called semi-discrete if the identity map on the predual 11,1* is the limit of a net of finite rank contractive completely positive maps in the pointwise topology (l1ere we say that a map on AI* is c.p. if its adjoint is c.p. on At). Equivalently, this means tllat tIl ere is a net of finite rank contractive completely positi\'e normal maps Uj: !II -4 !II such that, for all x in AI, Ui(X) tends to x in the a(M, At*)-topology.
Remark. In [CEl, p. 75], it is proved that if At is semi-discrete, there exists a net of maps (Uj) admitting a contractive completely positive normal factorization through matrix algebras of the form
At
__ U..:...i- - t )
!If
(i.e., Ui = W(l-'i and Vi, Wi are contractive, completely positive, and normal) such that, for all x in !If, Ui(X) tends to x in the a (At, !If*)-topology. Effros and Lance [EL] proved that semi-discreteness implies injectivity. Relying heavily on Connes's work, Choi and Effros ([CE4]) proved the converse on a separable Hilbert space (see [Wa5] for a simpler proof covering the nonseparable case). The key to these developments is Connes's proof that aU injective factors on a separable Hilbert space are hyperfinite. A von Neumann algebra is called hyperfinite if there is a directed net of finite-dimensional subalgebras Ala: C !If the union of which is weak-* dense in M. It is not too hard to show (bllt, beware, it is easy to give a false proof of this; see Remark 11.13 later) that hyperfinite implies injective (see [T02] for a proof based on the fact that AI is injective iff its commutant is). As a consequence of Connes's work on factors, one finally gets the converse for von Neumann algebras (on a separable Hilbert space). Thus, on a separable Hilbert space, injectivity, semi-discreteness, and hyperfiniteness are all equivalent properties for von Neumann algebras. (See Elliott's papers [EIl-2] for the nonseparable case, where hyperfiniteness has to be replaced by a different notion of "approximately finite-dimensionaL") See [Tor] for an exposition. Simpler proofs of Connes's most difficult results were obtained more recently in [H5] and [Pol-2]. It is rather easy to show that nuciearity passes to ideals (see Corollary 11. 9 later). It is known (and nontrivial) that nuciearity passes to quotients (cf. [CE2]), but not to subspaces ([Ch2, Blal]). Also (d. [CE2]) lluciearity is preserved under "extensions." This means that if I c A is an ideal in a C* -algebra, and if both I and AI I are nuclear, then A is nuclear. These assertions are easy consequences of Theorem 11.6 and the fact that, for any ideal I C A, we have a C* -isomorphism A** ~ (AII)** EB [**.
234
Introduction to Operator Space Theory
Indeed, the latter isomorphism shows that A ** is injective iff both (AII)** and 1** are injective. A major difference between the two "extreme" tensor products is that, while the maximal one is projective (see Exercise 11.2), the minimal one is not. On the other hand, the minimal tensor product is injective, but the maximal one is not. :More precisely, it can happen that, 'when B, At, A2 . are C*-algebras with B c AJ (as a C*-subalgebra), the norm induced by Al ®max A2 on B ® A2 does not coincide with the maximal norm on B ® A 2. Of course, this difficulty disappears when either B or A2 is nuclear, since we then have a unique CoO-norm on B ® A 2 • Since this "defect" is a key to the subsequent theory, we now review, for emphasis, some elementary sufficient conditions for the injectivity of the maximal tensor product.
Proposition 11.8. Let AbA2 be two C*-algebras, let B C Al be a C*subalgebra, and let E c B be a subspace. Eadl of tIle following conditions is sufficient to ensure that
(i) There is a contractive completely positive (in sllOrt c.p.) projection P: Al --+ B from Al onto B. (ii) There is a net of contractive c.p. maps Ti : Al --+ B sudl tllat Ti(x) --+ x for all x in E. (iii) There is anet of decomposable maps Ti : Al --+ B SUell that IITilldec --+ 1 and Ti(X) --+ x for all x in E. Proof. This is an immediate consequence of (11.1) and (11.7). (Recall that, for any c.p. map u between C*-algebras, we have lIulicb = Ilull = Ilulidec.) Corollary 11.9. Let AI, A2 be C* -algebras. For any (closed, two-sided, selfadjoint) ideal I c A we have an isometric embedding
I ®max A2 C Al ®max A 2· Proof. As is well known, I satisfies (ii) in the preceding statement (here we take B = E = I). Indeed, I has an approximate unit (cf. [Ta3] or Lemma 2.4.4), namely, there is a net (ai) in I such that Ilaill ::; 1 with aiX --+ x and xai --+ x for any x in I. Thus, if we let Ti(x) = aixai, we find (ii0· • Remark. Another useful situation of the same kind as in the preceding two statements is provided by the canonical inclusion Al C Ai*. For any C*-algebra A 2, we have an isometric inclusion Al ®max A2 C Ai'" ®max A2· For the proof, see the solution to Exercise 11.6. It will be useful to record here the following fact.
(11.8)
11. C*-Norms on Tensor Products. Decomposable l\Iaps. Nuclearity
235
Proposition 11.10. Let u: Al --> A2 be a decomposable linear map between two C* -algebras, and let B be anotller C* -algebra. If f B Q9 u: B Q9min A I --> B Q9max A2 is bounded, then
In particula,r, this bolds for any finite rank map u. lHoreover, for any finite rank c.p. map u, we have
Proof.
B
Q9max
Let us denote by
fB Q9max u
the mapping
fB Q9 u:
B
Q9lllax
Al
-->
A 2 . By (11.7), we have
Observe that the natural morphism q: B Q9lllax Al --> B Q9min Al is surjective (since it has dense range); hence it defines an isometric isomorphism from B Q9max AI/ ker(q) to B Q9lllin AI. Let u = fB Q9 u: B Q9lllin Al --> B Q9max A 2 . Since is assumed bounded, we must have q = fB Q9max 11; hence (IB Q9max u)(ker(q)) C {O}, and hence fB Q9max u defines a mapping of norm ::s: Ilulidec from B Q9lllax AI/ker(q) to B Q9lllax A 2 , which yields the announced result.
u
u
u
is bounded. Now, if u has finite rank, it is rather easy to verify that Indeed, it suffices to verify this for a map u of rank 1, but then u factors through
u
B
and hence Since
Q9min
If"' A I 1/J0VB --> Q9min =B IL,.
Q9lllax
1f"' 1/J0 W IL,.
-->
B
Q9lllax
A 2,
u is bounded.
Ilulidec = Ilull
when u is c.p., we obtain the last assertion.
•
Remark 11.11. Let C be a nuclear C*-algebra, and let A c C be a C*subalgebra. Consider a C*-embedding C C B(H). Then, for any C*-algebra B, we have an isometric embedding A
Q9lllin
B
-->
B(H)
Q9max
B.
(11.9)
Indeed, C Q9min B --> C Q9max B is isometric since C is nuclear; moreover, A Q9min B --> C ®min Band C ®max B --> B(H) ®max B are both clearly contractive. Thus (11.9) is contractive, and hence isometric. We will see in Chapter 17 that this property chamcterizes the (separable) C* -algebras that can be embedded in a nuclear one.
236
Introduction to Operator Space Theory
Remark 11.12. The C*- or von Neumann algebra setting leads to all sorts of implications that are surprising to a Banach space theorist. For instance, the fact that A nuclear (which, by Theorem 11.5, is a kind of approximation property for A) implies A** semi-discrete (which, by definition, is one for A*) is in sharp contrast with the fact that there exist a Banach space X (separable and with separable dual) with Grothendieck's Banach space approximation property but with dual X* failing it; see [LT1, p. 34J. Concerning injectivity, Connes's results are formally analogous to part of the theory of .coo-spaces (see [LiRJ). A Banach space X is called .coo if there is a net of finite-dimensional subspaces Xo. C X uniformly isomorphic to e~ spaces of the same dimension with union dense in X. In other words, X can be written as a norm dense directed unio!l of subspaces uniformly isomorphic to injective and finite-dimensional spaces. By [LiRJ, this holds iff the bidual X** is isomorphic to an injective Banach space, that is, iff X** is isomorphic to a complemented subspace of an Loo-space. Thus, this notion appears as the Banach space analog of nuclearity. Note that the difficulty in Connes's proof that injective implies hyperfinite, when compared to the analogous fact in [LiRJ, is that, starting from an injective infinite-dimensional object, one must construct a (suitably dense) directed net of finite-dimensional and injective *subalgebras; not surprisingly, it is much easier to construct (finite-dimensional injective) subspaces than *-subalgebras.
Remark 11.13. In yet another direction, we wish to point out that the proof that hyperfiniteness implies injectivity cannot be too simple in view of the following example (kindly pointed out to me by W. B. Johnson): There exists a constant C and a dual Banach space X*, with a weak-* closed subspace M C X* and a directed net of weak-* closed subspaces Mo. C M onto which there is a projection Po.: X* -+ Mo. with II Po. II :-:; C and with UMo. weak-* dense in 11I but such that there is no bounded projection from X* onto 1\1. As we will see in the next chapter, the properties of nuclear C* -algebras are valid more generally for a certain class of mappings between C* -algebras. It would have been natural to call these mappings "nuclear," but the risk of confusion with Grothendieck's notion of nuclearity would be too great. Still, in various discussions, the corresponding term is lacking, so (rather daringly!) we propose the following terminology. Definition 11.14. Let u: X -+ A be a mapping from an operator space X to a C* -algebra A. Let B be another C* -algebra. We will say that u is B-maximizing if IB ® u defines a bounded map from B ®min X to B ®max A. lVe will say that u is maximizing if this happens for all C* -algebras B. With this terminology, by definition A is nuclear iff its identity map is maximizing. More generally, we will see later (see Theorem 17.11) that A is "exact" iff the inclusion map A C B(H) is maximizing. In the next chapter
11. C*-Norms on Tensor Products. Decomposable Maps. Nllclearity
237
(see Corollary 12.6) we show that the maximizing maps u: X -+ A are characterized by a certain approximation property by (finite rank) maps factoring through matrix algebras. Note that the analogous notion of "maximizing" map makes perfectly good sense in the Banach space category: A map u: X -+ Y between Banach spaces v
could be called maximizing if, for any Banach space Z, the map I z 0 u: Z 0 A
X -+ Z 0 Y is bounded. It is then an easy (and instructive) exercise to show that this holds iff u is the pointwise limit of a net of finite rank maps Un: X -+ Y with uniformly bounded nuclear norms. Equivalently, u is an integral operator in Grothendieck's sense; that is, if we view u as acting into Y**, then u can be factorized as
where J: L oo (J.1)
-+
LI (J.1) is the inclusion map relative to a probability space
(O,J.1). Exercises Exercise 11.1. Let A, B be C* -algebras, and let I sided) ideal. Prove that the sequence
{O}
-+
I0 ma x A
-+
B 0 max A
-+
c B
(BII) 0 max A
be a (closed two-
-+
{O}
is exact. More precisely, prove that any x in (BII) 0 A with lifting in B 0 A such that IIxlirnax < 1.
x
Ilxli max <
1 admits a
Exercise 11.2. Let AI, A2 be C* -algebras, and let I C Al be an ideal. Prove the following (isometric) identity
Exercise 11.3. Let AI, A2 be two C* -algebras, and let II II be a C* -norm on Al 0 A 2. The goal is to show that lIal 0 a211 :::; lIatl! lIa211 for all ai E Ai (i = 1,2). Prove that 0 :::; al :::; bl and a2 2: 0 implies lIal 0 Q'211 :::; IIbl 0 a211· Then, using aix*xal :::; IIxll2aial, prove that IIxal0a211 :::; IIxlillal0a211, and deduce IIxal0ya211:::; IIxlillylillal0a211 for x E AI,y E A 2. Finally, choosing x = ai and y = a2' prove that lIal 0 a211 :::; lIalli lIa211. Exercise 11.4. Let AI, A2 be two C*-algebras, and let 11": Al 0 A2 -+ B(H) be a *-homomorphism. The goal is to show that there are representations 1I"i: Ai -+ B(H) with commuting ranges such that
Introduction to Operator Space Theory
238
One can proceed as follows: (i) Let K = 7r(Al 0 A 2 )H. Show that 7r decomposes as 7r(·)IK EB 0 with respect to H = K EB K ~. Thus the problem reduces to the case K = H. (ii) Let {xa} (resp. {Y/3}) be an approximate unit in the unit ball of Al
(resp. A 2), so that XaX --+ x (resp. Y/3Y --+ y) for any x in Al (resp. yin A 2 ). Assuming K = H and using the preceding exercise, prove that, for any ai in Ai, the nets 7r(al 0 Y/3) and 7r(xa 0 U2) converge in the strong operator topology of B(H). We denote by 7rl (al) and 7r2(a2) the respective limits. Show that these limits do not depend on the choice of the approximate units. (iii) Show that 7rl, 7r2 satisfy the announced properties.
Exercise 11.5. (i) Let p, q, a be operators on H. Assume p, q 2:
p ( a*
iff l(ax,y)1 2 ::; (py,y)(qx,x) V'x,y
o.
Show that
a)q 2:0 E
H.
(ii) In particular,
(iii) Let u: A
B be a c.p. map between C* -algebras. If A is unital, show that Ilull = Ilu(l)lI. If aa 2: 0 is a net in A such that a~2aa~j'2 --+ a for any a in A, show that Ilull ::; sUPa Ilu(aa)ll· Conclude that Ilulicb = lIull· --+
In any case note that Ilull = sup{llu(a)111 a 2: 0, lIall ::; l}.
Exercise 11.6. (i) For any C*-algebras AI, A2 we have an isometric embedding
(ii) More generally, we have an isometric embedding
(iii) The canonical inclusion j: Ai 0
(contractive) representation
max
A2
--+
(Ai 0 max A2)** extends to a
11. C*-Norms on Tensor Prodllcts. Decomposable Maps. Nllclearity
239
Exercise 11.7.
(i) Let A be a C*-algebra. Consider a directed net (Ac,,) of C*-subalgebras with dense union. Show that if all the Ac:t are nuclear, then A is nuclear. (ii) Deduce from this that Loo(O, A, p,) is nuclear (for any measure space (0, A, 11,)) or, equivalently, that any commutative von Neumann algebra is nuclear.
(iii) Show that any commutative C* -algebra is nuclear. Exercise 11.S. Show that, if II IImin = II on C~(G) 0 C*(G), then G is amenable.
Hint: Use Exercise 8.4.
Ilmax either on C~(G) 0
C~(G)
or
Chapter 12. Nuclearity and Approximation Properties The aim of this chapter is to give a reasonably direct proof of the fact that a C* -algebra A is nuclear iff it has the approximation properties mentioned in Theorem 11.5. Actually, we will show the following more general statement: Let A l , A2 be two C*-algebras, and let u: At ---+ A2 be any linear map. Fix a constant C > O. Then the following assertions are equivalent: (i) For any C*-algebra B,
(ii) There is a net Ui: Al ---+ A2 of finite rank maps tending pointwise to and such that SUPi Iluilidec ::; C.
U
This will be proved in Corollary 12.6. The proof will require a dual description of the maximal C*-norm (see Theorem 12.1) for which we need to introduce a specific notation.
Notation. Let E, F be operator spaces, and let a: E B(H) be linear mappings. We denote by a· rr: E ® F
---+
---+
B(H) and rr: F---+
B(H)
the linear mapping defined by
The main ingredient for this chapter is the following.
Theorem 12.1. Let A c B('Jt) be a closed unital subalgebra, and let E be an operator space. Consider an element y in E ® A. Let ~(y) =
sup{lIa· rr(y)IIB(H)},
where the supremum runs over all Hilbert spaces H and all pairs (a, rr), where rr: A ---+ B(H) is a completely contractive unital homomorphism and a: E---+ rr(A)' is a complete contraction. On the other hand, let (12.1)
where the infimum runs over all possible n and all possible representations of y of the form n
Y
=
L ij=l
Xij ®aibj.
(12.2)
12. Nuc1earity and Approximation Properties Then
~(y)
241
= c5(y).
Remark. Let q: A0E0A -+ E0A be the linear mapping taking a}0x0a2 to x 0 ala2' Then, by definition of the norm on A 0h E 0h A, it is easy to check that
c5(y) = inf{IIYIIA®I.E®I.A
lyE A 0
E 0 A, q(Y) = V}·
Note in particular that this shows that 15 is a norm. Equivalently, if we denote by E08A the completion of E0A for the norm 15, then the mapping q extends to a norm 1 mapping from A 0h E 0h A onto E 08 A, and the resulting mapping (which we still denote - abusively - by q) is a metric surjection from A 0h E 0h A onto E 08 A, that is, it induces an isometric isomorphism from (A 0h E 0h A)/ker(q) onto E 08 A. The preceding remark (which I had overlooked) is due to C. Le l\Ierdy. It allows us to shorten the proof of Theorem 12.1.
Proof of Theorem 12.1. We first show the easy inequality ~(y) ::; c5(y). We may assume by homogeneity that c5(y) = 1. Let (0-,7l') be as in the definition of ~(y). Then we have, assuming (12.2),
and hence
whence
~(y)
::; c5(y).
To show the converse, since c5 and it suffices to show that
~
~*
So let e: E ® A such that
-+
are norms (see the preceding remark),
::; 15*.
C be a linear form such that c5*(e) :S 1, or, equivalently, '
E
E®A
le(y)1 ::; c5(y).
Introduction to Operator Space Theory
242
e
By the preceding remark, we can associate to an element (A ®h E ®h A)* determined by the identity
f in the unit ball of
f:
Now, by Corollary 5.4 and Remark 5.5 applied to A ®h E ®h A --+ C, we can find two representations 7l"1: B(H) --+ B(H1) and 7l"2: B(H) --+ B(H2) together with a completely contractive map v: E --+ B(H2' H 1) and unit vectors 6 E H1 and 6 E H2 such that, for any a, b in A and any x in E, we have e(x ® ab) = f(a ® x ® b) = (7l"1(a)v(x)7l"2(b)6,e1). (12.3) If we now replace
7l"2(-)
by 7l"2(-)17I"2(A)~2 and
7l"1(-)
by the homomorphism a--+
(7l"1(a*)17I"dA')~)*' we may as well assume that 7l"2(A)6 and {7l"2(a*)6 I a E A} are dense in H2 and H 1, respectively.
Then, writing (ac)b = a(cb) into (12.3) and using the density just mentioned, we find for all c in A "IxEE
Let
7l":
A
--+
(12.4)
B(H1 EEl H 2) and v: E
--+
7l"(a)
= (7l"l(a)
v(x)
=
o
B(H1 EEl H 2) be defined by
0)
7l"2(a)
and
(~ V~»).
Then (12.4) implies v(x)7l"(c) = 7l"(c)v(x) for x E E, c E A, and 7l" is a completely contractive unital homomorphism on A. Finally, letting = (~~) and
"'2
"'1 = (~), we find e(X ® ab)
= (7l"(a)v(x)7l"(b)"'2' "'1).
Hence (taking for a the unit of A) for all y in E ® A
therefore 1~(y)1
::;
IIv· 7l"(y) II
::; ~(y).
This completes the proof that ~ * ::; 8* and hence 8 ::; ~.
•
We will mostly restrict attention here to the case when A is a C* -algebra. We will return to the non-self-adjoint case in Chapter 25.
12. Nuclearity and Approximation Properties
243
Proposition 12.2. Consider an operator space E. lYe view E as embedded into C*(E) (as defined in Chapter 8). Then, if A is a unital C*-algebra. we llave (12.5) 8(y) = iiyllc*(E)0m"x A' Proof. Note that any (completely) cont.ractive unital homomorphism 7r: A ----> B( H) is necessarily a *-homomorphism, or, equivalently, a (C* -algebraic) representation. Then 7r(A)' is also a C* -algebra, and any completely contractive map a: E ----> 7r(A)' extends to a representation C*(E) --+ 7r(A)'. Hence we dearly have Vy E E @ A
a:
~(y) :S sup lIa· 7r(y)ii :S lIyllc*(E)0m"xA'
•
The converse is dear.
Corollary 12.3. In the situation of Theorem 12.1, assume that A is a C*a.lgebra and let F be another operator space and B anotller C* -algebra. C011sider '111 E CB(E, F) a11d U2 E D(A, B). Then, for all y ill E@ A, we llave (12.6)
Proof. Assume iiuliicb = 1. Note that 1/,1: E --+ F extends to a C*representation from C* (E) to C* (F). Then (12.6) is an immediate conse• quence of (11.6) and (12.5). In the particular ease when E
=
AI;', Theorem 12.1 becomes:
Corollary 12.4. In Theorem 12.1 , assume A is a C* -algebra. and let E = 111;' for some N ~ 1. Then, for all y in E @ A, witll associated linear map iJ: AIN ----> A, we have (12.7) 8(y) = iiiJIID{lHN,A)' Proof. Let t E M;' @ MN be associated to the identity of lIlN . Let (eij) be biorthogonal to the standard basis (eij) in M N, so that t = L eij @ eij. Then we can write and hence
Hence, by Corollary 12.3, we have 8(y) = t5«J @ Y)(t)) :S lIiJiidec8(t) :S iiiJiidec'
244
Introduction to Operator Space Theory
On the other hand, if we have (12.2), then let v: E* -+ Mn be the map defined by v(e) = (e(Xij» (e E E*) and let w: Mn -+ A be defined by w(eij) = aibj. Then it is fairly easy to verify (see Exercise 12.2) that (12.8)
whence, if we let
fJ =
wv by (11.4) and (11.5),
Taking the infimum over all representations of the form (12.2), we obtain
IlfJlldec ::; 8(y).
•
Corollary 12.5. In the situation of Theorem 12.1, assume A is a C* -algebra. Consider an element y in E 0 A. Let fJ: E* -+ A be tIle associated linear map. Then
where the infimum runs over all possible factorizations of fJ of the form
with v weak-* continuous. lVloreover, if E = F* for some operator space F, let y: F -+ A be the restriction of fJ to F c F**. Tllen again
where the infimum runs over all possible factorizations of y of the form
Proof. First observe that the weak-* continuity of v implies that we have = (v*)· for some map v.: Af~ -+ E. Then, if fJ = wv with v, w as above, we can write y = (v* o IA)((J) , v
where (J E M~ 0 A is associated to w. Hence we have by (12.6) and (12.7)
whence 8(y) ::; inf{IIvllcbllwlldec}. For the converse inequality, we argue as in the preceding proof.
12. Nudearity and Approximation Properties
In the case E the reader.
245
= F*, the argument is similar, and we leave the details to •
Corollary 12.6. Let oX be a positive constant. Consider two C* -algebras Al and A2 and an operator subspace X c AI. Let u: X --+ A2 be a linear mapping. The following assertions are equivalent.
(i) For any C*-algebra B, IB®u detines a bounded linear map from B®min X to B ®max A2 witll norm ~ oX. (ii) Same as (i) with B = C* (E*) for all tinite-dimensional operator subspaces E CAl. (iii) For any tinite-dimensional subspace E C X, the restriction UIE admits, for any c > 0, a factorization of the form E~JI.[n~A2 with 1IVllcbllwildec ~ oX + c. (iv) There is a net of tinite rank maps U a : X --+ A2 admitting factorizations through matrix algebras of the form
tlo
X
-----+
with that U a = Wa Va converges pointwise to u. (v) There is a net U a : Al --+ A2 of tinite rank maps that tends pointwise to U when restricted to X. SUell
Witll
sup Ilualidec
~
A
Proof. (i):::} (ii) is trivial. Assume (ii). Let E C X be an arbitrary finite-dimensional subspace, and let iE E E* ® X be the tensor associated to the inclusion of E into X. Let B = C*(E*). By (ii) we have II(1B ® U)(iE)IIB0max A 2 ~ AlliEIIB0min X ' But by the injectivity of the min-norm. we have IIiEIIB0mlnX = IliEIIE'0mlnX = IliEllcB(E,X) = 1.
Hence we have II(1B ® U)(iE)IIB0m"x A 2 ~ A. By (12.5) and Corollary 12.5, this implies that, for any c > 0, there is a factorization of UIE of the following form:
V/ E
","w
246
Introduction to Operator Space Theory
with 1IVllcb ::; 1and IIwlldec ::; l+c. This shows (ii) =? (iii). Assume (iii). By the extension property of l\ln, we can extend V to a mapping v: X ---+ l\ln with IIvllcb ::; 1. Thus, if we take for index set I the set of all finite-dimensional subspaces E C X (directed by inclusion), we obtain nets v",: X ---+ ]l.!n" and W",: Mn o ---+ A2 such that IIw",v",(x) - u(x)1I ---+ 0 for all x in X and such that (after a suitable renormalization) sup IIv",lIcb ::; 1, sup IIw",lIdec ::; 1. This completes the proof that (iii) =? (iv). We may clearly assume (by Corollary 1.7) that v", is extended to At with the same c.b. norm; thus, recalling (11.4) and (11.5), (iv) =? (v) is immediate. Finally, the proof that (v) =? (i) is an immediate consequence of Proposition 11.10 in the preceding • chapter. Using an earlier version of the present work, :Marius Junge and C. Le Merdy PLM] obtained the following striking result. Theorem 12.7. ([JLMJ) Let u: B ---+ A be a finite rank map between two C* -algebras. Then, for any c > 0, there is an integer n alld a factorization u = vw of the form with (IIvllcbllwllcb ::;) IIvllcbllwlldec ::; lIulldec(1 + c). Therefore, if y E B* ® A is tIle tensor associated to u: B ---+ A, we have
lIulidec = 8(y). Remark. The following question seems interesting: Fix c > 0 and let k be the rank of u. Can we obtain the above factorization with n ::; I(k, c) for some function I? In other words, can we control n by a function 1 depending only on k and c? The proof will use an original application of Kaplansky's density theorem discovered by M. Junge PI], as follows. Lemma 12.8. Let A, B be arbitrary C*-algebras. Then any c.b. map 0': B* ---+ A can be approximated in the point-norm topology by a net of weak*continuous finite rank maps 0'",: B* ---+ A witll lIa",lIcb ::; lIalicb. Proof. Let !v! = A**®B**. By Theorem 2.5.2, CB(B*, A**) can be identified isometrically with M in a natural way. Let t E A**®B** be the tensor associated to 0' composed with the inclusion A ---+ A**. Then, by Kaplansky's density theorem (see any standard book, such as [Dil, Sa, Ta3, Ped, KaR]), there is a net (t",) in A ® B with Ilt",limin ::; IItliM such that t", a(M, M*)tends to t. Let 0'",: B* ---+ A be the finite rank map associated to t",. \Ve have lia",lIcb = IIt",lImin ::; IItliM = Iialicb. Moreover, for anye in B*, a",(e) must a(A**, A*)-tend to a(~"). But, since a",(e) and a(e) both lie in A, this means
12. Nuc1ea.rity and Approximation Properties
247
that ao:(~) tends to a(~) weakly in A. Passing to suitable convex hulls, we obtain a net (0'0:) such that, for any ~, ao:(~) tends to a(~) in norm. •
Remark. The same argument shows that, for any von Neumann algebra R, any c. b. map 0': R* --t A can be approximated by a net of finite rank maps 0'0:: R* --t A with lIao:lleb ::; lIalieb. Proof of Theorem 12.7. Ifu = wv as in Theorem 12.7, then, by (11.5), we have lIulidee ::; IIvlideellwlldee = IIvllebllwlldee, hence by Corollary 12.5,
lIulidee ::; 8(y). We now turn to the converse. We may assume u(x) = L~ ~i(x)ai with ~i E B* and ai E A, or equivalently y = L ~i ® ai' Assume lIulidee ::; 1. By Corollary 12.5 (and the equality 8 = ~) it suffices to show that, for any representation n: A --t B(H) and any completely contractive map 0': B* ---> n(A)', we have By Lemma 12.8, there is a net of finite rank weak* -continuous maps 0'0:: B* 0'. Hence
--->
n(A)' with lIao:lleb ::; 1 tending pointwise to
Let to: E B®n(A)' be the tensor representing 0'0:' By Proposition 11.10, since u has finite rank, we have
lI(u ® J)(to:)IIA®Ulax7r(A)' ::; lIulideellto:llmin = lIulidecllao:llcb ::; lIulld<,c' Since (u ® J)(to:) =
L
ai ® ao:(~i)' this yields
IlL ao:(~;)n(ai)11 ::; IlL ai ® aa(~i)llmax ::; lIulidec hence in the limit
IlL a(~i)n(a;)11 ::; lIulidec.
Thus, by Theorem 12.1, we obtain
~ (L ~i ® ai)
= 8(y) ::; lIulidec,
so we conclude 8(y) = lIulidee. Then, the first assertion follows from the second part of Corollary 12.5 (applied with F = B). • We now return to complete positivity. Recall (see Exercise 11.5(iii)) that a c.p. map u: Ai --t A2 between two C* -algebras is necessarily continuous and that lIuli eb = lIuli = sup{llu(x)1I1 x ~ 0, IIxll ::; I}. Moreover, if Al has a unit, we have simply lIuli = Ilu(l)lI. To derive the known results on the "c.p. approximation property," the following result, which is a simple adaptation of well-known ideas, will be useful.
248
Introduction to Operator Space Theory
Lemma 12.9. Let E c B(H) be a finite-dimensional operator space. Assume moreover that E is an operator system, that is, E is self-adjoint and unital. Let A be a C* -algebra. Consider a unital self-adjoint mapping u: E ----+ A . associated to a tensor t E E* 0 A. Fix c > o. Then, if 8(t) < 1 + c, we can decompose u as u = rp -l/J with rp, l/J c.p. such that Ill/JII ::; c and rp admits for some n a factorization of the form
V/ A
E where V, Ware c.p. maps with
IIVII ::; 1 + c and IIWII ::;
1.
Proof. By the definition of the norm 8 and by Theorem 1.6, we can assume that u = WV, where v: E ----+ Afn and w: Afn ----+ A are as follows: 'v'X
E
E
v(x)
=
V;7r(X)V2'
where 7r: E ----+ B(ii) is the restriction of a representation, VI, V2 are operators in B(£~,ii) with IIVIII = IIV211 < (1 +c)I/2, and W is defined by
with
Let
and a* = (al, ... ,an)
SO
that, in matrix notation, we have
u(X)
= a* . Vi7r(X)V2 . b.
Since u is self-adjoint, we have
hence we have (by "polarization")
u(X) = rp(x) -l/J(x)
12. Nuc1earity and Approximation Properties
249
with and
'Ijl(X) Clearly cp and
=
'1/) are c.p.
~[(via - 'V2b)*1l'(x)(vla - v2b)].
and
1 = u(l) = cp(l) - 'I/!(I); hence cp(l) ~ 1 and Ilcp(I)1I :<:::; IIvla~v2bll2 :<:::; (1 + c). This implies that Ilcp(1)111 :<:::; c, and hence we obtain (see Exercise 11.5 (iii)) II~)II = 11'1/)(1)11 :<:::; c. It remains to show that cp admits the announced factorization. Let (again in matrix notation)
V(x)
=
(1/2) (
Vi 1l'(X)VI
Vi1l'(X)V2)
v:i1l'(X)VI
V:i1l'(X)V2
=
(1/2)
(
v;*) 1l'(X)(VI,V2). v2
Clearly V is c.p. from E to Af2n and
IWII
:<:::;
!(ll vII1 2+ Ilv2112)
IVIoreover, if we define IV: Af2n
---t
:<:::;
1 + c.
A by
W(t) = !(a*,b*)tG),
then lV is c.p. with
IIWII
:<:::;
1 and we have
cp(X)
=
W(V(x)).
•
We now recover a result due to Choi and Effros and Kirchberg [CEl, Ki3] as follows. (This is the analog of Corollary 12.6 for c.p. maps.) Corollary 12.10. Let u: Al
---t A2 be a completely positive and unital linear mapping between two unital C* -algebras. The following assertions are equivalent. (i) For a.ny C* -algebra B, In 0 u defines a completely positi~'e linear map from B 0 m in Al to B 0 max A2 with norm = 1. (ii) Same as (i) with B = C*(E*) for all finite-dimensional operator subspaces E CAl. (iii) Tlwre is a net of finite Tank maps (un) admitting factorizations through matrix algebras of the form
250
Introduction to Operator Space Theory
with c.p. maps 1'0 and Wo satisfying Ilvo IllIwo II ~ 1 such that U o converges pointwise to u. (iv) There is a net u o : Al ---4 A2 offinite rank c.p. maps that tends pointwise to u. WoVo
Proof. (i)::::} (ii) is trivial. Assume (ii). Let E be a finite-dimensional operator system inside AI. Let B = C*(E*). Let y E E* ®AI be the element associated to the inclusion iE map of E into At. so that IIYlimin = IliEllcb = 1. By (ii) we have II(IB ®u)(Y)llmax ~ 1. Note that the linear map from E to A2 associated to the tensor (IB ® u)(y) is nothing but UIE. Hence, by (12.5) and by Lemma 12.9, for any e > 0 we can write UIE = cp - 'I/J with cp = WV as in Lemma 12.9. By the extension property of c.p. maps we may as well assume that V is a c.p. mapping of norm ~ 1 + e from Al to Mn. Then, using the net (directed by inclusion) formed by the finite-dimensional operator systems E C Al and letting e ---4 0, we obtain (iii) (after a suitable renormalization of V). Then (iii) ::::} (iv) is trivial. Finally assume (iv). Note that since Iluoll = Iluo(1)11 and Iluo (1)1I ---4 lIu(1)1I = 1, we have "automatically" Iluoll ---4 1. By Proposition 11.10, we have for any C* -algebra B
IllB ® uollB0tnlnAl-B0tnaxA2 hence, since
Uo
---4
~ lIuoll;
u pointwise,
•
This shows that (iv) ::::} (i).
Remark 12.11. In the preceding situation, every unital c.p. map of finite rank satisfies the equivalent four properties in Corollary 12.10. Corollary 12.12. A unital C*-algebra A is nuclear iff there is a net of finite rank maps of the form A~1\fno ~A, where vo, Wo are c.p. maps with IIvoll ~ 1, Ilwoll ~ 1, which tends pointwise to the identity. Proof. Apply Corollary 12.10 with u
= lA.
•
Remark 12.13. A C* -algebra A is nuclear iff its unitization is nuclear. This remark allows us to extend the preceding results to the nonunital case. (Indeed, A is an ideal in its unitization A, so that AIA ~ Co By Corollary 11.9, an ideal in a nuclear C*-algebra is nuclear. Thus A nuclear implies A nuclear. The converse is easy and left as an exercise for the reader.) Recently, A. Sinclair and R. Smith [SS3j have obtained a characterization of injective von Neumann algebras by decomposability properties of multilinear maps with values in the algebra. In the case of nuclear C* -algebras, the property they consider has the following analog.
12. Nuclearity and Approximation Properties
251
Theorem 12.14. Let A be a nuclear C* -algebra. Let E}, E 2, E3 be operator spaces. Then the product map of A defines a complete metric surjection from (El ¢SIminA) ¢SIhE2¢S1h (E3¢S1minA) into (El ¢SIhE2¢S1hE3)¢SIminA. In particular, if E2 = C, we have a complete metric surjection from (El ¢SIminA) ¢SIh(E3 ¢S1min A)
onto (El
¢SIh
E 3 ) ¢SImin A.
Proof. Let E = El ¢SIh E2 ¢SIh E 3. By Theorem 12.1 and the remark after it, if A is nuclear, the mapping q: A ¢SIh E ¢SIh A --+ E ¢SImin A is a complete metric surjection. The result then follows from the fact that q factorizes completely contractively through (El ¢SImin A) ¢SIh E2 ¢SIh (E3 ¢SImin A). Indeed, we have by associativity A¢SIhE¢SIhA = (A¢SIhEt}¢SIhE2¢S1h(E3¢S1"A), and we have canonical complete contractions A ¢SIh El --+ El ¢SImin A and E3 ¢SIh A --+ E3 ¢SImin A. Moreover, the product map defines (cf. Exercise 6.1) a complete contraction from (El ¢SImin A) ¢SIh E2 ¢SIh (E3 ¢SImin A) to (El ¢SIh E2 ¢SIh E 3) ¢SImin A. Thus we obtain the announced result. •
Exercises Exercise 12.1. Let A be a C* -algebra. Fix a, b in A. Let u: A --+ A be defined by u(x) = axb*. Note that u is c.p. when a = b. Then show that for any a, b
Ilulidec ::; Iiall IIbll· Exercise 12.2. Let u: Afn --+ A be a linear mapping into a C* -algebra. Assume that we can write u(eij) = aib; with ai,bj in A. Note that u is c.p. if ai = bi for all i. Then show that
Chapter 13. C*(IFoo) ® B(H) Recently,. E. Kirchberg [Ki2-5] revived the study of pairs of C* -algebras A, B such that there is only one C*-norm on the algebraic tensor product A ® B, or, equivalently, such that A ®min B = A ®max B. He constructed the first example of a nonnuclear C*-algebra such that A ®min AOP = A ®max AOP. He also proved the following striking result [Ki5]. A simpler proof is given in [PU], which we follow rather closely in the following. Theorem 13.14. ([Ki5}) Let F be any free group, C*(F) the (full) C*algebra of F, and H any Hilbert space. Tllen
C*(F)
®min
B(H) = C*(F)
®max
B(H).
This is particularly striking if one recalls that C*(F) and B(H) both are universal objects: Every C* -algebra is a quotient of C* (F) and a subalgebra of B(H) for suitable choices of F and H (see §2.12). More generally, we will prove Theorem 13.2. ([Pll}) Let (Ai)iEI be a family of C* -algebras (resp. unital C* -algebras). Assume that for eac11 i in I
Ai
®min
B(H) = Ai
®max
(13.1)
B(H).
We denote by *iEIAi (resp. by *iEIAi) their free product in the category of C*-algebras (resp. in the category of unital C*-algebras). Then we have (13.2)'
and in the unital case
* Ai) ®min B(H) = ( iEI * Ai) ®max B(H). ( iEI
(13.2)
Corollary 13.3. Let (Gi)iEI be a family of discrete amenable groups, and let G = *iEIGi be their free product. Then
C*(G)
®min
B(H)
Remark. Kirchberg's theorem for F
= C*(G) ®max B(H).
= FI
corresponds to Ai
= C*(Z)
for aU
i in I (in the unital case).
The main idea of our proof of Theorem 13.1 is that, if E is the linear span of 1 and the free unitary generators of C*(F), then it suffices to check that the min- and max-norms coincide on E®B(H). More generally, we will prove
13. C*(lFoo )
Q9
253
B(H)
Theorem 13.4. Let AI, A2 be unital C*-algebras. Let (Ui)iEI (resp. (Vj)jEJ) be a family of unitary operators that generate Al (resp. A2). Let El (resp. E2) be the closed span Of(Ui)iEI (resp. (Vj)jEJ). Assume 1 EEl and 1 E E 2. Then the following assertions are equivalent: (i) TIle inclusion map EI Q9min E2 -+ Al Q9max A2 is completely isometric.
(ii) Al
Q9min
A2 = Al
Q9max
A2.
We will use several elementary facts, as follows. The first one is a wellknown property of unitary dilations.
Lemma 13.5. Let U E B(1t), U E B(H) be lInitaries. Assume there is an isometry S: 1t -+ H be an isometry witll range I< c H, such that U
= S*uS.
Then I< = S(1t) is necessarily invariant mider u and u*, so tl18t U commutes witll PK. Proof. For simplicity we assume 1t c H (S being the inclusion map) and we decompose H as 1t ED 1t.l. Then u can be represented as a matrix
~=
11
b)
(11 cd'
where both 11 and u are unitary. Computing the (l,l)-elltry of u*u and uu*, we find c*c = 0 and bb* = 0; hence b = c = 0, which means that commutes w~~. •
u
The following simple fact is essential in our argument.
Proposition 13.6. Let A, B be two unital C* -algebras. Let (11; )iE/ be a family of unitary elements of A generating A as a unital C* -algebra. (i.e., tIle smallest unital C* -subalgehra of A containing them is A itself). Let E c A be the linear span of (11i)iEI and 1A. Let T: E -+ B be a linear operator such that T(lA) = 1B and taking each 11; to a unita.ry in B. Then, IITlicb ~ 1 suffices to ensure that T extends to a (completely) contractive representation (i.e. *-11011l0morphism) from A to B. Proof. Consider B as embedded in B(1t). Then, it clearly suffices to prove this statement for B = B(1t), which we now assume. By the ArvesollWittstock e~ension theorem (see Corollary 1.7), T extends to a complete contraction T: A -+ B(1t). Since T is assumed unital, T is unital; hence by Corollary 1.8 we can write T(x)
= S*ir(x)S,
254
Introduction to Operator Space Theory
where 7?: A ----+ B(H) is a unital representation (Le. *-homomorphism) and 8: 11 ----+ H is an isometry. Now, for any unitary U in the family (Ui)iEI, we have T(U) = f(U) = 8*7?(U)8. Hence, by Lemma 13.5, since T(U) is unitary by assumption, if K = 8(11), then PK commutes with 7?(U). Now since these operators U generate A, this implies that PK commutes with 7?(A), so that f is actually a *-homomorphism. Thus, f is an extension of T and a (contractive) *-homomorphism. This completes the proof. [Note that, a posteriori, the Arveson-Wittstock completely contractive extension f is unique. In short, the proof reduces to this: The multiplicative domain (see Lemma 14.2 below) of f is a unital C* -algebra (cf. [ChI]) and contains (UdiEI; hence it is equal to A.] • Remark. We will apply Proposition 13.6 in the following particular situation. Let A c A be the (dense) unital *-algebra generated by E. Consider a unital *-homomorphism u: A ----+ B. Then lIulElicb :::; 1 suffices to ensure that U extends to a (completeiy) contractive representation (Le. *-homomorphism) from the whole of A to B. Remark 13.7. In the same situation as in Proposition 13.6, note that, if T is a complete isometry, then f is a faithful representation onto the C* -algebra Bl generated by the range of T. Indeed, by Proposition 13.6 applied to T- 1 , f: A ----+ Bl is left invertible. This can be used to give a very simple proof of the fact due to Choi ([Ch2]) that the full C*-algebra of any free group admits a faithful representation into a direct sum of matrix algebras. By Proposition 13.6, it suffices to check this on the free generators, and this is quite easy (see Exercise 13.1). Proof of Theorem 13.4. The implication (ii) ::::} (i) is trivial, so we prove only the converse. Assume (i). Let E = E 1 i8l min E 2. We view E as a subspace of A = Al i8l m in A 2. By (i), we have an inclusion map T:· El i8l m in E2 ----+ Ali8lmaxA2 with IITllcb :::; 1. By Proposition 13.6, T extends to a (contractive) representation f from Al i8l m in A2 to Al i8l max A 2. Clearly f must preserve the algebraic tensor products Al i8l1 and 1 i8l A 2, and hence also Al i8l A 2· Thus we obtain (ii). • Remark 13.8. Let us denote by El i8l1 + 1i8l E2 the linear subspace spanned by elements of A 1 i8l A2 of the form {ali8l1 + li8l a2}. Then, in the situation of Theorem 13.4, Eli8l1 + 1i8l E2 generates A 1 i8lmin A2, so that it suffices for the conclusion of Theorem 13.4 to assume that the operator space structures induced on EIi8l1 + li8l E2 by the min- and max-norms coincide. Proof of Kirchberg's Theorem 13.1. Let Al = C*(F), A2 = B(H). We may clearly assume dim(H) = 00. For simplicity, this time we denote by (Ui)iEJ the free unitary generators (Ui liE I) of C*(F) augmented of the unit (so J is I plus one more point). We take E2 = B(H) and let EI be the linear span of the unit and (Ui)iEJ·
.
13. C*(lFoo) ® B(H)
255
Consider x E El ® E 2, with jjxjjmin < 1. By Lemma 8.9 we can write x = LiEJ Ui ® Xi with Xi E B(H), (Xi)iEJ finitely supported, admitting a decomposition as Xi = aibi with I L ai a: I < 1, I L b;bi II < 1, ai, bi E B(H). Now, let 11": Al ®lIIax A2 -+ B(H) be any faithful representation. Let 11"1 = 1I"1Al®1
and
11"2 = 1I"Il®A 2 •
We then have
iEJ
=L
11"1 (Ui )1I"2(ai)1I"2(bi ).
iEJ Hence, since
11"1
and
11"2
have commuting ranges, we have
Y=
L
11"2 (ai)1I"1
1I"(:r)
=y
with
(Ui )1I"2(bi ).
iEJ Now by (1.11) we have
Ilyll ,; II~ ~2(a;)~2(a;)'II'/211~ ~2(b')'~2(b.)II'/2 < 1. Hence we conclude that
jjxjjlllax
=
sup jj1l"(x)jj ~ 1. IT
This shows that the min- and max-norms coincide on El ® B(H), but since lIfn(B(H)) ~ B(H) for any n (recall dim(H) = (0), this implies "automatically" that the inclusion
El ®lIIin B(H)
-+
Al ®lIIax B(H)
is completely isometric. In other words, the operator space structures associated to the min- and max-norms coincide. Thus, we conclude by Theorem 13.4 (here E2 = A2). • Lemma 13.9. Let AI, A2 be two C*-algebras (resp. unital) satisfying (13.1). Let Al *A 2 (resp. Al *A 2) be their free product (resp. free product as unital C* -algebras), and let E C Al *A 2 (resp. E C Al * A 2) be tlle linear span in Al *A 2 (resp. Al * A 2) of all elements of the form al a2 with ai E Ai. Tllen tlle min- and max-norms of (AI *A 2) ® B(H) (resp. (AI * A 2) ® B(H)) coincide on E ® B(H). Proof. We may clearly assume dim(H) = 00. Consider X in E ® B(H) with Iixlilllin < 1. By Theorem 5.13, X corresponds to an element y in Al ®A 2 with liyli(Al®hA2)®mlnB(H) < 1. By Corollary 5.9, we have Y = Yl 0 Y2 with Yi E Ai ® B(H)
and
IiYilimin < 1.
256
Introduction to Operator Space Theory
We can write
Yi = Laf 0bf k
with af E Ai, bf E B(H) and
Now consider an isometric representation
1r: (At
* A 2) 0 max B(H) ---+ B(1t),
and let O't. 0'2, and p be its restrictions respectively to At 0 1, A2 0 1, and 10 B(H). We have (since the ranges of p and 0'2 commute)
k,t
=
(~O't (a1)p(b1») (~0'2(a~)p(b~»)
= 1r(Yt)1r(Y2).
Hence we conclude that
Ilxlimax = 111r(x)11
~
~
1I1r(Yt)IIII1r(Y2)11 IIYtllmaxllY2l1max,
and hence by our assumption on At and A2 ~
This shows by homogeneity that
IIYtilminllY211min < 1.
Ilxll max
~
Ilxllmin.
•
Remark 13.10. Let us denote by A the unitization of a C*-algebra A. Let (Ai)iEI be a family of C* -algebras. Then it is easy to check that the unitization of *iEIAi can be identified canonically with *iEIAi; in short we have
Proof of Theorem 13.2. We may clearly assume dim(H) = 00. It clearly suffices to prove (13.2) in case I is finite; hence, by iteration, we may as well assume that I = {I,2}. We will apply Theorem 13.4 to the subspace E 0 B(H) c (At * A2) 0 B(H) defined in Lemma 13.9. By the latter Lemma, since Mn(B(H» ~ B(H); the assertion (i) in Theorem 13.4 is satisfied (with
13. C*(lF ao ) Q9 B(H)
257
E I ,E2 now replaced by E,B(H)). Therefore, by Theorem 13.4, we have (13.2). To prove Theorem 13.2 in the nonunital case, we simply replace AI, A2 by their unitizations, which clearly still satisfy (13.1). Then, by the unital case, Al *A 2 appears (see the preceding remark) as an ideal in a unital C*algebra A such that A Q9min B(H) = A Q9max B(H). But, by Corollary 11.9, this property is inherited by closed ideals. • Curiously, the dec-norm considered in the preceding two chapters admits several slightly different descriptions. \Ve start with the most convenient one. Lemma 13.11. Let XI, .•• ,X n be elements in a C* -algebra. A, and let u: A be tIle linear map defined by u( (O'i)) = L 0'; X;.. Then
f~ --+
(13.7)
where the infimum TUns over all tIle factorizations Xi = aibi witll aj E A and bi E A. Proof. If u is positive, that is, if Xi ~ 0 for all i, this is very easy: The · I decomposltlOn ... . I 1/2 Xi1/2 . op t Ima IS simp y Xi = Xi Let us denote temporarily by !!!(x;)!!! the right side of (13.7). Assume first that !!U!!dec < 1. Then, going back to the definition of the dec-norm, we can find Yi,Zi ~ 0 in A with Yi,Zi. ~ 0, !!LYi!! < 1, !!Lzill < 1, and such that for all
i.
Then we have (rectangular matrix product)
Xi = (O,I)ti Let "Yi = (0, l)t:/2 and
Let "Yi =
(Ci' d i )
E
G)·
1\f1,2(A) and t5i = t:/ 2@
and t5i = (:;) so that
Xi = Ciri
E
lIhtCA). We have Xi = "Yit5i
+ diS; with
Assume A unital for simplicity. Let c > 0 and let ai = (cici + didi + c1)1/2 and bi = ai1xi. We can choose c > 0 small enough so that
258
Introduction to Operator Space Theory
II Eaiaill < 1. We then have bi = (a;lci,a;ldi )8i ; hence bibi = 8iwiwi8i, where Wi = (a;lci,a;ldd. Note that wiwi = a;l(cjci + djdi)a;l :s: a;l(a~)a;l = 1; hence bibi :s: 8i8i = riri + sisi, whence liE bibill < 1, so that we conclude 111(Xi)111 < 1. By homogeneity, this shows 111(xj)111 :s: Ilulidec. The converse follows from Exercise 12.2 (using the usual diagonal embedding f~ c Mn). • As a consequence, we easily derive the following (known) fact.
Lemma 13.12. Let n isometric identity
~
1.
Let A be a C*-algebra.
D(fn00' A)**
= D(fn0 0 A**) '·
Then we have all (13.8)
Proof. By Lemma 13.11, this statement reduces to the following fact: Given an n-tuple (Xi) in A** we have 111(Xi)111 :s: 1 iff there is a net (xf) of n-tuples in A with 111(xf)111 :s: 1 such that xi -4 Xi a(A**,A*) for each i. The "if" part is easy and left to the reader. To prove the "only if" part assume without loss of generality that 111(Xi)111 < 1, so that Xi = aibi with ai,b i in A** such that IIEaiail1 < 1 IIEbibil1 < 1. Let a E Mn(A**) (resp. bE Mn(A**» be the matrix admitting (ai) (resp. (b i )) as its first row (resp. column) and zero elsewhere. Let a'" (resp. b"') be a net in the unit ball of Mn(A) tending a(lIfn(A)**, Mn(A)*) to a (resp. b). By Kaplansky's Theorem we can even find a net for which the convergence is in the strong sense. We define xi = a"'(I, i)b"'(i, 1). Clearly 111(xf)111 :s: 1 and xi - 4 Xi in the a(A**, A*)sense. (The strong convergence of a"', b'" implies the weak convergence of x"'; moreover, on bounded sets weak and a-weak convergences coincide.) • Remark. The reader should be warned that the analogous identity for c. b. maps, nainely C B (f~, A) ** = C B (f~, A **), fails to be isometric in general when n > 2 and A is not locally reflexive. See Chapter 18 for more on this theme. We include here the following simple observation, which is implicit in [HI, Lemma 3.5]. Lemma 13.13. Let F be a free group and let (Ui)iEI be the family of free unitary operators in C*(F) associated to the.generators, to which we add the unit element. Let (Xi)iEI be a finitely supported family in a C* -algebra A, and let u: foo(1) -4 A be the mapping defined by U((Oi)iEI) = EiEIOjXi. Then we have Ui ® Xiii = Il L iEI C*(F)®maxA
lIulidec.
(13.9)
Proof. Let E = span[Ui , liE I). Let y = E Ui ® Xi E E ® A. Recall that, by Theorem 12.1, 8(y) = ~(y). By Theorem 12.7 and Lemma 8.9 (together
13. C*(lFoo ) 0 B(H) with the remark after it), we have
lIulidec = ~(y),
259 and by definition
where the supremum runs over all representations 11': A ---4 B(H) and all families (Vi) of contractions in 1I'(A)' C B(H). By the Russo-Dye Theorem ([Ped, p. 4]), the latter supremum remains unchanged if we let it run only over all the families (Vi) of unitaries in 1I'(A)'. Equivalently, this means:
~(y)
=
IlL
Ui 0
Xiii
I.E!
' C*(F)0mRx A
•
which establishes (13.9).
In [Ki5], Kirchberg also proves a general result on the tensor products C 0 N when N is an arbitmry von Neumann algebra. In that case (but with C an arbitrary C*-algebra), we can define a C*-norm II Iinor on C 0 N as follows:
where the supremum runs over all pairs of representations a: C ---4 B(1t) 11': N ---4 B(1t) with commuting ranges and with 11' normal. We denote by C 0 nor N the completion of C 0 N for this norm. (See [ELl for more information.) The preceding method also yields this result, but we only state it: Theorem 13.14. ([Ki5J) Let F be any free group and let C = C*(F). Let N be any von Neumann algebra. Then C 0
110r
N
=C
0rnax N.
The reader will find in Chapter 15 a characterization of the C* -algebras A such that C*(Foo) 0rnin A = C*(Foo) 0rnax A and in Chapter 16 a characterization of those such that A 0rnin B(H)
= A 0rnax B(H).
Exercise Exercise 13.1. Let G be any free group with free generators {gi liE I}. Show that C* (G) can be embedded into a direct sum of matrix algebras ([Ch2]).
Introduction to Operator Space Theory
260
Hint: Let {11" 1 11" E Go} denote the collection of all the finite-dimensional unitary representations of G (without repetitions). We define VtEG
a(t) = EB{1I"(t) 111" EGo}.
Clearly, a extends to a (contractive) representation a: C* (G) --+ E97r EG~0 B (H7r ). We claim that is isometric. Let A be the (incomplete) *-algebra generated bya(G). To prove this claim, it clearly suffices to show that (a)j1 is (completely) contractive. By Proposition 13.6, for the latter it suffices to show that the restriction of (a)-l to the linear span E of I and {a(gi) 1 i E I} is completely contractive, which indeed follows from Theorem 8.9.
a
Chapter 14. Kirchberg's Theorem on Decomposable Maps We have seen (see (11.1) and (11.6)) that c.p. maps (and more generally decomposable maps) "tensorize" for the maximal tensor product. Actually, as Kirchberg proved, the converse also holds, and this is the main result of this section: Theorem 14.1. ((Ki6J) Let A, B be C*-algebras and let u: A ----> B be a linear map. Let liS denote by iB: B ----> B** the canonical inclllsion map of B into B** viewed as a von Neul11ann algebra as usual. TIle following are equivalent:
(i) The map iBu: A
--4
B** is decomposa.ble witll
IliBUlldec ::; 1.
(U) For any C*-algebra D, we have
VXED®A
II(ID ® u)(x)IID0",axB
::;
IlxllD0lUaxA.
Remark. It is easy to check that (i) holds iff Ilu**lldec{A**.B**)::; 1. The proof uses several results on c.p. maps relative to "multiplicative domains" ([ChI]), which have not yet been discussed here, so we start by a brief review. Lemma 14.2. Let u: A --4 B be a c.p. map between C* -algebras Ilull::; 1. (i) Then, if a E A satisfies u(a*a) = u(a)*u(a), we have necessarily VxE A
u(xa)
Wit11
= u(x)u(a),
and tIle set of such a forms an algebra. (U) Let C u = {a E A I u(a*a) = u(a)*u(a) and u(aa*) = u(a)u(a)*}. Then C u is a C* -subalgebra of A satisfying Va, b E C u
"Ix E A
u(axb)
= u(a)u(x)u(b).
Proof. First observe that u satisfies the following kind of Cauchy-Schwarz inequality (see [Kad, ChI] for similar results for positive or merely 2-positive maps); VXEA u(x*x) ~ u(x)*u(x). (14.1) This is easy for c.p. maps. Indeed, by Arveson's Theorem, we can write u as u(·) = V*7I'(·)V for some representation 71': A ----> B(ii) and V: H --4 ii with B C B(H). Then we have for all T with 0 ::; T::; 1 u(x*x) = V*7I'(x)*7I'(x)V ~ V*7I'(x)*T7I'(x)V.
Introduction to Operator Space Theory
262
Hence, choosing T = VV*, we obtain (14.1). This implies that the "defect" cp(x, y)
= u(x*y) - u(x)*u(y)
behaves like a B-valued scalar product. In particular, we clearly have (by Cauchy-Schwarz) v~ E
H
Vx,y E A
This shows (taking y = a) that, if cp(a, a) = 0, we have (cp(x, a)~,~) = 0 for all ~; hence cp(x, a) = 0 for all x in A. Changing x to x* (and recalling that a c.p. map is self-adjoint) we obtain u(xa)
= u(x)u(a) for all x in A.
Thus we have proved {a E A I u(a*a) = u(a)*u(a)} = {a E A I u(xa) = u(x)u(a)
Vx E A}.
(14.2) It is easy to see that the right side of this equality is an algebra. This proves (i). To check (ii), we note that by reversing the roles of a and a* in (i) we have u(aa*) = u(a)u(a)* iff u(ay)
= u(a)u(y) for all yin A.
Note that a E C u iff both a and a* belong to the set (14.2). Therefore, C u is a C* -algebra and we have for any a, b in Cu and any x in A u(axb)
= u(a)u(xb) = u(a)[u(x)u(b)].
As a consequence, we have
•
Lemma 14.3. Let C be a C*-algebra and let 11'": C -> 1I'"(C) c B(H) be a representation. Assume C C B(K). Then any contractive c.p. map (in
particular any unital c.p. map) u: B(K)
->
B(H) extending 11'" must satisfy
that is, u must be a C-bimodule map (for the action defined by 11'"). Proof. Indeed, with the notation in Lemma 14.2 we have clearly C C C u ; hence this follows from the second part of Lemma 14.2. • A basic (elementary) fact used in the sequel is that for any x E B(H) we have (see Exercise 11.5 (ii))
II xII :5 1
iff
(;.
~)
2: O.
The next two lemmas are extensions of this to the case when x is replaced by a c.p. map (resp. a bimodule map) from an operator space (resp. an operator bimodule) into B(H).
14. Kirchberg's TI1eorem on Decomposable
~l\Iaps
263
Lemma 14.4. Let v: E ---+ B(H) be a map defined on an operator space E C B(K). Let S C .M2(B(K» be the operator system consisting of all matrices
;1)
(~!
with A,1l E
S
---+
lIh(B(H)) be
the mappillg defined by
a)) - (AI v(a») v(b)* III .
III
Then
Ilvllcb :'S 1 iff V
is c.p.
Actually, we will use a generalization of this lemma in the setting of "operator modules," as follows (this result appears in [SuJ; see also [PaSuJ).
Lemma 14.5. Let C C B(K) be a C*-algebra given with a representation n: C ---+ B(H). Let E C B(K) be a C-billlodule, that is, an operator space stable by (left and right) multiplication by any element of C. C011sider a bimodule map w: E ---+ B(H), that is, a map satisrying W(C1XC2) = n(ct}'w(x)n(c2)' (C1,C2 E C,x E E). Let S c lIf2(B(K» be the operator system consisting of all matrices of tIle form Let TV: S
---+
with A,ll E C, a, bEE.
M2(B(H» be defined by TV
Then
(~ ~)
Ilwllcb :'S 1 iff W
((~ ~)) - C:f~)* ~((~n· is c.p.
Proof. The easy direction is W c.p. => Ilwllcb :'S 1. Indeed, if l-V is c.p., we have IIlVllcb = IllV(l)1I = 1, and a fortiori Ilwllcb :'S 1. We now turn to the converse. Assume Ilwllcb :'S 1. Consider an element S E
Mn(S), say,
S
(~ ~).
=
A,ll
E
lIfn(C), a,b
E
lIfn(E). Assume
S
~ O.
Then necessarily A, 11 E Mn (C) + and a = b. Fix c > 0 and let Ae = A + d and lIe = 11 + d (invertible perturbations of A and 11)' Let Se = S + d. Let us denote Xc = X;1/2QIl;1/2 and let Hrn = hI" Q9lV, Wn = hI" Q9w. vVe then have 1 ( x* e
0.)
Xe)_(A;1/2 1
-
0
-1/2
lIe
Se
(A;1/2 0
0).
-1/2,
lIe
(14.3)
hence the left side of the preceding equation is ~ 0, which implies by the basic fact recalled above that Ilxell :'S 1. Therefore, if IIwllcb :'S 1, we have Ilwn(xe)11 :'S 1, which implies that (
1
wn(xe)*
Wn(IXe»)
~
O.
264
Introduction to Operator Space Theory
But this last matrix is the same as Wn (;;
~e).
Now, applying Wn to
both sides of (14.3) and using the fact that w is a C-bimodule map, we find
Since the left side is;:::: 0, we have Wn(se) ;:::: 0, and letting that Wn(s) ;:::: 0, whence that W is c.p.
€ ->
°we conclude •
Kirchberg's Proof of Theorem 14.1. We assume A unital for simplicity. The implication (i) ::::} (ii) is easy using the isometric embedding C 0max B C C0 max B** (see Exercise 11.6). Hence the main point is the converse. Assume (ii). Let M = B** viewed as a von Neumann algebra embedded in B(H) for some H. We have B c B**. We will use D = Jl;f'. We will work with the mapping w: M' 0max A -> B(H) defined by w(L:c,; 0ai) = L:ciu(ai) = L:u(ai)ci. Note that IIwllcb ::; 1. Indeed, replacing D by Mn(D), we see that (if (ii) holds) ID 0 u: D 0max A -> D 0 max B is completely contractive and w is the composition of this map with the *-homomorphism a: M' 0max B -> B(H) defined by a(c 0 b) = cb = bc. Let K be a suitable Hilbert space so that M' 0max A C B(K). Let C = M' 01 c B(K), and let E = M' 0max A. Note that E is a C-bimodule and that w: E -> B(H) is a C-bimodule map. Hence, with the same notation as in Lemma 14.5, the mapping W: S -> M2(B(H)) must be completely positive (and unital). Recall (see Corollary 1.8) that for a unital mapping on a C*-algebra, "completely contractive" is equivalent to "completely positive"; hence, by Corollary 1.7, any unital c.p. map V: Al -> B(H) on a (unital) C*-subalgebra Al C A2 admits a (unital) c.p. extension V: A2 -> B(H). Thus, let W: M2(B(K)) -> M2(B(H)) be a completely positive extension of W, and let T: .M2(A) -> M2(B(H)) be its restriction to M 2(A). (Here we identify A and 1 0 A c B(K).) Let u*(x) = u(x*)*. We then claim that T is of the following form:
T(x)
=
(
TU(XU)
U.(X2t}
U(XI2)) T22 (X22)
,
where Tu and T22 are c.p. maps from A into M = B**, with IITull, IIT2211 ::; 1. Taking this claim for granted, it is easy to conclude the proof since we can
.
14. Kirchberg's Theorem
011
Decomposable Maps
265
then define Va E A
R(a) = T
(:
:
) = (
Tn(a) u*(a)
Since T is c.p. and the mapping a
----> ( :
:)
is clearly c.p., R must be c.p.
Hence, by the definition of the dec-norm, we have
Thus it only remains to prove the claim. For this purpose, observe that, by its definition, T-V is a *-homomorphism on the algebra of matrices
(ci ~ )
with Cl, C2 in C = !vl' 0 1. Indeed, let C denote the set of all such matrices, and let 71": C ----> lIh(M') be defined by 71" ( Cl
~ 1 ~ 1) = C2
(ci
~).
Then the definition of W shows that WI~= 71". Therefore, since W is a c.p. extension of 71", Lemma 14.3 implies that W must be a C-bimodule map; that is, we have
Applying this with Cl, C2 scalars, we find (by linear algebra) that W is necessarily~uch that W(X)ij depends only on Xij. A fortiori the same is true for T = lVj1lf2 (A). Thus we can write a priori
Taking X11 = X22 = 0 and recalling that lV extends lV, we find T12 = u and = u*. Moreover, since T is unital and c.p., the same is true for Tn and T 22 , hence, we have IIT1111 ~ 1 and IIT2211 ~ 1 by Exercise 11.5 (iii). Finally, it remains to check that Tn and T22 take their values into B** = M. For that it suffices to check that T n (X11) and T 22 (X22) commute with M'. But this is an easy consequence of (14.4). Indeed, let Xll E A, X22 E A and let
T21
x-
(
10Xll
0
and
c=
(
C1
01
0
C2
0) 0 1
.
266
Introduction to Operator Space Theory
Then, with the same notation as in (14.4), we have ex = xc, and hence rr(c)T(x) = T(x)rr(c), which implies that T 11 (X11) and T22 (X22) commute with M'; hence they take their values in M" = Af = B*". This ends the proof of the claim and of Theorem 14.1. • The same proof yields an extension theorem as follows: Theorem 14.6. Consider two C* -algebras A, B, a subspace X linear map u: X --+ B. The following are equivalent:
(i) The map u admits adecomposable extension u: A 1. (ii) For any C*-algebra D, we have
--+
c A,
and a
B** with IIulidec
:::;
VxED®X
Proof. We simply repeat the preceding proof with E
= AI' ®X c
Af' ®max A .
•
Chapter 15. The Weak Expectation Property (WEP) In this short chapter we discuss a property introduced by Lance [Lal-2] as a generalization of nuclearity, namely, the weak expectation property (WEP). Definition 15.1. A C*-algebra A l18s the WEP (or "is WEP") if tIle inclusion map iA: A ---> A** factors completely positively and completely contractively through B(H) for some H; that is, we have completely positive complete contractions T 1: B(H) ---> A** and T 2: A ---> B(H) such that TIT2 = iA. Remark 15.2. (i) If we assume A** embedded as a von Neumann subalgebra in B(H). Then the WEP implies that there is a completely positive and completely contractive mapping T: B(H) ---> A** such that T(a) = a for all a in a. Thus we can take Tl = T and T2 equal to the inclusion mapping of A into B(H). Note that in general T will not be a projection. But of course, if A** is injective, then we can find a projection from B(H) onto A** so that a fortiori A **, and hence A, has the WEP. (ii) A von Neumann algebra AI has the WEP iff it is injective. Indeed, if Al has the WEP, it is injective since there is a projection P: AI** ---> lIf. which is actually a normal *-homomorphism (P can be described as the canonical normal *-homomorphism extending the identity M ---> 1If). Conversely. if Me B(H) is injective, there is a c.p. projection from B(H) onto it, so AI is
WEP. Since we know that A** is injective whenever A is nuclear (see Theorem 11.6), w~ find that nuclearity implies the WEP. A more direct proof can be given using the following result. Proposition 15.3. Let A be a C* -algebra. The following are equimJent:
(i) A has the WEP. (ii) For any embedding j: A c B into a C* -algebra B and for any C*algebra C, the associated morpllism A Q9lllax C ---> B Q9lllax C is isometric. (iii) Tlwre is an embedding A C B (H) such tllat, if C = C* (IF 00), tIle associated morphism A Q9lllax C ---> B(H) Q9lllax C is isometric. (iv) Tllere is an embedding A C B(H) such that the associated morpllism A Q9lllax C ---> B(H) Q9max C is isometric for any C*-algebra C. Proof. Assume (i). Let Tl and T2 be as above. By the extension property (see Corollary 1.7), T2 admits a completely contractive extension T2: B ---> B(H) and by (11.5) weJIave IIT211dec = IIT211cb ::; 1. Then, by (11.3) and (11.4), the mapping T = T 1T2 : B ---> A** is decomposable with IITlIdec ::; 1 and satisfies 71A = iA· By (11.7) we have
liT ®
Ie: B
®Illax
C
--->
A**
®Illax
CII ::; 1.
268
Introduction to Operator Space TheOIY
Hence, for any t in A ® C, we have
but since the inclusion A®maxC -+ A** ®maxC is isometric (cf. Exercise 11.6), this implies IltIIA® C ::; IltIIB® C, and the converse inequality is obvious. This shows that (i) ::::} (ii). The implication (ii) ::::} (iii) is trivial. Assume (iii). We first claim that (iii) remains true when IFoo is replaced by an arbitrary free group IF. Indeed, it is easy to see that, for any finitedimensional subspace E C C*(IF), there is a C*-subalgebra D with E cDc C* (F) such that D ~ C* (IF 00) and there is a completely positive contractive projection from C*(F) onto D. (Hint: Any element in C*(IF) can be described using only count ably many elements in IF, and hence only countably many generators. Then we can invoke Proposition 8.8.) By (11.1), our claim follows. Now, let C be any C*-algebra. Let IF be a large enough free group so that, if B = C*(IF), we can write C = BII for some ideal I (see Exercise 8.1). By Exercise 11.2, we have IllRX
I1IRX
B®max A (B I I) ®max A ~ I A' ®max from which it is easy to see that, if (iii) holds for C C = B II. This ~hows that (iii) implies (iv).
=
B, it also holds for
Finally, to show that (iv) implies (i), one way is to apply Theorem 14.6 to our embedding A C B(H). Assuming A unital, it follows that there is a completely contractive map T: B(H) -+ A** extending the inclusion A -+ AU. Since T is unital, it must be c.p. by Corollary 1.8. For the convenience of the reader, we will give a more explicit argument based on the same idea but avoiding decomposable maps. Let 1t be large enough so that A** can be realized as a von Neumann subalgebra of B(1t). Let 7r: A** -+ B(1t) be the inclusion map, and let C = 7r(A**)'. Let u: A ®max C -+ B(1t) be the representation defined by u(a ® c) = 7r(a)c. Since A ®max C C B(H) ®max C, there is a (unital) complete contraction B(H) ®max C -+ B(1t) extending u with lIull = Ilull = 1. By Corollary 1.8, u is c.p., and by Lemma 14.3, u must be an (A ® C)-bimodule map. In particular, if we define TI : B(H) -+ B(1t) by TI(b) = u(b ® 1) for any b in B(H), then we must have for any c in C
u:
Therefore TI(b) commutes with C = 7r(A**)', and hence TI(b) E 7r(A**)" = 7r(A**). Thus we obtain a c.p. map TI: B(H) -+ A** with IITIII = 1 such that TIIA = iA, which proves that A is WEP. This shows that (iv) implies (i) . •
.
15. The Weak Expectation Propert'y (WEP)
269
Corollary 15.4. An'y nuclear C* -algebra lIas the WEP. Proof. With the same notation as in Proposition 15.3(ii), if A is nuclear, there is a unique C* -norm on A 0 C, but the norm induced by B 0 max C on A 0 C is a C* -norm; hence (ii) in Proposition 15.3 must hold. • The preceding statement can also be derived from the next one, which gives Kirchberg's striking characterization of the WEP (cf. [Ki2]).
Theorem 15.5. Let C = C*(lF oo ). Then a C*-algebra A has the WEP iff
C0
m in
A = C 0 max A.
Proof. Assume C 0 m in A = C 0 max A. \Ve claim that (iii) in Proposition 15.3 holds. Assume A C B(H). Consider t E C 0 A. Then we have
Hence (since Iltlb8>max A 2: Iltllc0max B (H) is trivial) we obtain our claim, so that A has the WEP by Proposition 15.3. Conversely, assume A WEP. Then, by Proposition 15.3, the inclusion
C 0 max A
---t
C 0 max B(H)
is isometric. On the other hand, of course, by the injectivity of the min-norm. the inclusion C 0 m in A ---t C 0 m in B(H) is isometric as well. But now, by Theorem 13.1 we know that C0 I1lax B(H) = C 0 m in B(H), hence we must have equality of the norms induced on C 0 A, which yields C 0 max A = C 0 m in A. •
Remark. The same argument shows that if A has the WEP, then for any free group F we have C*(F) 0 m in A
= C*(F) 0 max A.
The ¥lEP does not seem to have been studied much in the operator space context, although its definition extends rather naturally, as follows.
Definition. Let X be an operator space. Let ix: X ---t X** be the canollical inclusion. Let >. 2: 1. lVe sa'y that X has the >.- \YEP if there are maps T 1 : B(H) ---t X** and T 2 : X ---t B(H) SUell that ix = TIT2 and satisfYing
Introduction to Operator Space Theory
270
Note that 1 ::; IITdlcbllT21lcb; hence, when oX = 1, this implies that we may choose Tl ,T2 so that IITlllcb = IIT211cb = 1. It is not at all obvious that, if a C* -algebra A has the oX- WEP for some oX ~ 1, then it has the WEP in the sense of Definition 15.1, but fortunately this is indeed true. It follows from an unpublished result due to Haagerup [H4] (see Remark 15.12). Thus we may say that an operator space has the WEP if it has the oX- WEP for some oX ~ 1 or, equivalently, if i x factors completely boundedly through B(H). Haagerup's result [H4] is as follows.
Theorem 15.6. The following properties of a C* -algebra A are equivalent: (i) A has the WEP. (ii) For any n and any aI, ... ,an in A we have
(iii) There is a constant C such that, for all n and all have
al, . . .
,an in A, we
We only give indications on the proof. In particular we will omit the proof of the most delicate implication (ii) ::::} (i). Remark. It is worthwhile to recall:
A
~
AOP via the correspondence a
--+ a*.
Hence we have in particular
The following key result plays an important role in the proof.
Theorem 15.7. Let A be any C*-algebra. Fix an integer n ~ 1. Let Eo (resp. E l ) be the space of n-tuples (al,"" an) in A equipped with the norm
We then have (15.1) Note that here either Eo or E1> or the complex interpolation space (Eo, E l )8, are the same spaces (namely An) equipped with different norms. The preceding equality computes the "interpolated norm" for () = 1/2. Haagerup's unpublished results extend those of the previous paper [P7] restricted to the case when A is a semi-finite von Neumann algebra, as follows.
15. TIle Weak Expectation Property (WEP)
271
Theorem 15.8. Let A be a von Neumann algebra., equipped with a normal, faithful semi-finite trace T, so that we can define in a standard way tIle spa.ce L2 (T). Tllen, if we denote by a ---> L( a) (resp. a ---> R( a)) tlle operator of left (resp. right) multiplication by a acting 011 L 2(T), tllen for any at, ... ,an ill A we llave II(ai)II(Eo ,Etll/2
=
IlL L(ai)R(anl/:(:2(T)) .
(15.2)
Note that (15.2) immediately implies that II(ai)II(Eo,Edl/2 :::; III: ai Q9 aill~;x' Conversely, we have by (1.11)
Remark.
and also since ai
Q9
1 and 1 Q9 bi commute
Hence, by the complex interpolation theorem ([BeL, p. 96]) for bilinear maps (noticing that (Eo, Edl/2 = (El , Eoh/2)' we must have
Finally, taking bi = ai we obtain
which yields half of (15.1). Now, in the case A = B(H), by Corollary 7.12 we have lI(ai)II(Eo.Etll/2 = II I: ai
Q9
aill~i~' Hence when A
= B(H) we obtain
(15.3) More generally, (15.3) holds if A is WEP. Indeed, if
are unital c.p. maps as in Definition 15.1, we have for any bi in B(H)
Introduction to Operator Space Tlleory
272
Hence we conclude that (15.3) holds in any WEP C* -algebra A. This proves the implication (i) =? (ii) in Theorem 15.6.
Remark. The proof that (iii)
=?
(ii) in Theorem 15.6 is elementary: Let
t = Lai ® ai. Since, for any C*-norm, we have IItl1 2k = II(t*t)kll and since
(t*t)k is again of the form
L Cl:j ® Cl:j' we have
Iltll~ax = II(t*t)kll max ~ CII(t*t)kll min = Clltll~~n; hence Iltll max ~ C,ft Iltllmin and letting k go to infinity we obtain (ii).
Remark. The implication that A ®min A = A ®max A
=?
A has the WEP
had been previously observed by Kirchberg in [Ki2]. The next two corollaries follow from Theorem 15.6.
Corollary 15.9. Let u: A --> B be a linear map between C* -algebras. Assume that u is c. b. (actually, up to the constant, it suffices to assume that IG ® u: G ®min A --> G ®min B is bounded when G = REB C). Then, for any finite sequence aI, ... ,an in A, we have (15.4)
Proof. Let C = Ilulicb. We have
IlL u(ai)*u(ai )11 1/ 2 ~
ClI(ai)IIEo and
IlL u(ai)u(ai)*111/2 ~
CII(ai)IIEI'
Hence, by interpolation (applying Theorem 15.7 both in A and in B), we deduce (15.4). •
Corollary 15.10. In the situation of Corollary 15.9, assume that u is a complete isomorphism. Then, if A has the ""VEP, so does B. Proof. The analog of (15.4) for the minimal C* -norm is clear since, if (Ti) is any orthonormal basis in 0 H, we have
Hence, if A has the WEP, we find
IlL u(ai)
® u(ai)llmax
~ Ilull~b
ilL ai
~ Ilull~bllu-lll~b and by (iii)
=?
® aiLax
=
ilL u(ai)
ilL ai u(ai) I . ,
Ilull~b
®
® aiLin
mm
(i) in Theorem 15.6 we conclude that B has the WEP.
•
The next result was originally proved, independently and differently, in [P3, CS3].
15. The Weak Expectation Property (WEP)
273
Corollary 15.11. Let M c B(H) be a VOll Neumann algebra. Assume that there is a projection P from B(H) onto M that is c.b. (or merely such that fa Q9 P: G Q9min B(H) --4 G Q9min M is bounded when G = REB C). Then Al is injective. Proof. By Corollary 15.9 applied to P: B(H) AJ, we can write
< C -
--4
IlL ail ai Q9
M, for any al, ... ,an in
-'
B(H)0max B (H)
Hence, by (15.3),
Thus, by (iii) algebra, WEP
*
*
(i) in Theorem 15.6, AI is WEP, and, for a von Neumann injective by Remark 15.2. •
Remark 15.12. Assume that the maps TI, T2 appearing in Definition 15.1 are merely c.b. (actually, for Tl it suffices even to assume that IaQ9Tl: GQ9 m in B(H) --4 GQ9 minA** is bounded when G = REBC). Then, A necessarily has the WEP. Indeed, by Corollary 15.9 (and the validity of (15.3) when A = B(H)), for any ai in A we have
Exercise 11.6 (applied twice), the left side is equal t.o IlL: ai Q9 aillA0max A ; hence we conclude by Theorem 15.6 that A has the WEP. Remark. Ozawa proved recently that B(£2) Q9min B(£2) fails the WEP (see But
by
[07.5]). We discuss several important open problems about the WEP at the end of Chapter 16.
Exercises Exercise 15.1. Let C
= C*(Foo). Show that a C*-algebra A is nuclear iff A**
Q9min C
= A** Q9max C.
Exercise 15.2. Consider the following property of a C* -algebra A: For any pair of C*-algebras Bll B2 with Bl c B 2, we have A ®max Bl C A ®max B 2. Show that this property characterizes nuclear C* -algebras.
274
Introduction to Operator Space Theory
Hint: By Theorem 11.6, it suffices to show that this property implies the injectivity of A **. To do that, use the well-known fact ([To2]) that a von Neumann algebra .111 c B(H) is injective iff its commutant AI' is injective. Exercise 15.3. Let M be a C* -algebra that is a quotient of a WEP C*algebra by an ideal (QWEP in short). Let AI c M be a C*-subalgebra such that there is a completely positive contractive projection P from M onto .111. Show that AI is QWEP. Exercise 15.4. Using the preceding exercise, show that if M is any (von Neumann sense) ultraproduct of matrix algebras as in Theorem 9.10.1, then any von Neumann subalgebra AI C M is QWEP.
Hint: Use the existence of a conditional expectation from M onto .1If. Exercise 15.5. Let G be a discrete group. If either C>:( G) or V N( G) CX(G)" is WEP, then G is amenable.
Hint: Use Exercise 8.4 and Theorem 15.6 (see also Exercise 11.8).
Chapter 16. The Local Lifting Property (LLP) In Banach space theory, the lifting property of £1 is classical: For any bounded linear map u from £1 into a quotient Banach space XIY and for any € > 0, there is a lifting u: £1 ~ X with lIuli ~ (1 + €)lluli. We will discuss in Chapter 24 the operator space analogs of this (see also [KyR]). But in this chapter we concentrate on the C* -algebraic analog of the lifting property as introduced by Kirchberg. Following [Ki2], we will say that a unital C* -algebra C has the lifting property (LP in short) if any unital c.p. map u from C to a quotient AI I (here A is a C* -algebra and I a two-sided closed ideal in A) admits a unital c.p. lifting u: C ~ A. Kirchberg (see [Ki2]) proved that, for any countable free group F, C*(F) (and also 111n ( C* (F» for any n ~ 1) has the LP. But actually very little is known about the LP, and it is rather the local version defined in the following that has proved fruitful. Definition 16.1. We say that a unital C*-algebra C llas the "local lifting property" (LLP ill sllOrt) if the following bolds for any C* -algebra A and any (closed two-sided) ideal I c A: For any unital c.p. map 1£: C ~ All and for any finite-dimensional subspace E c C, tbere is a complete contraction u: E ~ A tbat lifts UIE: E ~ AI I. In otber words, altllOugb a priori 1£ is not liftable, it rougbly "locally lifts." Wben A is l10t unital, we say tbat it bas tbe LLP if its unitization 1188 it. \Vhile the LP remains a rather elusive property up to now, Kirchberg [Ki2] discovered the following nice characterization for the LLP. Theorem 16.2. TIle following properties of a C* -algebra C are equivalent:
(i) C bas tIle LLP. (ii) C
Q9min
B(H) = C
(iii) C (iv) C
Q9min
B(H) = C Q9max B(H) for H = £2· B = C Q9max B for any C* -algebra B with the lVEP.
Q9min
Q9max
Proof. We will first prove (iii)
B(H) for any Hilbert space H.
=}
(i).
Assume (iii). Let A,I,E c C and u: E ~ All be as in Definition 16.1. Let t E C 181 E* be the tensor corresponding to the inclusion map E ~ C, and let s E (AI I) 181 E* be the one corresponding to UIE: E ~ AI I, so that S
=
(1£
181 I)(t).
We assume E* C B(H). Then by (2.3.2) we have Iitlic0mlnB(H)
= IIIElicb =
1.
276
Introduction to Operator Space Theory
Therefore, since we assume (iii), Iltlb~m ..xB(H)
= 1,
and therefore by (11.1) IIsll(A/I)®maxB(H) = lI(u i8l I)(t) II (A/l)®maxB(H) :::; 1 But now, by Exercise 11.2, we have
(Aj I) i8l max B(H) = (A i8l max B(H))j(I i8l max B(H)). Hence, by Lemma 2.4.6, there is an element sin the unit ball of Ai8l max B(H) such that, if q: A -+ Aj I denotes the quotient map, we have s
= (q i8l I)(s),
and of course a fortiori IISlIA®mlnB(H) :::; 1. By Lemmas 2.4.8 and 2.4.7, this implies that there is an element in the unit ball of A i8l min E* such that s = (q i8l I)(S). Let u: E -+ A be the linear map defined by s. Note that by (2.3.2) again lIullcb = Ilslimin :::; 1,
s
and s = (q i8l I)(S) equivalently means that ulE = quo Thus we conclude that C has the LLP. This completes the proof that (iii) => (i). Note that, by Theorem 13.1, this proves in particular that C*(IF) has the LLP for any free group IF. The proof that (i) => (ii) is based on Theorem 13.1. Assume (i). Consider t E C i8l B(H). We can assume tEE i8l B(H) with E c C finitedimensional. Let F be a free group such that C ~ C*(F)jI for some ideal Ie C*(F). Let us denote A = C*(F), let q: A -+ AjI ~ C be the quotient map, and let u: E -+ Aj I denote the natural inclusion. By the LLP of C, u admits a completely contractive lifting u: E -+ A, so that qu = u. We then have t = (q i8l I)(u i8l I)t, and hence IltIlC®maxB(H) :::; lI(u i8l I)(t)IIA®maxB(H) = II(u i8l I)(t)IIA®mlnB(H)
(by Theorem 13.1)
:::; IltIlE®mlnB(H) = IItIIC®mlnB(H). Hence we obtain (i) => (ii). (ii) => (iii) is trivial. Assume (iii). Let B be as in (iv). We may as well assume B C B(H) for some H. Now, if B has the WEP, we have C i8l m ax B C C i8l m ax B(H)
(isometrically)
16. The Local Lifting Property (LLP)
277
and obviously also
C
Q9min
Be C
Q9min
B(H)
(isometrically).
Hence, if (iii) holds, the induced norms on C 0 B must coincide so that Q9min B = C Q9max B. This establishes (iii) :::} (iv). Since (iv) :::} (iii) is obvious, and we already showed (iii) :::} (i), the proof is complete. •
C
Remark 16.3. (i) The preceding proof of (iii) :::} (i) shows that, if C has the LLP, any decomposable map u: C --> AI I with Ilulidec ::; 1 locally lifts in the following sense: For any finite-dimensional E c C, 111E: E --> AI I admits a completely contractive lifting E --> A.
u:
(ii) The preceding proof of (i) :::} (ii) shows that C has the LLP iff there is a free group F and a surjective representation q: C*(F) --> C such that the resulting isomorphism u: C --> C*(F)I ker(q) locally lifts in the preceding sense. Thus, if we restrict ourselves to unital representations u: C --> AI I in Definition 16.1, the resulting property is the same.
(iii) For a useful characterization of complete metric surjections that "locally lift" in the preceding sense, see Exercise 2.4.1. (iv) Consider a separable C* -algebra A with the metric CBAP (that. is, there is a sequence of completely contractive finite rank maps tending pointwise to the identity). Then, if A has the LLP, by Exercise 2.4.2 it must have the LP. (v) By Theorem 16.2, all nuclear C*-algebras have the LLP. Actually, in the separable case, they have the LP by the preceding point and by Theorem 11.5.
Remark. Fix A 2: 1. We could say that C has the A-LLP if, given any unital c.p. map u: C --> AI I and any finite-dimensional subspace E c C, the restriction UIE: E --> AI I admits a lifting E --> A such that Ilulicb ::; A. But this is not needed: Indeed, the preceding argument shows that if C has this A-LLP, then II ·lb8lmnxB(H) ::; All ·11c,~minB(H)'
u:
and since there is only one C* - norm on a C* -algebra, we must have C Q9max B (H) = C Q9min B (H) isometrically. In other words, for C* -algebras, this A-LLP implies the LLP (or, equivalently, the 1-LLP). We will now turn to the operator space version of the LLP introduced in
[Oz3J. Definition 16.4. Let A 2: 1. We say that an operator space X 1]88 the AOLLP if, for ally unital C* -algebra A, for any ideal I c A, for any complete contraction u: X --> A/I, and for any finite-dimensional subspace E C X, the restriction UIE: E --> A/I admits a lifting it: E --> A such that lIitli c b ::; A.
278
Introduction to Operator Space Theory
We say that X has the A-OLP if we can always take E = X in the preceding definition.
Remark. By Exercise 2.4.2, if X has the (A + e)-OLLP (resp. X has the (A + e)-OLP and is separable) for any e > 0, then it also has it for e = O. Moreover, if X is separable and has the metric CBAP (that is, there is a sequence of completely contractive finite rank maps tending pointwise to the identity), then (again by Exercise 2.4.2) X has the A-OLP if it has the A-OLLP.
Remark. The spaces max(f 1 ) and Sf have the 1-0LP (and a fortiori the 1-0LLP). Indeed, it is easy to check that they are "projective," that is, they satisfy a more general lifting property valid for all quotients of operator spaces (see Chapter 24) and not only for C* -algebra quotients. Theorem 16.5. ([Oz3}) Let X be an operator space. Then X has the 1OLLP iffC:(X) has the LLP. Proof. By the universal property of the embedding X '--+ C:(X), any com(X) --+ AI I. plete contraction u: X --+ AI I extends to a representation U: Thus, if C:(X) has the LLP (resp. LP), it is clear that X must have the 1-0LLP (resp. 1-0LP).
C:
Conversely assume that X has the 1-0LLP. Let 11': C: (X) --+ AI I be a unital representation. Let {E; liE I} denote the family of all the finitedimensional subspaces of X directed by inclusion. Recall (see Exercise 8.2) that the inclusion Ei C X extends to an isometric unital representation C:(Ei ) --+ C:(X). Since by assumption C~(X) has the LLP, for each i, the restriction 1I'1 E i: Ei --+ AI I admits a lifting rr: Ei --+ A with Ilrrllcb ::; 1, which extends to a unital representation 1I'i: C:(Ei ) --+ A. Let q: A --+ All denote the quotient map. Clearly q1l'i coincides with the restriction of 11' to C:(Ei ) C C:(X). Since the (directed) union of the spaces C:(Ei ) is clearly dense in C~(X), we have obtained a dense subspace AI C Cu(X) such that, for any finite-dimensional E C AI, the restriction 1I'1E: E --+ AI I admits a completely contractive lifting from E to A. By a simple perturbation argument (see Lemma 2.13.1), this implies that 11' locally lifts; hence we conclude, by Remark 16.3(ii), that C: (X) has the LLP. •
Remark. If X is separable, then X has the 1-0LP iff C:(X) has the LP. Indeed, if X is separable, then so is C:(X); hence (see Exercise 8.1) we can write C: (X) :;: AII for some ideal I in A = C*(lFoo )· Let 11': C=(X) --+ All be the corresponding isomorphism, and let u denote its restriction to X. Assume that X has
16. The Local Lifting Property (LLP)
279
the 1-0LP. Then, since Ilulicb = 1, u admits a completely contractive lifting X ---+ A. Let rr: C~ (X) ---+ A be the unital representation extending U. We obtain a factorization of the identity of C~ (X) as follows:
u:
But since A = C*(lFoo ) has the LP (cf. [Ki5]), we conclude from this factorization that C~ (X) must also have the LP. This proves the "only if" part. For the "if" part (which does not require separability), see the above proof of • Theorem 16.5. Theorem 16.6. ([Oz3]) Let A :::: 1. The following properties of all operator space X are equivalent.
(i) X has the A-OLLP. (ii) X 0min B(H) = X 06 B(H) for any Hand 8(t) ::; Alltllmin for any tin X 0B(H). (ii)' Same as (ii) for H = €2' (iii) For any finite-dimensional subspace E c X and any E > 0, tllere is, for some N:::: 1, a subspace G c MN and a factorization E~G*~X of tIle inclusion map E C X sllch that Ilvllcbllwllcb < A + E. Proof. Assume (i). Let F be a free group so that C~(X) ~ A/I with A = C*(F) (see Exercise 8.1). Let E C X be a finite-dimensional subspace and let u: E ---+ AI I be the inclusion map. Since we assume (i), there is a E ---+ A with Ilulicb ::; A. Hence V tEE 0 B(H) and we can write lifting t = (q 0 I)(u 0 I)t, and hence
u:
8(t) = IltllcZ(X)®umxB(H) hence by Theorem 13.1 hence by (2.3.2)
::; II(u 01)tIIA®maxB(H) = II(u 0 I)tIIA®miuB(H) ::; Ilullcblltilmin,
so that we obtain 8(t) ::; Alltllmin and (ii) holds. (ii) =} (ii)' is obvious. We now prove (ii)' =} (iii). Assume (ii)'. Let E be as in (iii). Since it is separable, we may embed E* (completely isometrically) into B(€2). Let tEE 0 B(€2) be the tensor associated to the inclusion map E* ---+ B(€2)' By (ii)' we have
Hence, by Corollary 12.5, we have for any X*
----+
E*
a'\.
E
> 0 a factorization of the form
c B(€2) /{3
MN
280
Introduction to Operator Space Theory
with Ilall cbll,8llcb < A + € and a*(MN*) eX. Hence, letting G = a(X*) and denoting bya1 and ,81 the restrictions of a and ,8, we find a factorization X* --+ E* . a1 '\. /,81 G with lIa1l1cbll,811lcb < A + €. Taking the adjoint diagram, we obtain (iii) with v = ,8i and w = ai. Assume (iii). Let u: X -+ AI I be a complete contraction and let E C X be finite-dimensional. Let G, v, w be as in (iii). Let U = uw: G* -+ AI I. Note that IIUlicb ::; lIullcbllwllcb ::; IIwllcb. By Lemma 2.4.8, since G C M N, we have (isometrically) . G ®min AI1= (G ®min A)/(G ®min I) Consequently, by (2.3.2) and by Lemma 2.4.7, U admits a lifting U: G* -+ A with IIUlicb ::; Ilwllcb. Then the mapping = Uv: E -+ A lifts ulE and satisfies lIulicb ::; Ilvllcbllwllcb. This shows that X has the A-OLLP. •
u
The next statement is immediate (take E = X in Theorem 16.6(iii)). Corollary 16.7. A finite-dimensional operator space E has the A-OLP (or, equivalently, the A-OLLP) iff for any € > 0 there is, for some N 2:: 1, a subspace G C .MN such that dcb(E*,G) < A +€. Remark 16.8. By a classical result (see, e.g., [HWW, p. 59]) if X is a separable Banach space with the metric approximation property, then any contraction u: X -+ AI I admits a contractive lifting X -+ A. Obviously this implies that the operator space max(X) has the 1-0LP. It is likely that this becomes false without assuming some kind of approximation property, but no couterexample is known (see Proposition 18.14 later). Here is another example:
u:
Proposition 16.9. ([Oz3]) Let M be a von Neumann algebra with predual M*. Then M* has the 1-0LLP, and even the 1-0LP if it is separable. Proof. Let q: A -+ AII be as before. We have a canonical surjective representation M ®min A -+ M ®min (All). By Lemma 2.4.6 and (2.3.2), M* -+ A with for any finite rank map u: M* -+ AI I there is a lifting lIulicb = lIullcb. (Indeed, let t E M ® (AI I) be the tensor associated to u. By Lemma 2.4.6, t has a lifting tin M ®min A with 1It]lmin = IItll min, and thus we can take for u: M* -+ A the linear map associated to t.)
u:
16. The Local Lifting Property (LLP)
281
Now, by the remark following Lemma 12.8, any complete contraction -+ AI I can be approximated pointwise by completely contractive finite rank maps U a : 1\[* -+ AII. By the preceding observation, each U a admits a completely contractive lifting a : 11[* -+ A, so that -+ U pointwise. Now, by Exercise 2.4.2, if M* is separable, this implies that u admits a completely contractive lifting; hence M* has the 1-0LP. In general, if we restrict to a separable subspace E C },,[*, then Exercise 2.4.2 implies that UIE admits a • completely contractive lifting, and hence a fortiori M* has the 1-0LLP. u: !v[*
u
quo
Let us denote 8 = B(C 2 ). We now turn to the lifting property relative to the "Calkin algebra," that is, the quotient 81K. The following result comes from Ozawa's thesis.
Theorem 16.10. ([Oz6]) Let A ~ 1. A separable operator space X has the A-OLLP iff every complete contraction u: X -+ 81K admits a c.b. lifting X -+ 8 with /lu/lcb ::; A.
u:
Proof. Assume that X has the A-OLLP and let u E CB(X,8IK). By the OLLP and the injectivity of 8, there is a net U a : X -+ 8 of maps with /lu a II cb ::; A such that quo tends pointwise to u, where q: 8 -+ 81 K denotes the quotient map. Then, by Exercise 2.4.2, we conclude that there is a "global" lifting U with /lullcb ::; A. This proves the "only if" part. To prove the "if" part, we will need the following lemma, for which we first introduce some notation. Consider a finite-dimensional operat.or space E C B(H), a separable unital C*-algebra B, and an ideal J C B. Let (): B -+ 8 be a unital c.p. map such that ()(J) C K. Let
0: and
v
(): (E
Q9min
B)I E
BIJ ~ 81K
Q9min
J
-+
(E Q9min 81(E
Q9min
K)
denote the associated complete contractions. \Ve can now state:
Lemma 16.11. ([Oz6]) For any y ill E IlyIIE0miu(B/J)
Q9
(B I J) we hare
= sup 11(1 Q9 0)(y)/lE0miu(13/K)
(16.1)
9EC
and v
Ily/I(E0mluB)/(E0mlnJ)
= sup /I ()(y)II(E0miu13)/(E0miuK).
(16.2)
9EC
wllere C denotes the set of all unital c.p. maps (): B
-+
8 stich that ()(J) C K.
Proof. This argument is inspired by an idea of Kirchberg. We first observe that in both cases the norm of Y is ~ the supremum because IIOlicb ::; 1 for
Introduction to Operator Space Theory
282
each (). We now turn to the converse inequality. We will only prove (16.2), which is the only one needed in the sequel (the proof of the first one is similar). We will show that, if the left side of (16.2) is > 1, then the right side is also > 1. Let vEE 0 m in B be such that d(v, E 0 m in .1) > 1.
(16.3)
We claim that there is a () in C such that
d«J 0 ())(v)
+ E 0 m in K) > 1.
From this claim, (16.2) follows easily (modulo our initial observation). Thus it remains to prove this claim. Since.1 is separable, there is an element h in .1 such that 0:::; h :::; 1 and.1 = h.1h (when.1 = K any strictly positive element of K will do). Let 9n be the indicator function ofthe interval [n -1, 1J and let Pn = 9n(h) E .1**. Then Pn is a projection, and, since lim 11(1 - Ph)hll = 0, we have: (16.4) v x E.1. lim II (1 - Pn)x(1 - Pn) II = 0 Let In E Co«O,
mbe the function equal to 9n on [nIn (t) = nt
Then h n
= In(h)
1,
IJ and such that
if 0 < t :::; n -1 .
E .1. Since (1 - Pn) ~ (1 - h n ) ~ 0, for ea~h n we have
1I[10(I-Pn)Jv[10(I-Pn)JIIE0min B ** ~ 11[10(I-hn)Jv[10(I-hn )IIIE0mln B > 1 (by our assumption (16.3) because
are all in E 0 m in .1). Note: Here the left and right multiplications are done in B(H) 0 B**, but in any case vEE 0 m in B implies [10 (1 - Pn)Jv[1 0 (1 - Pn)J E E 0 B**. Hence (see Proposition 2.12.1), for each n, for some suitable N(n) ~ 1, there is a unital c.p. map Tn: (1 - Pn)B(1 - Pn) -+ AfN(n) such that
11(10 Tn)[1 0 (1 - Pn)Jv[1 0 (1 - Pn)]ll > 1.
(16.5)
(Note: Here (I-Pn)B(I-Pn) is viewed as a C*-algebra with unit I-Pn; hence we have Tn{1 - Pn) = 1.) Then we define unital c.p. maps ()n: B -+ MN(n) and (): B -+ EBn MN(n) by
()n(b) = Tn«I- Pn)b(1 - Pn» and ()(b)
= E9n ()n(b).
16. TIle Local Lifting Property (LLP) We will view
ffin lIJN (n)
2R3
as an embedded block diagonally into
so that co({MN(nd) C K. By (16.4), we have B(J) c K, and using the fact that (16.5) holds for each n, we can easily verify that
dist((I®B)(v),E®min K) > 1.
•
Thus, we have proved our claim.
End of the Proof of Theorem 16.10. It remains to prove the "if" part. Assume that the lifting property into BIK holds as in Theorem 16.9. Consider a complete contraction u: X -4 B I I, as in Definition 16.4. Since X is separable, we may assume (by an elementary argument) that B is also separable. Let F c X be finite-dimensional. Let y E F* ® (B 11) be associated to ulF: F -4 BII. Let E = F*. Applying the lifting property for BIK to the tensor [I ® O](y) E F* ® BIK (for each B in C), we find (recall Lemma 2.4.7) v
II
B(Y) II (F* <8Ilni" B)/(F" <8Ilni n K)
:::;
A.
Hence, by (16.2), Ilyll(F*<8Irui"B)/(F*<8Ilni"l) :::;
A,
which means (again recall Lemma 2.4.7) that ulF admits an extension B with lIulicb :::; A. Thus we conclude that the X has the A-OLLP.
u:
F-4 •
Kirchberg's conjecture. The following conjecture due to Kirchberg [Ki2] is equivalent to many fundamental open questions about von Neumann algebras. In particular (see [Ki2]) it is equivalent to the Connes problem mentioned at the end of §9.1O. Its solution would be a major step forward. Conjecture 16.12. ([Ki2])
(i) Tlle C*-algebra C*(lFoo ) llas the WEP. The next statement gives several equivalent reformulations (see [Ki2] for more). Proposition 16.13. Eacll of tlle following assertions is equivalent to t1le preceding conjecture.
(ii) [fC = C*(Foo), we have C ®min C = C ®max C. (iii) Every C*-algebra is a quotient of a C*-a.Jgebra witll the WEP (such algebras are called QWEP).
284
Introduction to Operator Space T11eory
(iv) LLP =} WEP. (v) IfC is any C*-algebra with the LLP, we have C ®min C = C ®max C. (vi) For any finite-dimensional operator space E =/:. {O} wit11 dsdE*) = 1 (as defined in (0.2)), C*(E) bas the WEP. (vii) The C*-algebra C~(C) bas the WEP. Remark 16.14. For C*-algebras with the LLP, QWEP =} WEP. Indeed, if A has the LLP and if A := WI I with W having the WEP, the isomorphism A ---+ WII locally lifts up through W, from which it follows (say, by Theorem 15.5) that A has the WEP. Proof of Proposition 16.13. By Theorem 15.5, (i) <=> (ii). Moreover, (i) holds iff, for any free group F, C*(F) has the WEP. Since any C*-algebra A is a quotient of C*(F) for a suitable F (see Exercise 8.1), (i) implies that it is QWEP, and hence (i) =} (iii). Moreover, (iii) =} (iv) by Remark 16.14. Then (iv) =} (v) follows from Theorem 16.2 and (v) =} (ii) is clear since C*(IFoo) has the LLP. Thus we have proved that (i)-(v) are equivalent. Then (iv) =} (vi) is clear since (by Theorems 16.5 and 16.6) C~(C) has the LLP; (vi) =} (vii) is trivial. Thus it remains to show (vii) =} (i). It is easy to see that C~(C) is the universal unital C*-algebra of a contraction T. Any singly generated unital C*-algebra is a quotient of C~(C). Let C = C[-l, 1]. We claim that the full unital free product C * C is a quotient of C~ (C). Indeed, since C is singly generated by a self-adjoint contraction x, C*C is generated by a pair {Xl, X2} of self-adjoint contractions, hence C * C is singly generated by (Xl + iX2)/2, and therefore is a quotient of C~(C). A fortiori C~(C) admits as a quotient C * C, where C is any finite-dimensional commutative C* -algebra. Thus, for any pair G ll G 2 of finite Abelian groups, C*(G l * G2 ) is a quotient of C~(C), and hence (vii) implies that C*(G l * G 2 ) must have the WEP. But now it is well known that IFoo is a subgroup of Z3 * Z3; hence, by Proposition 8.8, C* (IF(0) itself must have the WEP. • Remark. By adapting the proof of Lemma 3.3 and Theorem 3.4 in [EHJ, one can show (as pointed out by Kirchberg [Ki2, Proposition 2.2(iv)]) that, whenever J is a WEP ideal in C = C*(IFoo), for the quotient CIJ the LLP implies "automatically" the LP. See also [Harl, Corollary 3.3] for a related useful fact. Note that obviously J has the WEP if C does. Thus, if Conjecture 16.12 is true, we must have LLP =} LP (hence also OLLP =} OLP) in the separable case. \Ve discuss several important open problems concerning the OLLP in Chapter 18.
Chapter 17. Exactness The notion of exactness was int.roduced by Kirchberg as early as 1977 [Ki4J, but. his major contributions on that topic did not circulate until the late 1980s; cf. [Kil-2].
Definition. An operator space X is called exact if, for any C* -algebra B and any (closed two-sided) idea.l Ie B, we have all exact sequence {O}
-+
I 0 min X
-+
B 0 min X
-+
(BII) 0 min X
-+
(17.1 )
{O}.
Remark. The analogous property for the max-tensor product holds if X is replaced by an arbitrary C* -algebra A. Indeed, first I 0 max A embeds isometrically into B 0 max A and second, by maximality, the max-norm on (BII) 0 A must dominate the C*-norm induced by (B 0 max A)/(I Q9max A) (see Exercise 11.1 for more details). Hence B0max A (BII) 0 max A ~ I A Q9nlax
and consequently the sequence
{O}
-+
I 0 max A
-+
B 0 max A
-+
(BII) 0 max A
-+
{O}
(17.2)
is always exact. This goes back to [Gul]. In particular, this remark shows that alllluclear C*-algebras are exact. By the properties of C* -representations, if X is a C* -algebra, it suffices for the exactness of the sequence (17.1) that the kernel of B0 minX -+ (B II)0 rn in X coincides with I 0 min X. But in the operator space case, the exactness of (17.1) requires in addition that the map B 0 min X -+ (BII) 0 min X be surjective. (Since C*representations all have closed ranges, this is automatic in the C* -case.) l\Joreover, in the C*-case all morphisms have norm 1, which of course is no longer true for general c.b. maps, so we need to introduce the "constant of exactness" as follows: We have a complete contraction B 0 min X -+ (B II) 0 min X associated to the quotient map q: B -+ B II. Since this map vanishes on I 0 min X, it defines a (completely contractive) map
Tx: (B 0 min X)/(I 0 m in X)
-+
(BII) 0 m in X.
By definition, X is exact iff T..'( is an isomorphism for any B and I. Note that, assuming Tx injective, B Q9rnin X -+ (BII) Q9min X is surjective iff Tx l is bounded so the norm II T.X- 111 measures the degree of exactness of (17.1).We then denote (17.3) ex(X) = sup IIT.ylll, where the supremum runs over all C* -algebras B and all ideals I
c B.
286
Introduction to Operator Space Theory
It is easy to see that ex(MN) any nuclear C* -algebra A.
= lor,
more generally, that ex(A)
= 1 for
Kirchberg discovered that exactness is closely connected to the existence of good embed dings into nuclear C* -algebras and discussed various notions of "subnuclearity." For operator spaces, it is advantageous to first restrict the discussion to the finite-dimensional case: Let E c B(H) be a finitedimensional operator space. Recall the notation
dS1C(E) = inf{dcb(E, F) IF c K}.
(17.4)
By a simple perturbation argument (see Lemma 2.13.4), it can be shown that
= inf{dcb(E,F) IF c MN,N
dsdE)
~
(17.4)'
dimE}.
Theorem 17.1. Let X c B(H) be an operator space. For any fixed A ~ 1, the following assertions are equivalent:
(i) X is exact and ex(X) :::; A. (ii) Let u: X ---+ B(H) be the inclusion map. For any C* -algebra C, tll€ mapping Ie ® u is bounded from C ®min X to C ®max B(H) with norm :::; A.
(iii) For any finite-dimensional subspace E C X we have dsdE) :::; A. In particular, for any n-dimensional subspace E C B(H) we have dsdE)
= sup{IIE ®min C
---+
B(H)
®max
CII},
(17.5)
where the supremum runs over all possible C* -algebras C. First Part of the Proof of Theorem 17.1. We will show (i) ::::} (ii) ::::} (iii). Assume (i). Let C be any C*-algebra. By Exercise 8.1, if G is a sufficiently large free group, then C is a quotient of C*(G). So if we denote B = C*(G), we have C = BII for some ideal I C B. Let us denote Q(X) = (B ®min X)/(I ®min X). By Theorem 13.1 (and Corollary 11.9), we have Q(B(H» = (B®maxB(H»II®maxB(H). Hence, by the exactness of (17.2),
Q(B(H»
= (BII) ®max B(H) = C ®max B(H).
Then Ie ® u coincides with the following composition:
C
®min
X
= (BII) ®min X
---+
Q(X)
---+
Q(B(H»
= C ®max B(H),
which shows that
IIIe
® u:
C
®min
X
---+
C
®max
B(H)II :::; II T i
l :
(BII) ®min X :::; ex(X).
---+
Q(X) II
17. Exactness
287
Thus we obtain (ii). Assume (ii). Let E c X be a finite-dimensional subspace. Then, by Corollary 12.6, for any c > 0, there is a factorization of UIE of the form E~l'vfn~B(H) with Ilvllcbllwllcb < A + c. Let E = v(E) C Mn, and let B: E ---> E be the map v but with range E. Then B- 1 : E ---> E is the restriction of Wi hence IIB- 11Icb ~ Ilwllcb and obviously IIBllcb ~ Ilvllcb. Thus we obtain dcb(E, E) ~ IIBllcbllB-11lcb ~ Ilvllcbllwllcb < A + c. Since c > 0 is arbitrary, this implies dsdE) ~ A. Thus we have shown that (ii) =} (iii) . • To complete the proof, we need the following useful lemma from §2.4, and we first recall the specific notation used there. Let I C B be a (closed twosided) ideal in a C* -algebra B. Let E be an operator space. As before, we denote for simplicity Q(E) = B ®min E. I®min E Then, if F is another operator space and if u: E ---> F is a c.b. map, we clearly have a c. b. map UQ: Q(E) ---> Q(F) naturally associated to I B ® u such that
Lemma 17.2. If u is a complete isometry, tIlen uQ is also one. This was already proved above as Lemma 2.4.8.
End ofthe Proof of Theorem 17.1. It remains to show (iii) =} (i). Assume (iii). We claim that, for any t in (BII) ® X, we have IltIIQ(x) ~ Alltllmin. Indeed, we can assume t E (B II) ® E with E c X finite-dimensional. By (iii) for any c > 0 there is an n and a subspace E C lIIn such that dcb(E, E) < A+c. Consider first an element tin (B II) ® E. We have by Lemma 17.2
but
hence
II~IQ(E) But then using dcb(E, E) < >.
+ c,
=
1I~I(B/I)®"';nE'
it is easy to see that this implies
Introduction to Operator Space Theory
288
and hence we obtain the announced claim. From this claim it follows that Tx is invertible and IIT.y-lll :::; A; hence ex(X) :::; A, which concludes the proof that (iii) :::} (i). Finally, (17.5) follows immediately from the equivalence of (ii) and (iii) . • Warning. Let I c Band q: B - B II be as before. For any C· -algebra A, the morphism q ® I A : B ®min A - (BII) ®min A is a metric surjection. Hence, for any x in (B II) ® A with II x II min < 1, there is an in B ®min A with IIxlimin < 1 such that q®IA(x) = x. However, in general such an x cannot be cllosen in the algebraic tensor product B ® A. Only when A is exact can we choose in A ® B! We emphasize this point because it is a potential source of confusion and is at the heart of "exactness."
x
x
By Lemma 17.2, if E
c F,
we have ex(E) :::; ex(F).
On the other hand, it is easy to see that ex(X) :::; sup{ ex(E) lEe X, dim E < 00 }, and consequently ex(X) = sup{ex(E) lEe X,dimE < oo}.
(17.6)
A careful look at the preceding proof shows that we have actually proved Theorem 17.3. An operator space X is exact iff
ds,oc{X)~fsup{ds,oc{E)
lEe X, dimE < oo} <
00.
~Moreover, for any finite-dimensional operator space E we llave ex(E) ds,oc{E) and
ex(X)
= sup{ds,oc{E) lEe X, dimE < oo}.
Remark 17.4. As observed by Kirchberg, if X is a C* -algebra and is exact as an operator space in the sense of Theorem 17.1, then necessarily ex(X) = 1. Indeed, the mapping Tx is a C* -representation; hence, as soon as it is injective it is isometric. Therefore we have Corollary 17.5. A C*-algebra A is exact iff ds,oc{A) = 1, or, equivalently, iff ds,oc{A) < 00. We will say that an operator space is A-exact if ex(X)(Le. ds,oc{X)) :::; A. With this terminology, an exact C*-algebra is automatically I-exact. Remark 17.6. In (17.1), let us make the specific choice I = K(.e 2 ) and B = B(.e2)' It c~n be shown that if (17.1) is exact with this specific choice of
17. Exactness
289
B and I, then it is exact for any choice (see Lemma 16.11), and this choice is equal to ex(X).
IIT,yl11
for
Another interesting choice of B and I is given by the consideration of ultraproducts. Let
and let Io C Boo be the ideal formed of all sequences x = (.Tn) in B')O such that IIxnll1\I" ~ 0 when n ~ 00. Again, for the exactness of X, it suffices to consider B = Boo and I = Io in the above definition, and IIT,y11l also coincides in this case with the exactness constant ex(X). Now let U be a nontrivial ultrafilter on N, and let Iu be ideal of all sequences x = (xn) in Boo such that limu IIxnllAl.. = O. The quotient C*-algebra Boo/Iu is nothing but the ultraproduct that we denoted earlier in §2.8 by DnEN Mn/U. As we will show in the next statement, this class of quotients also suffices to control the exactness of X. For this result, we advise the reader to read first Remark 2.8.3.
Theorem 17.7. Let X be an operator space. TIle following properties are equivalent.
(i) X is exact. (ii) For any family (Ei)iEI of operator spaces and any ultrafilter U on I. we llave a completely isomorpllic embedding:
(n iEI
E;/U)
Q9min
Xc
n(E
i. Q9min
X)/U.
iEI
(iii) Tllere is a constant A sucll tllat, for any (Ei)iEI as in (ii), tlle embedding considered in (ii) llas norm :S A. (iv) Same as (iii) bllt restricted to I = N and to families (En)nEN of finitedimensional spaces of tIle same arbitrary dimension. (v) Tllere is a constant A sucll tllat, for all finite-dimensional subspaces E C X, we llave dsK.{E) :S A. Proof. (i) => (ii): Assume (i). We can assume Ei C Ai for some C*-algebra Ai. By the injectivity of ultraproducts (cf. (2.8.2» and of the minimal tensor product, it suffices to prove (ii) when Ei = Ai' But, then we have DiEI A;/U = B/I with B = EBiEI Ai and I = {x I limu IIxili = OJ, so that (ii) follows from the definition of exactness. (ii) => (iii) is routine, and (iii) => (iv) is trivial. We now prove the key implication (iv) => (v).
290
Introduction to Operator Space Theory
Assume (iv). Let E c X be a fixed d-dimensional subspace. Let j: E--+ B(f2) be a complete isometry. Let Pn : B(f2) --+ lIfn be the standard projection that leaves eij invariant if i, j ~ n and takes it to zero otherwise. Let En = Pnj(E) and let Un: E --+ En be the restriction of Pnj to E. Clearly sup lIunlicb ~ 1, n
and Un is invertible for all n large enough. Moreover, x --+ (un(X»n::=:l obviously defines a complete isometry u: E --+ TIEn/U. Therefore, Ilu-1IlcB(llE,,/U,E) = 1. But u- 1 can be identified with a norm 1 element of (TIEn/U)* ®min E. By Lemma 2.8.1, the latter space is the same as (ilnE]II E~/U) ®min E. Hence, since we assume (iv), we can associate to u- 1 a sequence Vn with Vn E E~ ®min E = CB(En, E) with limu IIvnllcB(E,,,E) ~ A. Since (v n ) comes from u- 1 , we have necessarily
VxEE But since E is d-dimensional, this also implies that limu VnU n = Ie in the c.b. norm, or equivalently, that limu IIvn - u;;,-llicb = o. Finally, we have clearly dsdE) ~ Ilunllcbllu;;,-lllcb for all n, and hence
so we obtain (v). Finally, (v) =:> (i) is already included in Theorem 17.1.
•
We will also use the following refinement: Lemma 17.8. Let r be any set and let H = f 2 (r). For any subset 8 c r we view f2 (8) as a subspace of f2 (r). The following properties of an operator space Xc B(f 2 (r)) are equivalent:
(i) X is exact with ex(X) = l. (vi) For anye > 0 and any finite-dimensional subspace E c X, there is a finite subset 8 c r such that the natural completely contractive "restriction (or compression) map"
defines a complete isomorphism from E to Ps(E) II(PS)-l: Ps(E) -+ Ellcb
c
B(f2(8» such that
< 1 + e.
Proof. Clearly (vi) =:> dsdE) = 1, so (vi) =:> (i) follows from Theorem 17.1.
17. Exactness
291
Conversely, assume (i). Consider E as in (vi). Since dsdE) = 1 (by Theorem 17.1), we can find nand E c lI1n such that dcb(E, E) < 1 + c. We have then by Proposition 1.12
But if n remains fixed and if S tends to r along the net of finite sllbsets of r, it is easy to see (see Exercise 2.13.1) that II(Ps)-l: Ps(E) --> Elln --> 1. • We will now apply the ideas in Chapter 13 to exactness for C* -algebras. Let E be a finite-dimensional operator space, and let UE:
B I 1 0 m in E
E . E
B0min
-->
1
0
nlln
be the canonical isomorphism. As shown in (17.6), we know that for any exact operator space E
dsdE) = sup{lluell} = SUp{IIUEllcb},
(17.7)
where the supremum HillS over aU possible pairs (1, B) with 1 c B. (Actually, it suffices to consider 1 = K(£2) and B = B(f 2).) The point ofthe next result is that it suffices for the exactness of A to be able to embed (almost completely isometrically) the linear span of the unitary generators of A and the unit into K(£2) (or into a nuclear C*-algebra). Theorem 17.9. Let E C A be a closed subspace of a unital C*-algebra A. We assume that 1A E E and that E is the closed linear span of a family of unitary elements of A. Aforeover, we assume that E generates A (i.e., that the smallest C* -subalgebra of A containing E is A itself). We denote by [E* EJd the subspace spanned by all the products of the form XiY1X2Y2 ... xdYd with xt, Yl, ... , Xd, Yd E E. Then, the following are equivalent. (i) A is exact (i.e., dsdA) = 1, or, equivalently, dsdA) < 00). (ii) dsdE)
= 1.
(iii) limsuPd-+oo(dsd[E* EJd))l/d = 1.
Proof. Assume (ii). Let (I, B) be as above with B unital. By (17.7), if dslC(E) = 1, the unital *-homomorphism 11":
BII0 A
--->
B 0 m in A I0 m in A
becomes completely contractive when restricted to (BjI) 0 m in E. By Proposition 13.6, 11" extends to a continuous (contractive) *-homomorphism on
292
Introduction to Operator Space Theory
(BjI)®minA. Hence A is exact. More generally, assume (iii). Let T E BjI® E. To compute IIrr(T)II we use Ilrr(T)11 2d = IIrr((T*T)d)ll. Note that (T*T)d E BjI® [E*E]d, and hence by (17.7) Ilrr((T*T)d)1I ~ ds,d[E*E]d)II(T*T)dll, which yields Ilrr(T)1I ~ (ds,d[E*E]d))1/ 2d IITII, and by (17.7) again
ds,dE) ~ (ds,d[E* E]d))1/2d.
•
Thus (iii) implies (ii). The converses are clear by Theorem 17.1.
Corollary 17.10. For any free group G, the reduced C*-algebra C>:(G) is exact. Proof. Note that, for any D::::: 1, [E*E]D is the closed span of {>.(t) 2D}). Hence, by Remark 9.7.5, ds,d[E* E]D) ~ 4(2D)(2D
I It I ~
+ 1),
•
so that (iii) holds.
We now turn to the connection between exactness and the approximation property. It will be convenient to introduce several variants of the approximation property for a mapping.
Definitions 17.11. Let u: X
--+
Y be a c.b. mapping between operator
spaces. (i) We will say that u is strongly approximable if there is a net Ui: X --+ Y of finite rank maps such that, for any C*-algebra B, tbe maps IB ® Ui: B ®min X --+ B ®min Y converge pointwise to IB ® u. Note that it clearly suffices to consider B = B(£2)' (ii) We will say that U is exact if, for any ideal I c B in a C* -algebra B, the mapping _ B®min Y u: (B jI) ®min X ---+ I . Y (17.8) ®mm
that takes (b
+ I) ®
x to b ® u(x)
+I
®min
Y is bounded. We denote
ex(u) = sup{lItill}, wbere the supremum runs over all I and B. (iii) Let E be a finite-dimensional operator space and let u: E denote I's,du) = inf{lIvllcbllwllcbds.a::(F)},
(17.9)
--+
Y. We
where the infimum runs over all factorizations of u of the form E~ Equivalently, we can write I's.a::(u) = inf{lIvllcbllwllcb} and
F~Y.
17. Exactness
293
restrict F to be a subspace of K = K(f2). Now consider again X arbitrary and u: X --+ Y. We then define "Isdu) = supbsdulE) lEe X,dimE < oo}. Note that if 'u is the identity on X, we recover the previous definition, that is, we have "IsdIx) = dsdX). Then the proof of Theorem 17.1 can be adapted to this setting with very little changes (this was observed by Marius Junge). In particular, we have Theorem 17.12. Let u: X Then u is exact iff "Isdu) <
--+
00
Y be an operator between operator spaces. and we llave
ex(u) = "Isdu).
(17.10)
Definition. If the identity map on a C*-algebra (or an operator space) A is strongly approximable, we will say that A has tIle "strong operator space approximation property" (the strong OAP, in short). If A has the strong OAP, it is rather easy to verify that, for any ideal :1 c B in a C* -algebra B, any element in the kernel of the mapping B @min A --+ (B/:1) @min A must be approximable by elements in :1 @min A, and hence it actually lies in :1 @min A. Thus the strong OAP implies exactness for a C* -algebra.
More generally, we have Theorem 17.13. Let A be a C* -algebra. Let X, Y be operator spaces. Then, for any c.b. map w: X --+ A and any strongly approximable map v: A --+ Y, the composition vw is exact and satisfies (17.11)
Proof. Let:1 c B be as before, and let q: B @min A --+ (B/:1) @min A be the morphism associated to the quotient map from B onto B/:1. We denote for any operator space A R(A)
=
B@min A . ker(q)
Note that, by definition, A is exact iff we have R(A) = Q(A) for any :1 and B (here Q(A) = (B @min A)/:1 @min A, as defined in Lemma 17.2). Now assume first that Vi: A --+ Y is a finite rank map. Consider Vi = IB @Vi: B@min A --+ B ®min Y. Then, for any () in ker(q), Vi«(}) E :J ®min Y.
Introduction to Operator Space Theory
294
(Indeed, it is easy to check that, for anye in A*, we have (1B®e)(ker(q» c .:J.) Now assume that Vi tends pointwise to B®minA -4 B®min Y. This certainly implies v( 0) E .:J ®min Y. Therefore any strongly approximable map v: A -4 Y defines a mapping v: B ®min A -4 B ®min Y such that v(ker(q)) c .:J ®min Y. R(A) -4 Q(Y) that (since Passing to the quotient spaces, we obtain a map IIvll ::; Ilvllcb) satisfies Ilvll ::; Ilvllcb. But now, since A is a C* -algebra, the morphism q is onto and we have necessarily
v:
v:
(B/.:J)
®min
A = R(A).
On the other hand, for any c.b. map w: X IIIB/.1
®
w: B/.:J ®min X
-4
-4
A we clearly have
B/.:J ®min Allcb::; Ilwllcb.
Therefore, we conclude that, if U = vw, the map (17.8) can be factorized as 0 (1B/.1 ® w), and hence it has norm::; IIvllcbllwllcb. By (17.9) and (17.10), this completes the proof. •
v
Applying this to
U
= v = w = lA, we obtain:
Corollary 17.14. If a C*-algebra A has the strong OAP, then A is exact. More generally, as observed in [JIJ, we have Corollary 17.15. A C* -algebra A c B( H) is exact iff the inclusion A -4 B(H) can be approximated pointwise by a net of finite rank maps Ui: A-4 B(H) with SUPi IIUillcb < 00. Proof. The "only if" part follows e.g. from Theorem 17.1 and the factorization of maximizing maps described in Corollary 12.6. Conversely, assume that there is a net (ud as in Corollary 17.15. Of course, each Ui: A -4 B(H) is strongly approximable (since Ui has finite rank); hence, by Theorem 17.13, we have 'Y8dUi) ::; IIUi IIcb' so that SUPi sup 'Y8d Ui) < 00. Clearly this implies that A is exact by a perturbation argument (see Lemma 2.13.2). • Corollary 17.16. Let E be a finite-dimensional subspace of an arbitrary C*-algebra A. Let 'x(E, A) = inf{llPllcb}, where the infimum runs over all possible linear projections P from A onto E. Then
dsdE) ::; 'x(E, A). Proof. Let P: A -4 E be a projection onto E and let j: E inclusion map. We then have IE = Pj. Hence, by (17.11),
(17.12) -4
A be the
dsdE) = 'YsdE) ~ IlPllcblljllcb = IlPllcb, whence (17.12).
•
1 7. Exactness
295
Remark 17.17. (i) By [DCHj, for any noncom mutative free group IF the reduced C*-algebra C~(IF) (as defined in §9.7) has the strong OAP. Hence, this is another proof that it is exact. (ii) In sharp contrast, the full C*-algebra C*(IF) (considered in Chapters 8 and 13) is not exact. This was first proved in [Wa4]. From the (more recent) viewpoint of operator spaces, this can be seen as a consequence of the following estimate from [PlOj. Let lF n (resp. lFoe ) be the free group with n (resp. countably many) generators'(n = 1,2, ... ), and let U}, ... ,Un be the unitary operators in C*(lFn ) associated to the free generators. Let E[j = span(U}, ... , Un). Then we have dsdE[j) ~ n(2Jn=-1)-1, so that, if n ~ 3 this number is > 1; hence the containing C* -algebra C* (IF n) cannot be exact (a variant shows that n ~ 2 suffices for this). Then the fact that C* (IF) is not exact follows from the following two well-known facts: First, any noncommutative free group IF contains lF n and lFoo as subgroups (cf., e.g., [FTP, p. 15]). Second, for any discrete group r and any subgroup r' c r, the full C*-algebra c*(r') appears naturally embedded as a C*-subalgebra of c*(r) (see Proposition 8.8).
Remark. As mentioned after (17.2), the exactness of a C*-algebra can be summarized by saying that for any I, B the kernel of the map q: B Q9min A ----> (BII) Q9min A coincides with I Q9min A. This is what is called a "slice map property." The slice map property for C* -algebras was introduced in [T03] and further studied in [Wa4, Kr, Ki7]. More generally, let X, Y be operator spaces and let E c X, FeY be closed subspaces. The Fubini-product of E, F in X Q9min Y is the subspace
F(E, F) eX formed of all the elements t in X (x*
Q9
Q9min
Iy)(t) E F
Q9min
Y
Y such that Vx* E X* and Vy* E Y*,
and
(Ix
Q9
x*)(t) E E.
Then, the slice map problem (see [Wa4, Kr, Ki7]) is to decide when we have
F(E, F)
=E
Q9min
F.
For instance, for the map q: B Q9min A ----> (BII) Q9min A, we have ker(q) = F(I, A). Thus, by definition, A is exact iff F(I, A) = I Q9min A for any Band any ideal I C B, or iff this holds for I = K and B = B(£2). We will say that an operator space X has the slice map property (Sl\IP in short) with respect to FeY if F(X, F) = X Q9min F. We will say that X has the Y-SMP if this holds for all (closed) subspaces FeY. The SMP is closely related to the OAP. Here is a summary of the main known facts:
296
Introduction to Operator Space Theory
(i) The Co-SMP is equivalent to the AP (this essentially goes back to Grothendieck). (ii) The K-SMP is equivalent to the OAP ([Kr]). (iii) The B(f2)-SMP is equivalent to the strong OAP ([Kr]).
Remark. Haagerup and Kraus ([HK]) and Kirchberg ([Ki7]) independently proved that a C* -algebra A has the strong OAP if (and only if) it is exact and has the OAP (or actually iff it is locally reflexive - in the sense of the next chapter - and has the OAP). Moreover ([Ki7]) the OAP is stable under extensions; that is, if I and BII both have the OAP, then so does B (this can be proved by using (ii) and the exactness of K). However, by [Ki2] there is an example of a B such that I ~ K, B II has t.he strong OAP (actually the metric CBAP), but B is not exact. This shows that neither exactness nor the strong OAP is stable under extensions.
Haagerup an~ Kraus ([HK]) say that a discrete group G has the AP if has the OAP and the approximating net can be chosen formed of finitely supported multipliers (see Remark 8.4 for background on multipliers). They show that G has the AP iff C~(G) has the OAP. Moreover, for C~(G) (discrete case), the OAP implies automatically the strong OAP. They also. study the AP for general locally compact groups. C~ (G)
Up to now, the Gromov group 9 that will be discussed in Remark 17.22 is the only known example of a discrete group failing the AP. It is unknown whether there is an exact C* -algebra failing the OAP (or even the AP!), but there is a prime suspect, namely, CA(8L(3,Z». Indeed, since 8L(3, Z) is a lattice in a (connected) Lie group, by a result due to Connes [Col], CA(8L(3,Z» embeds in a nuclear C*-algebra; hence it is exact.. However, by [H8], it fails the CBAP. This motivates the conjecture (formulated in [HK]) that it fails the OAP. Note that if G is the semi-direct product of 8L(2, Z) with Z2 (for the standard action), then C~(G) has the OAP ([HK]) but fails the CBAP ([H8]). We refer to [CoH] for important results on approximation properties using c.b. multipliers. Corollary 17.5 means that exactness is characterized by a "local embeddability" into the nuclear C* -algebra K. Actually, Kirchberg [Ki5, KiP] (see also [AnI]) recently obtained a considerably deeper result, which yields a global embedding: Theorem 17.1S. Any separable exact C*-algebra A embeds (as a C*subalgebra) in a separable nuclear C*-algebra B. Actually, we can take for B the Cuntz algebra 02' Moreover, if A is nuclear, there is an embedding A C O 2 for which there is a completely contractive (and completely positive) projection P: O 2 -+ A.
1 7. Exactness
297
Curiously, it remains open whether a nonseparable exact C* -algebra embeds in a nuclear one. Remark. It is clear by Theorem 17.3 that exactness passes to subspaces and is stable under the minimal tensor product. It is also true [Kil] that if a C* -algebra A is exact, then all its quotients AII are also exact.. Surprisingly, however, the only known proofs are extremely difficult. One way or the other, they are aU related to Kirchberg's embedding Theorem 17.18 (or previolls versions of it in [Kil]). The derivation is as follows: To show that AI I is exact if A itself is exact, it is easy to reduce to the case when A is separable. But then, by Theorem 17.18, AII is a quotient of a subalgebra of a nuclear C*-algebra, and such algebras are exact by Proposition 18.19 in the next chapter. In the case A = CA(G), Kirchberg's embedding theorem can be recovered as a corollary of the following striking result due to Ozawa [Oz2] (inspired by the previous paper [GK]). Let D(G) c B(€2(G» be the (commutative) C*-algebra of all diagonal operators. Note that obviously D(G) ':::' €oo(G). We will denote by UC*(G) the C*-subalgebra of B(€2(G» generated by C~(G) and D(G). Theorem 17.19. ([Oz2]) Let G be allY discrete group. The following are equivalent.
(i) C A(G) is exact. (ii) UC*(G) is Iluclear. For the proof we will need several results. The first statement provides us with a description of the "typical" operators in UC*(G). Lemma 17.20. Let
v = span{A(t)D I t E G, DE D(G)}. (i) Then V is a dense *-subalgebra in UC*(G). (ii)An)' operator v in V can be uniquely written as a fillite sum of the form
with De E D(G), where T eGis a finite subset. (iii) An operator v in B(f2(G» is in V iff there is a finite subset T SUell that the associated matrix {v(s, t) Is, t E G} satisfies v(s,t)
=0
v s, t
such tllat
SCI
fj. T.
C
G
298
Introduction to Operator Space Theory
Proof. (i) follows from the elementary fact that ,x(t)D(G),x(t)-1 C D(G) for any t in G. (ii) Clearly any v in V can be written as a finite sum v = LOET ,x(O)Do with Do E D(G). Then the associated matrix v(s, t) defined by v(s, t) = (v8t , 88 ) satisfies
v(Os, s) = Do(s, s), which shows that Do and T are uniquely determined by v.
(iii) Assume v(s, t) = 0 if srI f/. T with T c G finite. Then, if Do(s, s) = v(Os, s) for all 0 in T, we have v = LOET ,x(O)Do, so that v E V. The converse is clear by (ii).
•
Definition. Let G be any set. We will say that a function cp: G x G ---+ C is a positive definite kernel if, for any n and any tl, ... ,tn E G, the matrix (CP(ti' tj)) is positive definite, that is,
Remark. It is easy to see that this holds iff there is a Hilbert space H and a function x: G ---+ H such that cp( s, t) = (x( s), x( t)). Indeed, if cp is a positive definite kernel, we may equip the space of all finitely supported functions a: G ---+ C with the scalar product (a, (3) = L a(s)(3(t)cp(s, t); after passing to the quotient by the kernel {a I (a, a) = O} and completing, we obtain a Hilbert space H, and we have (8 8 ,8t ) = cp(s, t). The converse is obvious. We now come to the key point: Lemma 17.21. ([Oz2]) If C~ (G) is exact, then for any € > 0 and any finite subset S C G, there is a finite subset T C G and a positive definite kernel cp: G x G ---+ C such that Icp(s, t) -
11::; €
if srI E S
and
cp(s,t) = 0 if st- I ¢. T. Proof. Let E = span{A(s) I s E S}. Let j: E ---+ B(f2(G)) be the inclusion map. By Lemma 17.8, there is a finite subset S C G such that the compression map Ps: E ---+ E = Ps(E) is an isomorphism such that IIPSI : E ---+ Ellcb < 1 +€. Let u = j(PS)-I: E ---+ B(f2(G)). We have lIullcb < 1 + €. By the extension property (Corollary 1.7), u admits an extension w: B(f2(S)) ---+ B(f2(G)) with IIwllcb < 1+€. Recall see (11.5)) that Ilwllcb = IIwlldec. Hence, by Corollary 12.5, the tensor u E (E)* ® B(f2(G)) associated to u satisfies
t
17. Exactness
299
8(u) < 1 + c. We will now apply Lemma 12.9 to u: it -+ B(C2 (G)). By enlarging S if necessary we may clearly assume that S is symmetric and e E S. Then E, and hence also it, is unital and self-adjoint and u is a unital selfadjoint map. By Lemma 12.9, there is a c.p. map <1>: B(C 2(8)) -+ B(E2(G)) such that \:IxEE (17.13) 1I(x) - u(x)1I :::; cllxll· We then define Using the complete positivity of the composition Ps it is easy to check that cP is a positive definite kernel: Given t1,' . . ,tn in G we have (>,(tjtj1)) ~ 0;
hence (Ps(>,(ti t j 1))) ~ 0, which implies (cp(tj, tj)) ~ O. Moreover, if sC 1 E S, we have >'(st- 1) E E, and hence Ps(>,(sC 1)) E and uPs(>,(sC1)) = >,(sr 1 ). Hence, by (17.13),
which implies (since >,(sC 1)8t
it
= 88 )
11 :::; c.
Icp(s, t) -
Finally, let T = 88- 1 , that is, T = {ab- 1 I a,b E 8}. Then Ps(>,(sC 1)) =I- 0 iff there are a, bin 8 such that (>,(sr 1 )8b, 80 ) =I- 0, or, equivalently, iff 8t- 1 E T. Thus we obtain a finite set T such that
IIM",II = IIM",(I)1I = sup l
By Lemma 17.20, M",(B(f 2 (G))
c
UC*(G).
Let I be the set of pairs Q = (S, c) with S finite, e E S, and c > 0, directed so that Q -+ 00 corresponds to "S tends to G and c tends to 0." Let CPo. be the kernel associated to Q = (S, c) as in Lemma 17.21, and let uo.: UC*(G)
-+
UC*(G)
300
Introduction to Operator Space Theory
be the restriction of AI",,, to UC*(G). Note that Un: is c.p. and Ilun:11 ::; IIM",o II ::; 1 + e. We claim that v x E UC*(G) Ilun:x - xii ---4 O. Indeed, since lIun:1I ---4 1, it suffices to check this (by Lemma 17.20) for all x of the form x = )..(t)D with D E D(G). Then we find
Un: X - x = Ls D(s, s)(
Ilun:x - xii::; sup ID(s, s)II
and if a is chosen large enough, we have (tS)(S)-1 = t E S for all Sj hence suPs I
and since Un: is c.p., lIun:lldec = Ilun:11 ::; 1+e, so we conclude that A®min B = A®max B . Finally, to check that Un: ® IB is bounded from A ®min B to A ®max B it suffices to rewrite Un: as a finite sum of maps of the form A~C~A with C nuclear, v c.b., and with w such that w ® IB is bounded from C ®max B to C ®max A (which holds if w is decomposabl~). We will show this with C = D(G). Being commutative, this is a nuclear C*-algebra (see Exercise 11.7). Let P: B(C2(G» ---4 D(G) be the standard contractive projection taking (a(s, t»s,t to (a(s, t)l{s=t}). Note that P is c.p. by Exercise 1.5 (since P can be viewed as Schur multiplication by (8 s ,8t ). Then, if a = (S,e) and if T is a finite set for which
Un: X = L
),,(O)P(),,(O)-IUn:x).
9ET
Thus Un: can be written as announced, since v: x ---4 P(),,(O)-IUn:X) is a c.b. map into D(G) and w: y ---4 )"(O)y is decomposable from D(G) into UC*(G). (More directly, assuming B unital, w ® I B : C ®max B ---4 A ®max B is clearly contractive since it is the restriction to C ®max B of the left multiplication by ),,(0) ® 1 in A ®max B.) • This completes the proof that UC*(G) is nuclear. Remark 17.22. (i) It was a long-standing open problem whether C~(G) is exact for any discrete group G. (The groups for which it is true are called
17. Exactness
301
exact in [KiW].) However, this was disproved very recently by N. Ozawa [Oz2] (see also [An2]): His result (the above Lemma 17.21) shows that, for a special group 9 constructed previously by Gromov [Gro] using very delicate arguments (not yet fully understood at the time of this writing), CAW) cannot be exact and hence is a counterexample to the above problem. A fortiori, by [HK], CAW) fails the OAP; in other words, 9 fails the AP. (ii) On the other hand, it remains an outstanding open problem whether the full C* -algebra C* (G) is not exact for any nonamenable discrete group
G. The proof of Kirchberg's embedding (Theorem 17.18) is quite difficult and beyond the scope of this book. However, there is a much simpler operator space version, which we can fully prove, following [EOR]. Fix A ~ 1. We will say that an operator space X is A-nuclear if the identity on X can be approximated pointwise by a net of finite rank maps of the form
with sUPa Ilvallcbliwalicb ~ 1 and na < 00. When A assume that Va, We> are complete contractions.
= 1, we may as well
Note that, if X is a C* -algebra, it is nuclear iff it is I-nuclear in the preceding sense (see Theorem 11.5) and actually ([PI, p. 35]) iff it is A-nuclear for some A. Recall we say that X is A-exact if sup{ds.dE) lEe X,dim(E) < oo} ~ A.
We then have Theorem 17.23. ({EOR]) Every i-exact separable operator space X embeds completely isometrically in some i-nuclear separable operator space. Proof. ([EOR]) Let {En} be an increasing sequence of finite-dimensional subspaces of X with dense union. Let c(n) > 0 be chosen so that Lc(n) < 00. We claim that there is an increasing sequence N(l) < N(2) < ... of natural numbers and maps as in the following diagram: E1
i1
C
C
!
MN(l)
En
in it
'---+
;n-1
'---+
such that (i) in is a complete isometry,
C
!
l1fN (n)
En+! in+l
in
'---+
C
!
lIfN (n+l)
jn+l '---+
302
(ii) (iii)
Introduction to Operator Space Theory
Ilinllcb:::; 1 and Ili;~k.(E,)lcb :::; 1 + c(n), Ilin+1l En - ininllcb :::; c(n).
We will prove this by induction on n. Since dsdEl) = 1, we can find N(I) and an embedding il: El --+ lIfN(l) such that Ilidlcb :::; 1 and
Now suppose that in: En --+ MN(n) is given satisfying (ii) above. We will prolong the above diagram to the right by constructing N(n + 1), in+!: En+! --+ AIN(n+!), and in: AIN(n) --+ MN(n+!) satisfying the required conditions. Since dsd E n+!) = 1, we can find N ~ 1 and a complete contraction u: En+! --+ MN with IIU~~E,,+dllcb < 1 + c(n + 1). By the injectivity of MN(n), the map in: En --+ lI,fN(n) admits an extension i:: E n+1 --+ AIN(n) with Ili:llcb :::; 1. By the injectivity of M N , the map UIEni;;l: in(En) --+ MN admits an extension v: MN(n) --+ MN with Ilvllcb :::; 1 + c(n). Note that vin = UIE,,· Then let N(n + 1) = N(n) + N, and let i n +1 : En+! --+ lI1 N(n+l) and in: MN(n) --+ MN(n+l) be defined by the following block diagonal sums:
in+!(e) = in(e) EBu(e) (e E En+d in(x) = x EB (1 + c(n))-lv(x) (x
E
MN(n»)'
Then it is easy to check the desired conditions. This establishes our claim. Clearly we may now view the maps in: lIfN(n) --+ lIIN(n+!) as if they were inclusions (of operator spaces, but not of algebras!) and form the direct limit Y = UMN(n) (Le., the inductive limit of this system). Obviously the resulting o.s. Y is I-nuclear by construction. Moreover, the mapping i: UEn --+ Y defined by i(x) = lim n --+ oo in(x) is well defined and completely isometric. By density, it extends to a completely isometric embedding of X into Y. • Exercise Exercise 17.1. Prove that a C* -algebra is nuclear iff it is both exact and WEP.
Chapter 18. Local Reflexivity
Basic properties. In Banach space theory, the principle of local 7'eflexivity ([LiRJ) says that every Banach space X has the following propert.y, called "local reflexivity": B(E, X)** = B(E, X**) (isometrically) for any finitedimensional Banach space E. In sharp contrast, the O.s. analog is not universally true, and local reflexivity has turned out to be a very important property. Definition 18.1. Let A 2: 1. An operator space X is called A-locally reflexive if, for any finite-dimensional o.s. E, the llaturallinear isomorpllis1l1 GB(E, X**)
--->
GB(E, X)**
has norm ::; A. It is easy to check that the inverse map GB(E, X)** ---> C B(E, X**) always has norm L Hence X is I-locally reflexive iff, for any finite-dimensional E, we have an isometric identity GB(E, X)** = GB(E, X**), or equivalently, for any finite-dimensional o.s. F we have (F
®min
X)** = F
®min
X**.
Replacing E* and F by 1\/n(E*) and l\!1I(F), it is easy to see that the preceding identities are actually completely isometric, but we will not need this. If Y c X (isometrically), we have Y** C X** (isometrically). Hence, since GB(E, Y) c GB(E, X) (isometrically) if Y c X (completely isometrically), it is easy to check (just like for reflexivity) that Y C X is A-locally reflexive if X is. On the other hand, in general, the quotient XjY is not A-locally reflexive: Indeed, any separable o.s. is a quotient of SI that is locally reflexive by Theorem 18.7. Nevertheless, the quotient of a locally reflexive G* -algebra by an ideal inherits local reflexivity (see Exercise 18.4). Remark 18.2. By (2.5.1), we know that, for any operator space X:
(1\ l n ®min X)** = 1\/11
®min
X**
(isometrically).
Hence, for any subspace S C 1\ln we have (S
®min
X)**
= S ®min
X**
(isometrically),
or equivalently GB(S*, X)** = GB(S*, X**)
(isometrically).
The A-local reflexivity of X is used in the following equivalent reformulation:
304
Introduction to Operator Space Theory
Proposition 18.3. An operator space X is A-locally reflexive iff, for any finite-dimensional operator space E and any mapping u: E -+ X** witll lIulicb ::; 1, there is a net Ui in CB(E, X) with lIu;lIcb ::; A SUc11 that, for any x in E, Ui(X) -+ u(x) for the a(X**,X*)-topology. Proof. That this is indeed equivalent follows easily from the classical density of the unit ball of a Banach space B in the unit ball of its bidual B** for the topology a(B**,B*) applied to B = CB(E,X) = E* ®min X. • In this formulation, it is easy to see (replacing E by u(E» that it suffices to consider the case when E c X** and the map U to be approximated is the inclusion map E -+ Xu. In the Banach space case, when E c X**, a stronger principle holds: The maps Ua: E -+ X approximating the inclusion E c X** can be chosen so that lIuall ::; 1 and lIu~I~,,(E)11 ~ 1 when Q -+ 00 (see Exercise 18.2). The following consequence for operator spaces was observed in [GHJ. Here, for any map u: E -+ F, we denote lIulin = III ® u: Mn(E) -+ Mn(F) II· Proposition 18.4. Let X be an arbitrary operator space. Let E c X** be a finite-dimensional subspace. There is a net of maps Ua: E -+ X tending to the inclusion map E c X** in the point-a(X**, X*) topology and, moreover, such that, for each fixed n 2:: 1, lIu ali n ::; 1 and lIu;;lluo(E)lIn -+ 1. Proof. Consider the inclusion Mn(E) C Mn(X**). Recall that, by (2.1.4), Mn(X**) = Mn(X)**. By the local reflexivity principle for Banach spaces, there is a net of contractive maps Va: Mn(E) -+ Mn(X) such that Va(e) tends a(Mn(X)**, Mn(X)*) to e for any e in Mn(E). Let Va be defined as in Exercise 18.3. Then, by Exercise 18.3, we have Va = I ® U a and lIu a lin ::; 1. Clearly Va(e) still tends to e for a(Mn(X)**, Mn(X)*). Therefore, by Exercise 18.2 applied to Va, we find that Va is almost isometric when Q is large enough, and hence II U;; \ . " (E) lin -+ 1. • Proposition 18.5. ([EaR]) An operator space X is A-locally reflexive iff the same is true for every separable subspace. Proof. Since local reflexivity passes to subspaces, we only need to prove the "if" part. Suppose that each separable subspace of X is A-locally reflexive. Let E C X** be a finite-dimensional subspace. Fix a finite subset 6, ... , ~k in X*. By Proposition 18.2 there is a sequence of maps Un: E -+ X such that II Un II n ::; 1 such that lim(~j,un(e) n
- e) = 0
for each
j = 1, ... ,k.
18. Local Reflexivity
305
Let Xl be a separable subspace of X containing un(E) for all n ~ 1. Let U be a free ultrafilter on N. We define a mapping v: E ---+ Xi* by setting '<:Ie E E
v(e) = lim Un (e)
u
(limit in a(Xi*,Xi) sense).
Note that (f"j, v(e)) = (f"j, e) for each j = 1, ... , k. Clearly IIvllcb ~ 1 (since Ilu n lin ~ 1 for each n). Then, by the local reflexivity of the separable subspace Xl, there is a net of maps Va: E ---+ Xl such that Ilvalicb ~ 1 and '<:Ie E E
(f"j, v,,(e))
---+
(f"j' v(e)) = (f"j, e).
•
This shows that X is locally reflexive.
Remark 18.6. As pointed out to the author by Kirchberg, it is not hard to show that any Ll-space (in the usual, commutative sense) over a measure space (fl, It) is locally reflexive. However, the question whether the predual of a von Neumann algebra (for instance, the space Sl of all trace class operators on f2 (Le., the predual of B(f2))) is locally reflexive remained open for a while, but this was recently proved in [EJR] (see also [JI] and [JLl\I]). We include the simplified proof from [JLM]. Theorem 18.7. Tlle predua1 111* of allY VOll Neumal1n algebra 111 is locally reflexive. Proof. Let X = 111* so that X** = 111*. \Ve will use the criterion in Proposition 18.3. Let E be a finite-dimensional opf'rator space and let u: E ---+ AI* be a complete contraction. We may assume E* c B(H) (complf'tf'ly isometrically). Let i: E* c B(H) be the inclusion map. Let T: 1Il ---+ E* be the restriction of u* to M. Note that iT: 111 ---+ B(H) is a finite rank map with IliTllcb ~ 1. Hence, by Theorem 12.7, for any c > 0 there is, for some n, a factorization of iT of the form 111 ~1IIn~B(H) with IIvllcb = 1, Ilwllcb < 1 + c. Let S = v(M) C lIln . Since w(S) C E*, we have, by restriction, a factorization of T of the form T: M~S~E*
with Ilvlllcb = 1, Ilwlllcb < 1 + c. Taking adjoints, since u factorization of u of the form u:
=
T*, we find a
E~S*~Af*.
But now, by Remark 18.2, vi can be approximated in the point-a(AI*, AI) sense by a net of complete contractions '1/),,: S* ---+ AI*. Then, letting u" = t/J" wi, we find an approximating net as in Proposition 18.3, but with sup Ilu"lIcb ~ 1 + c. Since c is arbitrarily small, this implies that 1Il* is I-locally reflexive. •
A conjecture on local reflexivity and OLLP. We now return to the OLLP, which we encountered in Chapter 16.
306
Introduction to Operator Space Theory
Proposition 18.8. Let A 2: 1. Any operator space X for which X** has the A-OLLP is A-locally reflexive and has the A-OLLP. Proof. Fix c > O. Let E c X** be a finite-dimensional subspace. By Theorem 16.6, the inclusion E --+ X** admits a factorization E~G*~X** with IIwllcb = 1, IIvllcb < A + c, and G C JfN for some N. Since MN(X**) = MN(X)** (isometrically) (see (2.5.1)), we also have G ®min X** = (G ®min X)**; hence W can be approximated in the point-o-(X**, X*) sense by maps Wa: G* --+ X with IIwallcb ::; 1. Then the maps U a = WaV: E --+ X approximate in the same sense the inclusion E --+ X** and sup lIuallcb ::; A + c. Thus X is (A + c)-locally reflexive for any c > 0, and hence (by an elementary argument) it is A-locally reflexive. To prove that X has the A-OLLP, assume that E eX. Then we have, for any e in E, U a (e) - e --+ 0 weakly in X; hence, passing to convex hulls, we can obtain a net such that U a (e) - e --+ 0 strongly for anye in E. By a simple perturbation (see §2.13) argument, this gives us that X satisfies (iii) in Theorem 16.6, and hence X has the A-OLLP. • We will now show the equivalence of a number of interesting conjectures formulated by Ozawa [Oz3, Oz6] and closely related to previous work by Oikhberg [03]. Definition 18.9. We will say that an operator space X is submaximal if it embeds completely isometrically into a maximal o.s. Y. If X is separable and submaximal, we may as well take Y = max(B(£2)) (or, in the nonseparable case, Y = max(B(H)) for some H). Indeed, if X is separable, we can assume X C B(£2)' Then let X be the o.s. obtained by inducing on X the o.s. structure of max(B(£2))' Now assume Xc Y with Y maximal. By the injectivity of B(£2), there is a complete contraction T: Y --+ B(£2) extending the inclusion X --+ B(£2)' Since Y is maximal, liT: Y --+ max(B(£2))lIcb::; 1; hence, restricting to X, we find liT: X --+ Xllcb ::; 1. On the other hand, since max(B(£2)) --+ B(£2) is trivially completely contractive, we also have liT-I: X --+ Xllcb ::; 1. Hence, we conclude that:
X submaximal <:=::::}X
= X (completely isometrically).
Any maximal space X is (essentially by definition; see Proposition 3.3) completely isometric to a quotient of £1 (r) for some set r. If X is finite-dimensional, it is easy by a compactness argument to see that we can achieve this with a finite set r, provided we replace "completely isometric" by (1 + c)-completely isometric (see Exercise 2.13.2). Similarly, submaximal spaces can be identified with subquotients of £1 (r) with r an arbitrary set. However, in sharp contrast with the preceding case, it turns out to be apparently a quite delicate question of whether the corresponding finite-dimensional assertion is valid. This motivates the following
18. Local Reflexivity
307
Definition 18.10. Let A ~ 1. An operator space X will be called Ahypermaximal if, for any finite-dimensional subspace E c X, there is a finitedimensionalnormed space F and a factorization E~max(F)~X of the inclusion map E c X SUcll that Ilvllcbllwllcb < A + c. We will say that X is A-maximal if it is completely isomorphic to a maximal space and the c.b. norm of the identity X ---+ max(X) is ::; A. We claim that X A-hypermaximal implies X, A-maximal. Indeed, if X is as in Definition 18.10, any bounded u: X ---+ B(H) satisfies, for any finitedimensional subspace E c X, IlulEllcb = IIuwvllcb ::; Iluwllcbllvllcb, and hence (since Iluwllcb = Iluwll) a fortiori IIulEllcb ::; Iluwllllvllcb ::; Ilullllwllllvllcb ::; (A + c)lluli. Thus we must have Ilulicb ::; Allull, and we conclude that X is A-maximal.
Remark 18.11. By a perturbation argument it is easy to verify that, if a maximal o.s. X has the metric approximation property (resp. the A-BAP), then X is 1-hypermaximal (resp. A-hypermaximal). Remark 18.12. The reader should observe the analogy between the preceding definition and that of spaces with A-OLLP (see (i)~ (iii) in Theorem 16.6): If G = max( F) with dim F < 00, then, for each c > 0, G* (1 + c)-embeds into f! for some N. Thus, in some sense, the A-OLLP is the noncommutative analog of the preceding definition. Theorem 18.13. ((Oz3]) Fix A ~ 1. The following conjectures are equivalent: (i) Every maximal o.s. lIas the A-OLLP. (ii) Every maximal o.s. is A-locally reflexive. (ii)' Every submaximal o.s. is A-locally reflexive. (iii) max(B(f2)) is A-locally reflexive.
(iv) Any maximal o.s. is A-l1Ypermaximal. (v) For any separable Banach space X, any bounded map u: X ---+ Bjte admits a bounded lifting u with Ilull ::; Allull (recall the notation B = B(f2), te = K(f2)). Proof. (i) ~ (ii). If X is maximal, then (by Exercise 3.3) X** is maximal; so, by Proposition 18.8, (i) ~ (ii). (ii) {:} (ii)' is obvious since local reflexivity passes to subspaces. (ii)
~
(iii) ~ it embeds assume X subspace.
(iii) is trivial. (ii). By the preceding remarks, if X is separable and submaximal, into max(B(f2»; hence it is A-locally reflexive if (iii) holds. Now possibly nonseparable. Let Xl C X be an arbitrary separable By Exercise 3.8, if X is maximal, there is separable subspace X2
Introduction to Operator Space Theory
308
with Xl C X 2 C X that is also maximal and hence A-locally reflexive by the separable case. A fortiori, its subspace X I is A-locally reflexive. Then since local reflexivity is separably determined (see Proposition 18.5), we conclude that X itself must be A-locally reflexive. (ii) ::::} (iv). This uses a construction due to W. B. Johnson [Jo]. Consider a finite-dimensional subspace E C max(X). Fix € > O. Assume X separable for simplicity. Let {En} be an increasing sequence of finite-dimensional subspaces of X such that E C EI and UEn is dense in X. Let Y = el({En}). We claim that the identity on X** admits a factorization through y*" of the form I }(.... X** - -J+ Y** -P" - + X** ,
where J and P: Y
--+
X are complete contractions.
There is a natural metric surjection P: Y --+ X defined by P«x n )) Note that Y* = eoo({E~}). Let U be a free ultrafilter on N. Let us denote by J: X** --+ y** the mapping defined by
2:x n .
(J(x),e) = (x, [limen]),
u
where [limu en] E X* is first defined as a pointwise limit on the union of the spaces {En} and then extended by density to the whole of X. Clearly IIJII ~ 1. Note that P** J = Ix'" Let j = JIE: E --+ Y**, so that, denoting by iE: E --+ X** the inclusion map, we have P**j = iE. Since IIJII ~ 1, we have IIJ: max(X**) --+ max(Y)**lIcb ~ 1, and hence a fortiori Iii: E
--+
max(Y)**llcb ~ 1.
If we assume (ii), max(Y) is A-locally reflexive; hence there is a net of maps Uo: E --+ max(Y) with Iluolicb ~ A tending point-a(Y**, Y) to j. A fortiori, P**u o --+ P**j = iE in the point-a(X**,X*) topology. Hence, since P**u o = PUo,
we have Ve E E
Puo(e)
--+
e
(18.1)
with respect to a(X**, X*). But now, since both Puo(e) and e are in X, the convergence in (18.1) holds in the weak topology of X. Passing to convex hulls, we can modify the U a and ensure that, actually, the convergence (18.1) holds in norm. This gives us a (strong) approximate factorization of iE as follows: E~ max(Y)~ max(X) with lIuallcb ~ A. Now, by Remark 18.11, max(Y) is 1-hypermaximal, so we immediately deduce from this factorization that max(X) is also Ahypermaximal. This completes the proof that (ii) ::::} (iv).
18. Local Reflexivity
309
Since it is clear that a A-hypermaximal space has the A-OLLP (see Remark 18.12), we have (iv) => (i) and (i)-(iv) are equivalent. By Theorem 16.10, (v) is equivalent to the assertion that every separable maximal o.s. has the A-OLLP. Now let X be a nonseparable maximal o.s. and let E c X be a finite-dimensional sub~pace. By Exercise 3.8 there is Xl separable and maximal such that E c Xl eX. Thus, if we know that Xl always has the A-OLLP, it immediately follows (by definition of the OLLP) that X has the A-OLLP. This shows that (i) {::} (v). • It is a long-standing open question whether any ideal I c B in a separable C* -algebra is automatically complemented, that is, whether there is a bounded linear projection P: B ----> I. Thus the next statement constitutes a strong motivation to disprove the above conjectures. Proposition 18.14. ([Oz6]) If, for all values of A, the equivalent conjectures in Theorem 18.13 fail to be tme, there is an embedding K C B a.9 an ideal in a separable ()* -algebra without any bounded linear projection from B onto
K. Proof. Indeed, for any A, we can find Xx, and UA: X A----> f3IK with lIuAl1 ::; 1 such that any lifting UA must have lIuAIl > A. Let BA C f3 be a separable C*-algebra such that K C BA and UA(XA) C BAlK. If PA: BA ----> K is any projection, we must have IIPAII ~ A-I (otherwise we would get a lifting UA associated to 1- PA with norm::; A). We will take A = 1,2, ... , and consider I = Co (K) c Co ( {Bn} ). Clearly, I is an ideal in a separable C* -algebra, and any projection P: co({Bn}) ----> I must induce (after a well known averaging) a projection Pn from Bn to K with IIPII ~ IlPn II ~ n - 1; hence P cannot be bounded. Finally, if we let B C K @min f3 be the C* -algebra generated by K @min I and by Co ( {Bn} ), then K @min K is an ideal (isomorphic to K) in B, but there is no bounded projection P: B ----> K @min K (because otherwise, composing P with the natural contractive projection Q: K @min K ----> Co(K), we would obtain a bounded projection from B onto I and a fortiori from Co ( {Bn} ) onto i). • Properties C, C f , and C". Exactness versus local reflexivity. In the C* -algebra case, the ideas developed in this chapter go back to Archbold and Batty [AB] , who introduced the properties C and C' defined below. Their work was continued and extended to the operator space setting by Effros and Haagerup [EH], who also added C" to the list. The reader should beware: This subject is full of traps; see Exercise 18.9 for an illustration!
Introduction to Operator Space Tlleory
310
Before we give the formal definitions, we need to introduce, given a pair of C*-algebras A, B, the natural inclusion of A** ® B** into (A ®min B)**, as follows. (See also Exercise 11.6(iii).) Assuming A,B unital for simplicity, we obviously have embeddings A
-4
(A ®min B)**
(defined by a -4 a ® 1 and b representations A*OO
-4
-4
(A ®min B)**
and
B
-4
(A ®min B)**
1 ® b) that extend to a pair of normal and
B**
-4
(A ®min B)**
with commuting ranges. The "product" of these representations gives us an *-homomorphism J: A** ® B** -4 (A ®min B)OO* that is clearly injective (since AOO ® Boo c (A ®min B)OO). Perhaps a slightly more concrete description of J emerges from the following obs@rvation: For any . (a,b) in A** x B** whenever (a o ) and (b{3) are nets in A,B with lIa o ll::; lIall, II bo II ::; II bll, tending respectively to a and b with respect to the Mackey topologies r(A**, A*) and r(B**, B*), then J(a ® b) is the weak-* limit of J(a o ® b{3) in (A ®min B)**. We claim that for any tin A** ® B** we have (18.2) Indeed, if we assume IIJ(t)II(A®minB)** ::; 1, there is a net to in the unit ball of A ®min B tending weak-* to J(t). Let U o : A* -4 Band u: A* - 4 B** be the linear maps associated, respectively, to to and t. Then, for any in A*, uo(e) -4 u(e) in the a(B**, B*) sense. But, by (2.3.2), we have, lIuo ll c b = IIto ll min and lIull cb = IItll min; hence, we obtain
e
IItll min
= lIull cb
::;
sup lIuo ll cb ::; sup IIto llrnin ::; 1, o
0
which establishes our claim (18.2). Let A be a COO -algebra. We will say that (1) A has property C if, for any C*-algebra B, the norm induced on the algebraic tensor product A** ® B** by (A ®min B)** coincides with the minimal norm. In other words, we have an isometric embedding A ** ®min B**
c (A ®min B)**.
(2) A has property C f if, for any C"-algebra B, the norm induced by (A ®min B)OOOO on A ® BOO" coincides with the minimal norm. In other words, we have an isometric embedding
A ®min B**
c (A ®min B) .....
18. Local Reflexivity
311
(3) A has property G" if, for any G* -algebra B, the norm induced by (A Q9min B)** on A** Q9 B coincides with the minimal norm. In other words, we have an isometric embedding A**
Q9min
Be
(AQ9 m in
B)**.
The implications G => G' and G => G" are obvious by restriction. Conversely, it is not too difficult to verify (see Exercise 18.6) that A has property G iff it has G' and G". Clearly these not.ions (and the preceding observat.ions) remain valid if A is merely an operator space, but the embeddings are now completely isomorphic embeddings and we must introduce the relevant constants. Let A ~ 1 be a constant. Then we will say that an operator space X has property G (resp. G', resp. G") with constant A if, for any G*-algebra B, we have a completely isomorphic embedding X**
Q9min
B**
~ (X Q9
B)**
(resp. X Q9min B** ~ (X Q9min B)**, resp. X** Q9min B ~ (X Q9min B)**) with c.b. norm majorized by A. Note that it suffices to majorize the norms by A, since by changing B to lI/n (B) (n = 1,2, ... ) we recover the c.b. norm. Moreover, by Gelfand's embedding theorem for C* -algebras, it clearly suffices in all these properties to consider the case B = B(H) with H Hilbert. We will denote by G(X) (resp. G'(X), rresp. G"(X)) the smallest number A for which X has property G (resp. G', rresp G") with constant A. We have clearly G'(X) ::; G(X) and G"(X)::; G(X). It is useful to observe that we have trivially G'(X)
= sup{G'(E)
lEe X
dimE < oo}.
(18.3)
The most interesting constant seems to be the constant G"(X), which is just the local reflexivity constant. Proposition 18.15. Let A ~ 1 be a constant and let X be an operator space. The following are equivalent.
(i) X has property G" with G"(X) ::; A. (ii) X is A-locally reflexive. Proof. To show (i) => (ii), we may assume E c B for some G*-algebra B. Note that X** ®min E C X** ®min B and (X ®min E)** C (X ®min B)** are isometric embeddings, so that, by restricting to X ®min E, (i) implies
312
Introduction to Operator Space Theory
IIX** ®min E ---+ (X ®min E)** II the converse is similarly easy.
:::;
A, which is the same as (ii). The proof of •
Theorem 18.16. Let A be a G*-algebra.
(i) (ii) (iii) (iv)
If A has property G, then A is exact. A is exact iff it has property G'.
If A is exact, then A is locally reflexive, that is, it has property G". G' {:} G {:} exactness.
Part (i) is due to Archbold and Batty [AB], and (ii) and (iii) are due to Kirchberg. Part (iv) is then immediate from (ii) and (iii): Indeed, by Exercise 18.6 we know G {:} (G'&G"), but (ii) and (iii) together show that G' => G"; hence G and G' must actually be equivalent and by (ii) G' is equivalent to exactness. Apparently, no simple direct proof of (iii) is known. It is derived from Theorem 17.18 (or from a previous result of Kirchberg representing separable exact G* -algebras as quotients of sub- G* -algebras of nuclear ones; cf. [Kil, Wa3]). For part (ii), however, a simple argument is available: Marius Junge observed an extension to operator spaces of this result as follows. Theorem 18.17. Let X be an operator space. Then X has property G' with constant A iff (18.4) sup{ dsx:{E) lEe X, dim E < oo} :::; A.
In particular, dsx:{E) = G'(E) for any finite-dimensional operator space. Proof. Assume (18.4). Note that the space MN and a fortiori any of its subspaces obviously satisfy G' with constant 1. Therefore, any finite-dimensional operator space E satisfies G' with constant dsx:{E). Thus, if (18.4) holds, X satisfies G' with constant A, because of (18.3). Conversely, assume that X satisfies G' with constant A. We will show that X is exact with constant A. By (17.6) and (18.3) we can assume that X is finite-dimensional. Then let I be an ideal in a G* -algebra B. Note that B** :::: I** EB (B /I)**, so that we have an isometric identification (B/I)**
®min
X :::: (B**
®min
X)/(I**
®min
X).
Now, if X satisfies G' with constant A, the inclusion B** ®min X ---+ (B X)** has norm:::; A; hence, after passing to the quotient, the inclusion (B **,o, . X)/(I** '6'mm
®
. mm
X)
t--+
(B
®min
X)**
(I ®min X)**
®min
18. Local Reflexivity
313
has norm::::: A. By the preceding identification, this means that the inclusion
(BII)**
®min
X
f----*
(B ®min X)** = (B ®min X)** (I ®min X)** I ®min X
has norm ::::: A. By restriction to B II
BII ®min X
®min
---+
X, the inclusion
(B
®min
X)
I®lIIinX
also has norm < A, but this says that ex(X) ::::: A. conclusion.
Whence the desired •
The following obvious reformulation of C' is useful. Proposition 18.18. Let E be a finite-dimensional operator space. Let A = C'(E). Then, for any CO-algebra B and any u: E* ---> B**, there is a net Ui: E* ---> B with SUPiEI Iluilicb : : : Allullcb that tends to u pointwise with B** equipped with the a(B**, B*)-topology. Indeed, CB(E*, B**) = E ®min B** isometrically; hence, if A = an element of norm::::: Allullcb in CB(E*,B)**. Therefore there is a net (Ui) as above tending to u in the topology a(CB(E*,B)**, CB(E*, B)*). But it is easy to check that this equivalently means that Ui ---> U in the point-a( B** , B*) topology. • Proof.
C'(E),
u defines
Proposition 18.19. (fAB}) Any nuclear C*-algebra A is locally reflexive. More generally, any quotient C* -algebra of a C* -subalgebra of A is exact (and locally reflexive). Proof. By Exercise 18.7 (since A nuclear implies A** injective) A must be locally reflexive (equivalently, it must have (C")). Moreover, it is obvious that nuclear =} exact. Hence (by Theorem 18.17) nuclear implies (C" and C'). By Exercise 18.6, this means that nuclearity implies (C). But now (C) passes to subalgebras (obviously) and to quotient C*-algebras (by Exercise 18.8); hence any quotient of a subalgebra of A has (C) and therefore (Theorems 18.16 and 18.17) is exact. • Remark 18.20. It seems to be a delicate open problem to decide whet.her the converse of Theorem 18.16(iii) holds. In other words: Is it true that local reflexivity implies exactness for a C* -algebra? Note that for operator spaces this is clearly false, since any reflexive operator space is obviously locally reflexive, and of course it is not exact (take, e.g., OH; see Theorem 21.5). Nevertheless, we have the following o.s. analog of Theorem 18.16(iii):
314
Introduction to Operator Space Theory
Theorem 18.21. ({EOR)) Any 1-exact operator space is 1-locally reflexive. Proof. The proof relies on a deep result of Kirchberg [Kill that is beyond the scope of this book. The latter result says that any separable 1-nuclear operator space X can be realized (completely isometrically) as a quotient B/(L + R), where B is the CAR C* -algebra (infinite tensor product of 1'If2 in the C*-sense) and L, R are respectively a left and right closed ideal in B. By classical results on one-sided ideals in C*-algebras, the bidual of B /(L+R) embeds canonically completely isometrically into B**. Using this, it is not hard to show (see Exercise 18.4) that the I-local reflexivity of B (note that B is a nuclear C*algebra) passes to B/(L+R). Thus any I-nuclear operator space X is I-locally reflexive. Since we have seen (Theorem 17.23) that any I-exact separable o.s. embeds in a I-nuclear one, we conclude (invoking Proposition 18.5) that 1exact implies I-locally reflexive. • Note that, as a consequence, X is I-exact iff it satisfies either (C) or (C') with constant 1.
Remark. Apparently, aU of the known proofs that a nuclear C* -algebra A or an exact one is 10caUy reflexive (see Exercise 18.7) use rather delicate (and somewhat indirect) arguments. If one could find a direct reasonably simple proof for the local reflexivity of A, one would have a much simpler demonstration that nuclearity and exactness. pass to quotients. There is very recent work by S. Wassermann along this line. Exercises Exercise 18.1. Let A be a C* -algebra. Show that A ** is injective if A is both WEP and locally reflexive. Exercise 18.2. Let X be a Banach space. Let E c X** be a finitedimensional subspace, and let U",: E - t X be a net of mappings with lIu",1I :::; A such that, for any e in E, u",(e) - t e with respect to a(X**,X*). Show that u'" is injective when a is larger enough and
Exercise 18.3. Fix n ~ 1. Let E, X be operator spaces. Let V: 1'Ifn (E) - t Mn(X) be a bounded map. Let G be the (finite) group of all unitary matrices such that each row and each column has exactly one nonzero entry equal to ±1. We define V(x)
= 1~12
L g,hEG
g-l.
V(gxh) . h- 1 .
18. Local Reflexivity
315
Show that we can write for some u: E
--+
X with
lIull n
:::::
IIVII·
(Note: Actually the same result holds if G is the whole unitary group.) Exercise 18.4. Let X be a A-locally reflexive operator space with a closed subspace Y c X. Let q: X --+ X/Y be the quotient map. Assume that there is a completely contractive map r: (X/Y)**
--+
X**
such that q**r is the identity on (X/Y)**. Show that X/Y is A-locally reflexive. In particular, any quotient of a C*-algebra by a (two-sided, closed) ideal is locally reflexive if it is the case for the C* -algebra ([AB]). Exercise 18.5. ([AB]) Show that C*(1F2) and C*(lFoo) are not locally reflexive. Exercise 18.6. ([EH]) Show that for a C* -algebra (C)
{=}
(C' & C").
More generally, if an operator space X satisfies C' with constant A' and C" with constant A", then it has property C with constant A' A". Exercise 18.7. Let A be a C* -algebra. Show that if A ** is injective, then A is locally reflexive (hence nuclear implies locally reflexive).
Hint: Use the fact that A ** is semi-discrete, that is, there are nets of complete contractions Va: A** --+ Aln(a) and Wa: Afn(a) --+ A** such that WaVa --+ IA-' pointwise in the a(A**, A*)-sense. Exercise 18.8. Show that if a C*-algebra A has property (C), then any quotient C*-algebra A/I also has (C). Exercise 18.9. Find what is wrong in the following FALSE ARGUMENT: Let IC = K(.e 2 ) and B = B(.e 2 ), and let A be any C*-algebra. Since IC c B, we have (IC Q9min A)** c (B Q9min A)** (isometrically), but, on the other hand (see Exercise 5.6), (IC Q9 m in A)** ~ BQ9A**j hence we have IIBQ9A** --+ (B Q9min A)** II : : : 1 and restricting to B Q9 A ** we obtain liB Q9rnin A ** --+ (B Q9min A)**II ::::: 1, which "shows" that A is locally reflexive.
Chapter 19. Grothendieck's Theorem for Operator Spaces We first recall the noncommutative version of Grothendieck's theorem (in short, GT) due to U. Haagerup and the author (see [P4] for details). Let A, B be C* -algebras. Then any bounded linear map u: A --+ B* satisfies the following: For any finite sequences in A and in B
(bi)
(ai)
1~)u(ai),bi)1 Kllull max
: :;
{IILa;aiII1/2 , IILaia;111/2} max {IlL b;bi111/2 , IlL: bib;11 1/2},
where K is a numerical constant independent of u. When A, B are commutative C*-algebras, this is a classical result due to Grothendieck. The best constant K in this case is called the (complex) Grothendieck constant and is denoted KG. It is known that 1.338 :::; KG :::; 1.405. See [Ko] for the latest information on this. The operator space version of GT, which we prove below, applies to completely bounded maps; hence we assume more than in the classical version of GT, but on the other hand it applies to all mappings u: E --+ F* where E and F are exact operator spaces. Thus we seem to be requiring much less structure on the domain and range ofu. For convenience, we will use the following. Notation. Let E be an operator space. Then, for any finite sequence (X;)i~n in E, we set "(Xi)"Rc=max{IILe li ® Xi Equivalently (see Remark 1.13), if E
l
c
M,,(E)
,IILeil®xill. }. M .. (E)
B(H), then
II(Xi)IIRC = max {IILX:XiII1/2, IILXix:111/2}. Theorem 19.1. Let E, F be exact operator spaces. Let C = dsdE)dsdF). Then any c.b. map u: E --+ F* satisfies the following inequality. For any finite sequences in E and (b i ) in F we have
(ai)
IL(u(a
i ),
bi)1 :::; 4C1lull cbll(ai)IIRclI(bi )IIRC'
More precisely, if we denote for simplicity X
= C! (JF00), we have
\ (19.1)
/
I ~)u(ai)' bi)1 :::; Cllullcb IlL ai ® -\(9i)II E®mlnX IlL bi ® -\(9i)11 F®mln.'l.' (19.1)' Recall that, given a map v: Y --+ Z between Banach spaces, we denote by 'Y2(V) its norm of factorization through a Hilbert space, that is, 'Y2(V) = inf{lI vlllllv211}, where the infimum runs over all possible Hilbert spaces H and all factorizations of v of the form Y ~H ~Z.
19. Grotllendieck's Theorem for Operator Spaces
317
Corollary 19.2. In tIle situation of the preceding theorem, let A, B be C*algebras with completely isometric embeddings E C A and FeB. Then there are states It, gl on A, 12, g2 on B, and 0 ::; (h'(h ::; 1 such that for a.ny (a,b) in E x F
l(u(a),b)l::; 4C1lull cb[Bdl(a*a)
+ (1- Bdgl(aa*)]1/2 (19.2)
u:
Consequently, tIl ere is a bounded linear map A -+ B* l\'ithllull ::; 1'2(it) ::; 4Cllull c b that extends u in tIle sense that, ifn'e \Tiew u and as bilinear forms on E x F and A x B, respectively, tllen it extends u. A fortiori, we bave 1'2(U) ::; 4Cllull c b.
u
Proof. The proof of the first assertion is entirely analogous to the solution to Exercise 2.2.2, to which we refer the reader. One should simply note that
where the supremum runs over all states j, 9 on A and all 0 ::; B ::; 1. The second assertion is proved as follows. Equation (19.2) allows write
liS
to
(19.3) where ( , hand ( , h are scalar products on A and B respectively sllch that V (a,b) E A x B If we denote by HI and H2 the Hilbert spaces obtained after passing to the quotient and completing, we have contractive inclusions
so that we deduce from (19.3) a factorization of the form
(u(a), b)
V(a, b) E Ex F
where T: HI mapping
-+
= (JiT J 1 (a), b),
H2 is an operator such that
it
=
IITII ::;
4Cllull c b. Thus, the
JiT J 1
is an "extension" of u (in the sense of Corollary 19.2) satisfying
• The preceding statement explains why GT is often described as a factorization theorem. Actually, we will prove the following slightly more "abstract" result:
Introduction to Operator Space Theory
318
Generalized Theorem 19.1. Let E,F be exact operator spaces and let C = dsdE)dsdF). Let At, A2 be C*-algebras. Assume that either Al or A2 is QWEP (i.e., is a quotient of a C*-algebra with the WEP). Then any c.b. map u: E -+ F* satisfies, for any finite sequences (ai) in E, (bj ) in F, (Xi) in AI, and (Yj) in A 2, tlle following inequality:
To prove Theorem 19.1, the key ingredient will be the embedding of the von Neumann algebra of the free group into an ultraproduct described in §9.1O. In addition, we will use the following fact. Lemma 19.3. Let C = dsdE)dsdF) as before. Let At, A2 be C*-algebras and assume Ai = Bd h where Bi are C* -algebras and Ii C Bi closed twosided ideals. Let qi: Bi -+ Ai be the quotient map. Let cp E (Ai ® A2)* be a
linear form such that
Then, for any finite sequences (ai), (b i ), (Xi), (Yj) with (ai,bj ) (Xi, Yj) E Al x A2 and any linear map u: E -+ F*, we have IL(u(ai ), bj)CP(Xi ® Yj)1
~ Cllullcb
ilL ai
® xitin
ilL bj
E
Ex F and
® Yjtin .
(19.4)
Remark. This result will be applied to a form cP that is unbounded on Al ®min A 2, so we really need to consider CP(ql ® q2) instead of cpo Remark 19.4. If ql ® q2 maps (contractively) BI ®min B2 into Al ®max A 2, then (19.4) holds for all cp such that IIcpll(Ali81max A2)* ~ 1, since this implies a fortiori Ilcp(ql ® q2)II(Bli81mln B2)* ~ 1.
We will use the following simple fact (for a proof, see the solution to Exercise 19.1). / Lemma 19.5. Let E, F, G be operator spaces. Consider a linear map u: E CB(F,G). Then, for any C*-algebras Bt, B 2,U defines a bilinear form
u:
E ®min BI
X
F ®min B2
-+
G ®min BI ®min B2
satisfying Ilull ~ Ilulicb and u(a ® X, b® y) = u(a)(b) ® x ® y.
--+
19. Grotllendieck's Theorem for Operator Spaces
319
Proof of Lemma 19.3. We may clearly assume, without loss of generality, that E and F are finite-dimensional. We first assume E C 1Iln and F C 111m. Let x = E ai ® Xi and Y = E bj ®Yj. Assume Ilxllmin < 1 and Ilyilmin < 1. Then, by Lemma 17.2, there are x in E ®min Bl and fi in F ®min B2 such that (I®qJ)(x) = x and (I®q2)(fi) = y, with 1IXIlmin < 1, lifilimin < 1. Indeed, this is clear when E = Aln and F = 111m , but Lemma 17.2 ensures that it remains automatically true for subspaces. We may clearly assume = E ai 0 Xi, fi= Ebj ®fij with ql(Xi) = Xi and q2(fij) = Yj. Then, with the notation of Lemma 19.5 (taking G = q, we have
x
u(x, fi) = L(u(ai), bj)x; ® fij. i.j
Hence, applying c.p
0
(ql ® q2)
to this, we obtain
L (u(ai), bj )c.p(Xi ®Yj)
=
Ic.p 0 (ql ®q2)(U(X, fi))1
;,j
::; 11c.p 0 (ql ®q2)II(B1®"'inB2)*lluli Ilxll II fill ::; Ilulicb. Thus we obtain the desired inequality when E, Fare matricial spaces. The general case is then easy to prove using isomorphisms v: E -+ E C 1I1n and w: F -+ F c 111m. Replacing E by E and F by F produces an extra factor equal to Ilvllcbllv-lllcbllwllcbllw-lllcb, whence the presence ofthe constant C = dsdE)dsdF) in (19.4). • Proof of Theorem 19.1. We use the construction described in §9.1O. vVe can take for (Xi) either (A(gi» (with (g;) the generators of Foe) or a free semicircular (or circular) sequence and we let Yj = Xj' In both cases, we have (see §9.7 and §9.9)
IlL
ai
®Xiii::; 211(a;)IIRC
and
IlL b ®Yjl : ; 211(b )IIRC j
j
(19.5)
when ai, bj are elements of an arbitrary operator space. Let A! be the von Neumann algebra generated by (x;). By §9.10, there is a family of (finite-dimensional) matrix algebras M(n), a free ultrafilter U, and an ideal Iu such that M embeds into Bj1u where B = foe({JI!(n) I n ~ I}); and moreover, if q: B -+ B j Iu denotes the quotient map and T ( resp. Tn) the normalized trace on M (resp. M(n», we have for any t = (t n ) in B (19.6)
320
Introduction to Operator Space Theory
We identify M with a subalgebra of B /Iu as described in §9.10. We will apply Lemma 19.3 with Yi = Xi, (Xi) being as above with BI = B, B2 = B, ql = q,q2 = q, and with cP: B/Iu ® B/Iu - t C given by cp(s ® I) = limu CPn(sn, tn),
where CPn(Sn, tn) = rn(snt~).
Note that, since the M(n) are finite-dimensional matrix algebras, it is clear (see, e.g., Proposition 2.9.1) that IICPnll(M(n)®lIIin 1lf (n))* ::; 1 for any n, and hence Ilcpll(B®min B )* ::; 1. Then, recalling (19.5) and noting that (19.6) implies cp(Xi®Xj) = r(xixj) = 8ij , we can finally deduce (19.1) and (19.1)' from (19.4) and (19.5). • Proof of Generalized Theorem 19.1. We will use Lemma 19.3. Assume, say, that Al is QWEP. Let BI be WEP such that BdII ~ AI, and let B2 = C*(G) with G a suitable free group so that A2 ~ B2/I2 (see Exercise 8.1). By Kirchberg's Theorem (see Theorem 15.5 and the remark after it), we know that BI ®min B2 = BI ®max B 2, and hence a fortiori
Ilql ® q2:
BI ®min B2
-t
Al ®max
A211 ::; 1.
Thus Remark 19.4 applies in this case and gives us (19.1)".
•
Lemma 19.6. In the situation of Lemma 19.3, assume that Al = A2 = A, where A is a quotient of a WEP (in short QWEP) C* -algebra. Then, for allY tracial state '1/1: A - t C, V ai E E, V bj E F, V Xi E A, V Yj E A, we have
I~)u(ai)' bj )'1/1 (Xi Yj ) I ::; dsdE)dsdF)llullcb IlL ai ® xillmin ilL bj ® Yj Lin'
Proof. We may apply (19.1)" with Al = A and A2 = AOP. Now, since '1/1 is a tmcial state, we can write 'I/1(xy) = (7I"1(X)7I"2(Y)';,';), where 71"1. 71"2 are commuting representations of A and AOP associated to the GNS construction relative to '1/1 and 11';11 = 1. Thus we find that x®y - t 'I/1(xy) defines an element in the unit ball of (A ®max AOP)*, so that (19.1)" gives us \ the conclusion. / • Remark. Note that the von Neumann algebra M appearing in the proof of Theorem 19.1 is QWEP by Exercise 15.4. This explains our terminology "Generalized Theorem 19.1." As we observed earlier, the spaces R, C and their direct sum R EEl Care the only known examples of infinite-dimensional separable operator spaces E that are exact as well as their duals. It is a natural question to ask whether they are indeed the only ones. An affirmative answer was very recently given in [PiS]. The next statement is a first step in this direction.
19. Grotllendieck's Theorem for Operator Spaces
321
Corollary 19.7. If an operator space E is exact as well as its dual E*, then E must be isomorphic to a Hilbert space. More precisely, we have
Proof. This is a immediate consequence of Corollary 19.2.
•
Remark 19.8. In Lemma 19.3, the exactness of E (or F) is used only to lift elements of E 0 m in (BdIt) up into E 0 m in B I . Therefore, if II = {O} (resp. if both II = 12 = {O}), then Lemma 19.3 remains valid when E (resp. when each space E or F) is an arbitrary operator space, and (19.4) is valid with C = dsdF) (resp. with C = 1). In particular, we find: Corollary 19.9. Let E,F be arbitrary operator spaces. Let {cn I n 2: 1} be a system satisfying the CAR (see §9.3). Tllen any c.b. map u: E -+ F* satisfies for any finite sequences (ai) in E and (b i ) in F
Proof. Recall that {CI,"" cn } can be realized in a finite-dimensional C*algebra J\/(n), with normalized trace Tn. Since CiC; + C;Ci = oijI, the system {c;j21/2 Ii::; n} is orthonormal with respect to 1/Jn. Hence we may apply the preceding Remark 19.8 with Al = BI = lIl(n), A2 = B2 = M(n),
Proof. This is an immediate consequence of the preceding corollary and the fact that if E = min(E), we have by (9.3.1)
I Lai0Cili.
=
mm
and similarly for
L
bi 0
sup
~EBB'
Ci
if F
IIL~(ai)Cill::; = min(F).
sup
~EB~:.
(LI~(aiW)I/2,
•
Remark. The preceding proof works equally well if we use a spin system instead of the system (cn ).
322
Introduction to Operator Space Theory
Let E, F be Banach spaces. Consider a linear map u: E ---+ F*. \Ve denote by ')'2(u) the smallest constant C such that, for any finite sequences (ai) in E and (bd in F, we have
As the notation indicates, this norm is dual to the ')'2-norm (introduced before Corollary 19.2) in the following sense. For any v E E®F, let us denote E* ---+ F determined by v. It is by ')'2(V) the ')'2-norm of the linear map easy to see that ')'2 (v) < 1 iff v can be written as v = L ai ® bi with
v:
Therefore we have a duality formula:
')'2(u)
= sup{l(u,v)11 vEE ® F, ')'2(V) ::; I}.
(19.8)
Note that, by Exercise 2.2.2, ')'2(u) ::; 1 iff there are probability measures >. and JL on (BEo,a(E*,E» and (BFo,a(F*,F» such that V (a,b) E Ex F
l(u(a),b)l::; ( /
1~(aWd>'(~») 1/2. ( / 117(bW dl l(17») 1/2
(19.9) The duality between ')'2 and ')'2 originally goes back to Grothendieck [Gr]. The norm ')'2 is now fairly well understood in Banach space theory, and (19.9) can be reinterpreted in terms of 2-absolutely summing operators (see [P4] for more on this). Thus the next result gives (at least in a special situation) a meaningful equivalent of the c.b. norm (due to V. Paulsen and the author on one hand, see [Pa5], and also independently to M. Junge). Theorem 19.11. Let E, F be minimal operator spaces. Then, for any c.b. map u: E ---+ F*, we have \
./
(19.10)
Proof. By (19.8) (and the remarks preceding it), Corollary 19.10 implies
whence the left side of (19.10». As for the right side, it simply follows from the observation that, for any v in E ® F (here E, F can be arbitrary operator spaces), we have (19.11)
19. Grothendieck's Theorem for Operator Spaces Indeed, assume IIvIIE01\F < 1. Then the map admits a factorization of the form:
v:
E*
---+
323
F associated to v
'" 111 ~ ,I"n -----+ L" sn RJ> sn (3 F E * -----+ 2 -----+ 1 -----+ ,
Ilnllcb < 1, 1I.6Ilcb < 1 and where La(x) = ax, Rb(X) = xb with IIal12 . IIbll 2 < 1. Then, if we let V2 = Lan and Vl = /3Rb, we obtain v'= VIV2 and IlvI11 . IIv211 < 1; hence 1'2(V) < 1, which proves (19.11). By duality, lIsing • (19.8) and Chapter 4, (19.11) implies Ilulicb ::; 1'2(U).
where
Remark. Let E be again minimal, let G be a maximal operator space, let u: E ---+ G be a linear map, and let ic: G ---+ G** be the canonical inclusion. We then have Tll'~(icu) ::; Ilulicb ::; I'~(icu). Indeed, by Exercise 3.2, G* is a minimal operator space, so this follows from the preceding result. Remark. Note that the preceding statement gives a satisfactory description of CB(E, F*) as a Banach space, but its operator space structure remains unclear, in particular, the following is open.
Problem. Let E, F be minimal operator spaces (take for instance E = F = co). Is E*0 m inF* completely isomorphic to tile symmetrized Haagemp tensor product E* 0", F*? Actually, tllis might even be tme whenever E, Fare exact. The same question arises with two maximal operator spaces instead of
(E*, F*). In addition, there is evidence that the preceding question may have a positive answer when E, F is any pair of C* -algebras. A quite similar question is already raised in [B3] and at the end of [ER3]. Some very recent progress was made in [PiS]: It is proved there that if E, F are exact operator spaces, or if E, F are both C* -algebras, it is indeed true that E* 0 m in F* ~ E* 0", F* with equivalent norms. In particular, in the case E = F = Mn, it is proved in [PiS] that the norms of the identity maps M~ 0 m in M~ ---+ M~ 0", M~
are bounded uniformly over n. Unfortunately it remains unclear at the time of. this writing whether their c.b. norms are also uniformly bounded.
Exercises Exercise 19.1. Prove Lemma 19.5. Exercise 19.2. Show that dsdmax(f~)) ~
vn/4.
(Actually, this holds with any n-dimensional normed space in place of f~; see [JP].)
Chapter 20. Estimating the Norms of Sums of Unitaries: Ramanujan Graphs, Property T, Random Matrices In this chapter, we estimate the growth of a specific sequence of numbers C(n) that are closely related to the analysis of "asymptotic freeness" for sequences of n-tuples of (N x N) matrices with a common size N tending to infinity. More precisely, for each n 2': 1, we define C(n) as the infimum of the numbers C for which there exist integers {Nm I mEN} and an infinite sequence of n-tuples of N m x N m unitary matrices {(ui(m)h
{lit
ui(m) ® Ui(ml)11
}:::; C. min
t=l
The last norm is meant in AIN m ®min AIN m' , or, equivalently, rn the (operator) . norm of the space MNmxNm , of all matrices of size NmNm, x NmNm,. By the triangle inequality, we have IIL~ Ui ® Vi II min :::; n whenever Ui, Vi are all unitary; hence we have the trivial bound: C(n) :::; n,
which, as we will see, is far from the true value of C(n) when n is large. However, we have C(2) = 2. This follows from Corollary 20.2 (the reader is invited to find a direct proof, as an exercise). We will first estimate C(n) from below. We start with a result from [P14]. The alternate proof that we give here is due to Szarek. Theorem 20.1. Let Ul,' .. ,Un be arbitrary unitary operators in B(H) (H any Hilbert space). Then
2v'n=l:::;
IltUi ®Uil . t-l
mm
Remark. If dim H < 00, the triangle inequality II L~ Ui be improved, and we have
II
® Ui min :::;
n cannot
(20.1) Indeed, t = IH is an eigenvector for t -+ L Uitut associated to the eigenvalue n. More generally, it is easy to see that (20.1) still holds when dim H = 00 if Ub"" Un all belong to a finite injective von Neumann subalgebra Me B(H). However, (20.1) is not true if we drop the injectivity assumption, as shown when M is the von Neumann algebra (factor actually) associated to the free group Fn on n generators. We first recall that A: G -+ B(l2(G)) denotes
20. Estimating the Norms of Sums of Unitaries
325
the left regular representation that takes an element 9 in G to the unitary operator of left translation by g. Now, in t.he particular case G = F n , let gl, ... ,gn be the generators of Fn. Then it is known that (20.2)
Indeed, by Fell's absorption principle (see Proposition 8.1), the left-hand side is the same as IIE~ .\(g;)II, and the latter norm was computed in [AO] and found to be equal to the middle term of (20.2) (see the following remark). The results of [AO] were partly motivated by Kesten's thesis [K], where it is proved that (20.3) and also that (20.3) realizes the minimum of all norms IIEIEs .\(t)11 when 8 runs over all possible symmetric subsets of cardinality 2n of any discrete group G. Theorem 20.1 can be viewed as an abstract version of Kesten's lower bound.
Remark. We will not include the proof that IIE~ .\(g;)11 = 2yn=1. Note, however, that IIL~ .\(g;)11 2': Vn is obvious and IIE7.\(g;)11 :::: 2Vn follows from (9.7.1). Thus, we have at least proved that
Proof of Theorem 20.1. Let C = {t E 8 2 I t 2': 0 that, for any Ui in B(H),
IItll2
= 1}. We claim
IIEui@uill=sup{tr(Euitu;s) It,SEC}.
(20.4)
To prove this, first note that, for any t, S in C, (20.5) By definition, if T
= L:~ Ui @ Ui,we have
Moreover, every z in 82 can be written as Zl - Z2 +i(Z3 - Z4) with Zl, ..• ,Z4 all 2': 0 such that L:: IIzjll~ = IIzll~. From this fact and the preceding observation it is easy to check our claim (20.4).
Introduction to Operator 8pace Theory
326
Let T = L~ Ui ® Ui and let 8 = L~=l )..(gi). The idea of the proof is to show that, for any integer m ~ 1 and any t in C, we have (20.6) where 8e denotes the basis vector in t'2(lFn ) indexed by the unit element of IFn • Note that the normalized trace 7 in VN(lF n ) is given by the formula
To verify (20.6), note that we can expand (T*T)m as a sum of the form LOEI UO ® u o , where the UO are unitaries of the form UtI ujt Ut2 uh .... Now, for certain a, we have Uo = I by formal cancellation (no matter what the Ui are). Let us denote by I' c I the set of all such a. Then by (20.5) we have for all t in C
((T*T)mt, t)
=L
tr(uOtuo*t) ~
L
1 = card(l'),
oEI'
oEI
but by an elementary counting argument we have
Hence we obtain (20.6). Therefore IIT*TII ~ lim ((T*T)mt, t)l/m ~ lim (7((8* 8)m))l/m m~oo
= 118* 811,
m~oo
so that we obtain IITII ~ 11811, whence Theorem 20.1 by (20.2). See Exer• cise 20.1 for another proof. Corollary 20.2.
2Vn=l ~ C( n)
for all n ~ 1.
Proof. Let (ur )i$n be a sequence of n-tuples with (ur, . .. ,u~) unitary in the space AfNm of all N m x N m complex matrices. Let A be the space formed of all families x = (Xm)mEN with Xm E MN", and SUPm IlxmllMNm < 00. Equipped with the norm Ilxll = sup IlxmlIMN", , A becomes a C*-algebra. Let U be a nontrivial ultrafilter and let Iu c A be the (closed two-sided selfadjoint) ideal formed of all sequences x = (Xm)mEN such that limu Ilxmll = o. Then the quotient space AlIu is a C* -algebra called theultraproduct of {MNm I mEN} with respect to U. See also §2.8. By Gelfand theory we . can view AlIu as embedded into B('H) for some Hilbert space 'H. Let us denote by Ul, ... ,Un the unitary elements in AlIu associated to the families (Ur)mEN,"" (U~)mEN' We claim that, for any al, ... , an in B(H) (with H arbitrary), we have (20.7)
20. Estimating tile Norms of Smns of Unitaries
327
Indeed, the quotient mapping q: A ---+ Al1u being a C* -representation is completely contractive, and hence (recalling (2.6.2» IILu; 0 a;11 :S SUPm IlL ujn 0 a; II, but since the left side of (20.7) depends only on the equivalence class modulo U, this last sup can be replaced by the limit along U, and (20.7) follows. Now, if we apply (20.7) with a; = Ui E B(Jt), we obtain by Theorem 20.1
2vn=I :S
IlL Ui 0Uil :S ~?J IlL ui 0it; I =
=
~~D
IlL ujn 0u;11
~~JIILUi0Uillj
hence by (20.7) again
:S m,Um,U lim liJIl and the last term is of course
:S sup, m#m
IlL ui' 0uill '
lit ui'l · i=1
ui 0
•
Thus we conclude that 2Jn=l :S C(n).
We now turn to the much more delicate task of majorizing C(n). Until very recently, the best known estimates (see [Va]) used some deep number theoretic results ("Ramanujan graphs") due to Lubotzky, Phillips, and Sarnak ([LPS]), themselves based on Andre Weil's celebrated proof of "the Riemann hypothesis for curves over a finite field." Actually, it is somewhat easier for our exposition to derive our estimates from an application of these results (by the same authors [LPS]) to a "packing problem" on the sphere of 1R 3 , as follows. We will denote by 8 the Euclidean unit sphere in 1R 3 , equipped with its normalized surface measure a. We let
We consider the representation p: SO(3) \Ix E 8
---+
B(Lg) defined by:
[p(w)f](x) = f(w- 1 (x».
This is usually called the "quasi-regular" representation of 80(3), except that we restrict it to the orthogonal of the constant functions, namely, Lg. Theorem 20.3. ([LPS]) For any n of the form n = p + 1 witi] p prime ~ 3, there are elements tl'· .. ' tn in 80(3) such that
Il tp(ti)11 ,=1
:S 2Jn=l. B(L~)
328
Introduction to Operator Space Theory
Corollary 20.4. For any n 2:: 4 of the form n C(n):::;
= p + 1 with p prime we have
2Jn=l.
Proof. Since 80(3) is a compact group, the unitary representation p: 80(3) -+ B(Lg) decomposes as a direct sum of a sequence (1I"m)m~1 of irreducible finite-dimensional unitary representations of 80(3), that is, we have p
~
E9
1I"m·
m~l
The representation 1I"m is just the restriction of p to the spherical harmonics of degree m (for details see, e.g., p. 161 in [Fol]). In particular, for any t l , ... , tnE 80(3) we have
These are distinct (Le., mutually inequivalent) representations. l\-{oreover, it is classical that the collection {11" m I m 2:: 1} exhausts all the nontrivial irreducible representations of 80(3). (Note t,hat this is special to the dimension 3.) Therefore, if m =I- m', then 11" m ® 11" m' decomposes again as a direct sum of a certain subset (depending of course on m and m') of {lI'm 1m 2:: 1}. Note that by Schur's lemma, 1I"m i= 1I"m' guarantees that the trivial representation is not contained in 1I"m ® 1I"m" Consequently, we have
Therefore, if we choose the points tt, ... , tn as in Theorem 20.3 and let ui(m) = 1I"m(td with N m = dim(1I"m), then we obtain C(n) :::; 2v'n"'=l. • For the applications in the next chapter, the crucial point is that C(n) < n for suitably large n. The ideas revolving around Kazhdan's property T (cf. [DRV, Vol]) provide us with a different route to this fact:
Theorem 20.6. C(n) < n for any n > 2. Definition. A finitely generated discrete group G, wit11 generators g}' 92, ... , 9n, is said to have property T if the trivial representation is isolated in the set of all unitary representations of G. More precisely, this means that there is a number c > 0 such that, for any unitary representation p, the
condition
20. Estimating the Norms of Sums of Unitaries
329
suffices to conclude that p admits a nonzero invariant vector, or, in other words, that p contains the trivial representation (as a subrepresentation). It is easy to see that this property actually does not depend on the choice of the set of generators.
Remar k. Let 7f, a be two unitary irreducible representations of a discrete group G. Then Schur's classical lemma implies that, if 7f0a admits a nonzero invariant vector (in H'/r 02 H (7), then 7f is unitarily equivalent to a and both are finite-dimensional representations. Indeed, such a vector can be identified with a nonzero Hilbert-Schmidt operator T: H(7 --+ H'/r such that 7f(t)Ta(t)* = T for all t in G. Equivalently, T intertwines 7f and a (Le. 7f(t)T = Ta(t) for all t in G). By Schur's lemma, T must be an isomorphism, but since it is HilbertSchmidt, both H'/r and H(7 must be finite-dimensional. Moreover, we have T*7f(t)* = O'(t)*T*. Hence, multiplying this on the right with 7f(t)T = TO'(t), we find T*T = a(t)*T*Ta(t), which shows that T*T commutes with a and therefore is a multiple of the identity. Thus we may assume that T is unitary, and we conclude that 7f and a are unitarily equivalent. • We will use the following well-known fact.
Lemma 20.7. Let G be a discrete group with property T generated by 91,92, ... ,9n with 91 equal to tIle unit. Tlwn (20.8)
where the supremum rllns over all pairs of distinct (i,e., not unitarilyequivalent) irreducible representations of G. Proof. Let c > 0 be as in the preceding definition of property T. Actually, we will show more generally that (20.8) holds with the sup running over all pairs of disjoint representations, that is, pairs (7f,7f'), such that 7f 0 7f' does not contain the trivial representation (equivalently has no invariant nontrivial vector). By the preceding remark, this happens whenever 7f, 7f' are two distinct irreducible representations. We now complete the argument: If (20.8) fails, then, for each m, we can find a disjoint pair (7fm, 7f:n) and ~m with lI~m II = 1 such that
By the uniform convexity of Hilbert space (see Exercise 20.2), this forces the n-tuples of points [7fm(9i)07f:n(9i)](~m) (i = 1, ... , n) to all collapse to a single point, (Le., the diameter ofthis n-tuple must tend to zero when m --+ 00), and
330
Introduction to Operator Space Theory
since 91 is the unit element, we have 11"m (9t) ® 1I"in (91)( ~m) must have for each i = 2, ... ,n
= ~m j therefore we
In particular, when m is large enough we find that Pm = 1I"m ® 1I"in satisfies sUPi
O~
0° ) 1
and 93
= (0~ ~1 0) ~ .
Let {11"m I m ~ 1} be a sequence of distinct finite-dimensional irreducible representations of SL3('Z). (Note that the existence of such a sequence is immediate by considering the morphisms SL 3('Z) ----> SL 3('Z/p'Z). Indeed, the irreducible representations of SL 3('Z/p'Z) are finite-dimensional and give rise to irreducible representations of SL 3 ('Z), with finite range, which separate points. Since SL 3 ('Z) is infinite, there has to be infinitely many inequivalent representations appearing in this process.) If we set Ui (m) = 11"m (gi) as before, then (20.8) implies that C(n) < n. • We now turn to yet another proof that C(n) is small, this time based on Gaussian random matrices. In a preliminary version of this book, we used the difficult estimates from [HT1] to show that
C(n) ~ (311"/8)vn/2.
(20.9)
But, just as our manuscript was about to go to the printer, we learned of the improvements of [HT2], which prompted us to replace the proof of (20.9) by a brief outline of the results of [HT2] showing that C(n) = 2~ for all n ~ 2, as follows. Theorem 20.S. ([HT2]) For any n ~ 2, CA(lFn ) embeds, as a unital C*subalgebra, into the C·-algebraic ultraproduct IIaM(o:)/U of a family ofma- . trix algebras {M(o:) I 0: ~ 1} (that is, there are integers N(o:) such that 1\1(0:) = MN(a) and U is an ultrafilter on N). Note that the difficulty lies in the fact that the ultraproduct IIaM(o:)/U is meant here in the norm sense as in §2.8 (and not in the von Neumann
20. Estimating tIle Norms of Sums of Unitaries
331
sense as in §9.1O). The proof of Haagerup and Thorbj0rnsen [HT2] rests on a delicate computation of the norms of Gaussian random matrices with matrix coefficients. Then, using t.his, t.hey can embed the C* -algebra C* (Cl, ... , c n ) generat.ed by an n-tuple of free circular elements (see §9.9) into an ultraproduct of the form ITa JI,f (a) /U. Since C,\ (11"n) embeds into C* (CI , ... , cn ), the result follows a fortiori for C,\ (IFn). As an immediate consequence, we have Corollary 20.9. ([HT2J) For any n ::::: 1 and any a ::::: 1, we can find an n-tuple of unitary a x a matrices (U?h::;i::;n such that, for any N and allY aI, ... , an in AIN, n-e IlaFe lim o,U
II2:
n
I
ai ®
urll . = II2: min
n
I
a; ®
-\(g;)11 ..
(20.10)
IlUIl
Proof. According to Theorem 20.8, we may view C,\(IFn) as embedded int.o IToM(a)/U. Then, for each 1 ::; i ::; n, let (UnO~I be a representative of -\(gi) in ffio M(a). Since )..(g;) is unitary, we must have limu 111 - (un*(unll = 0; but an elementary argument (based on the polar decomposition of shows that, actually, we may as well assume that u? is unit.ary for each n. vVe have then, for any fixed N, an isometric embedding
un
In particular, for any aI,"" an in AIN we have
II2:
n a; I·
®)..
i
(g)
I
MN(q(lF,,)l
=
lim a,U
II2:
n ai I·
®
UO 1
I
II/N(II/(a))
•
•
Remark. Actually, using Proposition 13.6, one can deduce conversely Theorem 20.8 from Corollary 20.9 (but, at the time of this writing, no direct proof of Corollary 20.9 is known). By Corollaries 20.2 and 20.4, we have C(p+l) = 2/p for all prime numbers p::::: 3. By [Mor], this remains t.rue if p is a power of a prime number, but. t.he general case was settled only very recently by Haagerup and Thorbj0rnsen. Theorem 20.10. ([HT2J) C(n) ::; 2Jn=1 (and hence C(n) = 2Jn=1) for any n::::: 2. Proof. Fix e > O. Obviously it suffices to construct a sequence of n-tuples {(u;(m)h::;i::;n 1m::::: I} of unitary matrices (note: u;(m»l::;i::;m is assumed to be, say, of size N m x N m ) such that, for any integer p 2: 1, we have (20.11)
332
Introduction to Operator Space Theory
We will construct this sequence by induct!on on p. Assume that we already know the result up to p. That is, we already know a family {(ui(m)h~i~m I 1 ~ m ~ p} formed of p n-tuples satisfying (20.11). We need to produce an additional n-tuple (Ui(P + l)h~i~n of unitary matrices (possibly of some larger size Np+1 x N p+1) such that (20.11) still holds for the enlarged family {(ui(m)h
On the other hand, by the absorption principle (Proposition 8.1) and by (20.2), we have
Hence, if 0: is chosen large enough, we can ensure that, for all 1 simultaneously, we have
~ m ~ p
But then, if we set Np+l = N(o:) and Ui(P + 1) = ui, the extended family {(ui(m)h~i~n 11 ~ m ~ P + 1} clearly still satisfies (20.11). • Remark. Let U(o:) denote the group of all 0: x 0: unitary matrices (0: 2: 1). Let uiO:), ... , UAO:) be a sequence of independent matrix valued random variables, each having as its distribution the normalized Haar measure on U(o:). It is very likely to be true that, for all N and for all all"" an in J..JN, we have, for almost all w, (20.12) and in particular, if aI, ... ,an are all unitary, for almost all w, lim sup 0-+00
liEn 1
ai
® U?(w)
II .
mlD
= 2vn-=-I. ,/""
If true, this would yield a more direct proof that C(n) ~ 2vn=-T. The paper
[HT2] contains an analog of (20.12) with Gaussian random matrices instead of {UiO:(w)}. Remark 20.11. It is sometimes useful to replace C(n) by another constant, fJ(n). We define fJ(n) as the supremum of all the numbers fJ 2: 0 for which there are two infinite sequences {(xi(m)h;5i;5n I mEN} and
20. Estimating the Norms of Slims of Unitaries
333
{(Yi(m)}ts;sn I mEN} of n-tuples of operators in B(H) (H arbitrary) satisfying sUPm.; Ilxi(m)11 + IIYi(m)11 < 00 together with:
~~~,
lit
xi(m) ® Y;(ml)11 . :::; 1 and (3 :::;
,.-1
nlln
lit 1-1
xi(m) ® Yi(1n)11 . \1m E N. 111111
It would be more natural to assume Xi(m), Yi(m.) E B(Hm) with Hm HilbE'rt, but the resulting (3(n) would be the same (since we can replace each Hm by the direct smn H = EfJmHm). This definition is closely related to the notion of "coding sequence" introduced by Voiculescu in [Vol]. Note that we have obviously
n/C(n) :::; (3(n). Exercises Exercise 20.1. (Alternate proof of Theorem 20.1.) Let a(n) = 112:7 A(gi)ll. Let Ul, ... , Un be unitary operators. Using (7.2) and Proposition 8.1 (Fell's absorption), show that
and conclude that
Exercise 20.2. Let x = (Xl, ... , xn) be an n-tuple in the unit ball of a Hilbert space H. Let 1Il(x) = n- l 2:7 Xk and ~(x) = max1sihsn Ilx; - xjll· Show that 111I1(x) 112 + n- l L~ Ilxk -lIJ(x)1I2 :::; 1. Deduce from this that, if a sequence of n-tuples {x(m) I In E N} (in the unit ball) is such that 11l\f(x(m»11 -> 1 when m -> oc, then ~(x(m» -> O.
Chapter 21. Local Theory of Operator Spaces. Nonseparability ofOSn One of the most interesting benefits of operator space theory is the possibility to employ finite-dimensional methods in C* -algebra theory, and some ideas from the "local theory" of Banach spaces have already proved very fruitful in the study of operator spaces. The "local theory" is the part of the theory that studies infinite-dimensional Banach spaces through the asymptotic properties of the collection of their finite-dimensional subspaces. If E and F are Banach spaces of the same dimension, they clearly are isomorphic, but one can measure their "degree of isomorphism" using their "Banach-Mazur distance," which is defined as follows: d(E, F) = inf{llullllu-111}, where the infimum runs over all the isomorphisms u bet.ween E and F. \Ve have clearly for any G of the same dimension d(E, F) :::; d(E, G)d( G, F). Moreover, d(E, F) = 1 iff E and F are isometric (hint: use the compactness of the unit ball of B(E, F) to check this). Thus, if we identify E and F whenever they are isometric, we can equip the set Bn of all n-dimensional Banach spaces with the metric 8(E, F) = Log d(E, F). By a classical theorem due to Fritz John (1948) we have d(E,£~) :::; yin for every n-dimensional normed space (see the remark after Corollary 7.7), and consequently d(E, F) :::; n for every pair E, Fin Bn. Moreover: Theorem 21.1. For each n 2: 1, (Bn' 8) is a compact metric space. Proof. Let {Em} be a sequence in Bn. Choosing an appropriate basis and using Fritz John's aforementioned theorem, we may assume that Em = (en, II 11m) and that
v x E en where we set
Ixl ~ (~lx;lf'
/
Let 0 = {x E en Ilxl :::; I}, and let 1m: 0 -+ R be defined by Im(x) = IIxll m. Note that 11m (x) - Im(y)1 :::; Im(x - y) :::; vnl x - yl and Im(O) = 0; hence, by Ascoli's Theorem, the sequence {1m} is relatively compact in C(O). Thus, after passing to a subsequence, we may assume that 1m converges uniformly
21. Local Theory of Operator Spaces. Nonseparability of OSn
335
on n to a limit; therefore, by homogeneity, for any x in en, IIxll m converges to a limit Ilxli oo such that Let Eoo = (en, II 1100)' We claim that 8(Em, Eoo) -+ O. Indeed, if we let c(m) = sup{lllxll oo - Ilxllmll x En}, we have for any x in en IlIxll oo
-
IIxllml :::; c(m)lxl :::; c(m)llxll m;
hence
(1 - c(m))llxll m :::; Ilxll oo
:::;
(1
+ c(m))lIxll m,
which implies d(Eoo, Em) :::; ~~:~:~. Since c(m) -+ 0, we obtain d(Eoo, Em) -+ 1 and 8(Eoo, Em) -+ O. Thus we have shown that any sequence {Em} in (Bn, 8) contains a con• vergent subsequence, which means that (B n ,8) is compact.
Remark 21.2. Fix an integer N. Assume that all the spaces Em are given equipped with an isometric embedding J m : Em -+ B(Hm). For al,'" ,an E /t;/N we set Fm(aI,"" an)
=
IlL ai ® Jm(ei)IIMN(B(Hm
))'
Then it is easy to show by the same reasoning as above (see also Exercise 21.1) that there is a subsequence of (Em) such that the functions Fm converge uniformly on the unit ball of (MN)n. If we repeat this successively for N = 1,2, ... and pass to a further subsequence each time, we obtain a subsequence for which this holds for all N. By Ruan's Theorem (or by the stability of C*algebras under ultraproducts), there is (for some Hoo) an isometric embedding J oo : Eoo -+ B(Hoo), such that, for each N = 1,2, ... , lim IlL ai ® Jm(ei)11 l\fN(B(H m m--+oo
))
= IlL ai ® Joo(edll
MN(B(H=))
uniformly on the unit ball of (MN)n. In particular, we conclude that MN(Jm(Em)) tends to MN(Joo(Eoo)) in the metric space BN2n' We now describe what remains of all this in the category of operator spaces. The main difference is that the "quantized" scalars are the elements of Ie, which is an infinite-dimensional space; hence the compactness arguments are no longer available. Fix an integer n ;::: 1. Let us denote by OSn the set of all operator spaces of dimension n. We consider that two spaces are the same if they are completely isometric. Then OSn can be equipped with the metric
8cb(E, F) = Log dcb(E, F), which is analogous to the Banach-Mazur distance. As in the Banach space case, the space OHn is a center of this metric space (see the above Corollary 7.7):
Introduction to Operator Space Theory
336 Theorem 21.3.
Tllerefore VE,F E OSn
dcb(E,F)::; n.
We recall that, since dcb(Rn' Cn ) = n, this cannot be improved. The analogous result (Le., the optimality of the estimate) for Banach spaces is much more delicate (see [GIl). In another direction, it is natural to wonder whether a given finitedimensional operator space can be realized (at least approximately) as a subspace of the compact operators. For that purpose, we have introduced above the constant dsdE) (see (17.4) and (17.4) '). By Corollary 7.7, we have Theorem 21.4. For all E in OSn, dsdE) ::;
vn.
Quite surprisingly it turns out that this bound cannot be improved asymptotically. Indeed, let us denote
EiJ = (e~)* equipped with its dual operator space structure, or, equivalently,
By Theorem 9.6.1 (or Corollary 8.13), EiJ can also be realized as the span of the n unitary generators in the full C* -algebra of the free group with n generators. Note that when n = 2 we have dsdEiJ) = 1 by Proposition 3.10. However, this no longer holds when n > 2: Theorem 21.5. ([P6}) Let an
= n/(2v'n - 1). We have
an ::; dsdEiJ) ::; (a n )1/2 Note an
rv
vn/2 when n
-+ 00
::;
Vn .
dsdOHn) ::; n 1/ 4 .
and an > 1 as soon as n > 2:
Proof. We may assume EiJ = span[U(gi) /1 ::; i ::; n] as in §9.6. Note that (EiJ)* ~ e~, so that dsd(EiJ)*) = 1. Moreover, by the absorption principle (Proposition 8.1) and by (20.2), we have
21. Local Theory of Operator Spaces. NOllseparability of OSn Applying (19.1)' with u equal to the identity on dsdE[j )2vn-=I, which proves the first line.
Ef"
337
we obtain n :::;
We now turn to OHn . Let (Ti) be an orthonormRI basis of OHi>o Applying (19.1)' with u equal to the identity on OHn and using OH~ ~ OHn , we obtRin
n:::; dsdOHn?
I12: Ti @A(9;)limin 112: 7\ @'\(9i)limin
:::; dsd OHn )2112: T i
@
2 A(g;)11 InlI1 .
= dsdOHn)21/2: '\(g;) @
'\(g;)11 . ; nun
hence, by (20.2), n :::; d s dOHn)2a n , and thus (njon)I/2 :::; dsdOHn). Conversely, by (10.8), since dsdRnnCn ) = 1, we have dsdOHn) :::; n l / 4 .
•
This phenomenon is in sharp contmst with the BRnRch space CRse. Indeed, for every n-dimensional Banach space E and any E > 0, there is an integer N = N(E, n) > 0 and a subspace Fee:;' such that d(E, F) < 1 + E (see Exercise 2.13.2). Thus the analog of dSK(E) in this case is always I! Also, every separable Banach space (a fortiori every finite-dimensional one) embeds isometrically into a single sepamble space, namely, the space C([O, 1]). Thus it is natural to ask: Problem: Is there a single separable operator space X s1tch that every jinitedimensional operator space embeds completely isometrically into X? vVe have just seen that X = K does not work. \Ve will give the answer in Corollary 21.12. But before that, we observe thRt, if X is separable, a simple perturbation argument (see CorollRry 2.13.3) shows that, for any n, the subset of OSn formed of all the n-dimensionRI subspaces of X is R rS cb sepamble metric space. This brings us to the following question rRised by E. Kirchberg:
Problem: Is (OSn, rScb ) separable? A negative answer will gllRrantee thRt there does not exist any "universRl" sepamble operator space X as above. Contrary to the Banach space case, the space (0 Sn, rS cb ) , which is a complete metric space, is not compact if n > 2. Even the subset of all isometrically Hilbertian operator spaces is not compact! There is however a weaker metric structure that one can consider on OSn. For Rny N ::::: 1 and for u: E ---+ F, we denote IluliN
Then, 1::/E, F
E
=
IlhfN
@ uIIMN(E)--->lI/N(F).
OSn we define
dN(E,F)
= inf{lluIlNllu-1IlN I u:
E
---+
F isomorphism}.
338
Introduction to Operator Space Theory
It is easy to check (by a compactness argument) that for all E, Fin OSn we have deb(E, F) = SUpdN(E, F). (21.1) N
Hence deb(Ei , E) -+ 1 iff dN(Ei , E) -+ 1 uniformly in N. But we are interested in the topology for which
We call it the weak topology. It is associated to the metric •
ow(E,F) =
L
TNLog dN(E,F).
N~l
For the weak topology, the space OSn is compact by Lemma 21.7) and a fortiori separable. But the strong topology (the one associated to deb) is strictly stronger than the weak one, so that the identity
is discontinuous (at least if n > 2). But if OSn is assumed oeb-separable, the function f must be in the first Baire class. Indeed, by (21.1), if B is any closed ball in (OSn,Oeb), then f-l(B) is closed in (OSn, ow); therefore, for any (strongly) open set U, f-l(U) must be (weakly) an Fu-set, and the latter property characterizes functions of Baire class 1. Quite curiously, Kirchberg's question was first answered by applying the classical Baire theory to this map f. Theorem 21.6. ({JPJ) For n > 2, the metric space (OSn,Oeb) is nonseparable. We will use the following elementary facts. Lemma 21.7. The metric space (OSn,ow) is compact.
•
Proof. This is an easy consequence of Remark 21.2. Lemma 21.8. For any E, F in OSn we have /
and for any N 2: 1
Proof. The first equality is an immediate consequence of (2.3.3). Similarly, the second one follows from (2.3.4). •
21. Local Theory of Operator Spaces. Nonseparability of OSn
339
Lemma 21.9. For any E in OSn, there is a sequence {Em} in OSn tending weakly to E and such that Em is "matricial" (i.e .. Em C lIIN for a sllitable N = N(m)) for any m.
Proof. Since E is separable, we may clearly assume that E C B(H) with H separable. Then let {Hm} be an increasing sequence of finite-dimensional subspaces of H with UHm dense in H. Let Vm: E ---> B(Hm) be defined by vm(e) = PH",eIH"" and let Em C B(Hm) be the range of V m . Since (see Exercise 2.1.1) Ilvm(e)11 i l!ell, we have
for any e, e' in E, and since the unit ball of E is compact, we must have Ilvm(e)11 ---> lIell uniformly on BE. In particular, when m is large enough, Vm is an isomorphism from E to Em, and moreover d(E, Em) ---> 1 when 111. ---> 00. lVIore generally, for any fixed N ::::: 1, the mappings (Vm)N = hIN IS! ---> MN(Em) satisfy
Vm: MN(E)
for any x in lIIN(E). Repeating the same argument, we find that the convergence is actually uniform on the unit ball of lII N (E) and hence dN(E, Em) ---> 1 when 111. ---> 00. Thus we conclude that Em tends weakly to E. • We will also need to characterize (following ([P6]) the points of continuity of
f.
Lemma 21.10. An element E E OSn is a point of contilluity of the map iff dsdE) = dSK:(E*) = 1.
f
Proof. Let {Ern} be as in the preceding lemma. Then, if E is a point of continuity, Em must tend strongly to E, and hence deb (Ern , E) ---> 1. But since Ern is matricial, we have dsdErn) = 1; hence we also obtain dsdE) = 1. Applying Lemma 21.9 to E*, we find a sequence of matricial spaces {Em} that tends weakly to E*. Then, by Lemma 21.8, {E,';,} tends weakly to E; but if E is a point of continuity, the convergence must be strong, so that, by Lemma 21.8 again, {Ern} also tends strongly to E*, whence dsdE*) = 1. Conversely, assume dsdE) = 1. Fix E > O. Then there is Nand E C lIIN such that dcb(E, E) < 1 + E. By Proposition 1.12, this implies that for any F and any u: F ---> E we have Ilulleb ::; (1 + E)lluIiN. Now, if dsdE*) = 1, we may use (2.3.4) and we find by the same argument an integer N' such that, for any F and any v: E ---> F, we have Ilvllcb::; (1 +E)lIvIiN" Replacing
Introduction to Operator Space Theory
340
N by the largest of N, N', we obtain that for any F and any isomorphism ~ E we have
u: F
Thus we obtain From this last result, it is clear that if F tends to E weakly, that is, dN(E, F) ~ 1 for any N), then F to E strongly (Le., dcb(E, F) ~ 1). • Then we can use Chapter 19 to obtain: Lemma 21.11. Let E E OSn be the weak limit of a sequence (En) in OSn such that dsdEn) = dsdE~) = 1, for all 0:. Tilen d(E, £'2) ::; 4. Moreover, for any biorthogonai system (ei' ~i) (1 ::; i ::; n) in E x E*, we have
(21.2)
Proof. By the last assertion of Corollary 19.2 (applied to the identity on En), we have d(En' £2) = 'Y2(IEo ) ::; 4. Since weak convergence implies a fortiori d(En' E) ~ 1, we obtain the limit d(E, £2) ::; 4. Moreover, to show the second part, let X = C~(lFoo). Recall (Corollary 17.10) that X is exact with dsdX) = 1. By Exercise 21.5 applied with X = C~ (F00)' we have for each 0: a basis {ei (0:) I 1 ::; i ::; n} of En with dual basis {~i (0:) I 1 ::; i ::; n} in E~ such that
li~ IlL -'(gi) 0 ei(o:)llmin = IlL -'(gi) 0 eil min li~ IlL -'(gd 0 ~i(o:)llmin = IlL -'(gi) 0 ~illmin . By Theorem 19.1 applied to u
= lEn' we have
Hence passing to the limit we obtain
as announced.
•
21. Local Theory of Operator Spaces. NOllseparability of OSn
341
First Proof of Theorem 21.6. Assume OSn strongly separable. Then .f is of the first Baire class and OSn is weakly compact and hence weakly Baire. By Baire's famous theorem (see [Ku, 31, X, Theorem 1, p. 394]), .f must have a (weakly) dense set of points of continuity, but this is impossible, at least if n > 2. Indeed, by Lemma 21.11, the spaces appearing in Lemma 21.10 are too few to be weakly dense in OSn. To verify this, it suffices to produce a single element of OSn failing the conclusion of Lemma 21.11. Obviously, we cannot have d(E, e~) ~ 4 for any E in OSn, at least if n > 16: Indeed it is easy to check that d(ef,e~) = and this is> 4 when n > 16 !
vn,
l\lore precisely, for any n > 2, one can produce an E in OSn failing (21.2): We set E = E~, that is, E = span[A(g;) I i = 1, ... , n], and X; = A(gi). Then, by Remark 8.10 and (20.2), we have 112::7 A(g;) 0 :r;11 = 112::7 A(g; )11 = 2vn=-I, and since 2:: A(gi)0ei represents the identity on E, we must have (by (2.3.2)) 112::7 A(gi) 0 II = IIIdeb = 1. Hence, (21.2) implies n ~ 2vn=-I, or, equivalently, n ~ 2. •
e;
We will give two additional proofs of Theorem 21.6; see Theorem 21.14 and Corollary 21.15.
Remark. The only known examples of spaces E such that dsdE) = dslC (E*) = 1 are (in dimension> 1) the spaces e~,.e~ (say with their minimalactually unique-o.s.s.), the spaces Rn and C n (n > 1) or the spaces Rand C, and also the spaces
Problem. Are the spaces e~, e~, Rn, Cn, C EB Rn, C EB Cn, C EBl Rn, C EBl C n (resp. R, C, C EB R, C EB C,
342
Introduction to Operator Space Theory
Actually, the original proof in [JP] also yields the nonseparability of the subset of OSn formed of all the (isometrically) Hilbertian operator spaces. Even the further subset of all the "homogeneous" Hilbertian operator spaces, in the sense of §9.2, fails to be separable if n > 2. When n = 2, the nonseparability remains an open question. Note, however, that if we restrict ourselves to operator spaces E spanned by n linearly independent unit aries, then n > 2 is necessary, since in the case n = 2 it suffices to consider the spaces E spanned by 1 and a unitary, and these lie in a commutative C* -algebra; consequently they are minimal operator spaces, and hence they satisfy dSA(E) = 1 with A = Co (see Exercise 2.13.2) and therefore (Corollary 2.13.3) form a separable collection. Let Eo be an n-dimensional Banach space. Let OSn(Eo) be the subset of OSn formed of all the spaces isometric to Eo. (Equivalently, OSn(Eo) represents the set of all possible operator space structures on Eo compatible with its norm.) This set may be a singleton (for instance, if Eo = €~; see Chapter 3), in which case it is compact and separable. Curiously, the converses are open: Problem. Consider the following properties of an n-dimensional normed space Eo. (i) OS(Eo) is separable. (ii) OS(Eo) is compact. (iii) OS(Eo) is a singleton. A re these properties equivalent if We now generalize the numbers dsK.{.): For any operator space X (actually X will often be a C*-algebra) and any finite-dimensional operator space E, we introduce (21.3) dsx(E) = inf{dcb(E, F) I Fe X}. Of course, if X = B(H), dsx(E) = 1 for all E. By Corollary 17.5, if A is an exact C*-algebra, we have (21.4) for all E. Moreover if A is roughly "large enough" (precisely, if A contains, for each n and e: > 0, a subspace Xn with dcb(Xn, Mn) < 1+e:), then the converse also holds. Thus the number dSA(E) is the same for all '!1arge enough" nuclear or exact C* -algebras. In [Ki2] , Kirchberg's candidate for a "universal" separable X as above was the full C*-algebra C*(Foo) of the free group Foo with infinitely many generators, which the next result invalidates. Corollary 21.12. Ifn > 2, there does not exist any separable operator space X such that dsx(E) = 1 for any E in OSn o
21. Local Theory of Operator Spaces. Nonseparability of OSn
343
Proof. If X is separable, the subset {E E OSn I dsx(E) = 1} also is • separable (by Corollary 2.13.3); hence this follows from Theorem 21.6. In the next chapter, we will concentrate on the special case when X
=
C*(Foo), and to simplify the notation we will set
We now wish to measure the "degree of nonseparability" of OSn. For that purpose, we introduce, for any n 2: 1, the number 8(n), which is the infimum of the numbers 8 > 0 for which OSn admits a countable 8-net. More precisely, we prefer to use dcb instead of 8cb, so we define for any E E OSn and any subset V c OSn dcb(E, V)
= inf{dcb(E, F) I FE
V}.
Then we have by definition
8(n)
inf
=
(21.5)
sup dcb(E, V).
V countable EEDS"
Of course, we may replace "countable" by "separable" in this definition. Thus, for any separable operator space X, considering V = {E C X I dim E = n}, we obtain (by Corollary 2.13.3)
8(n)::; sup dsx(E).
(21.6)
EEDS"
The second and third proofs of Theorem introduced in the preceding chapter.
2~.6
make use of the constant C(n),
Recall that we have trivially C(1) = 1 and (exercise) C(2) =:= 2, but for n > 2 we have C(n) < n, and this is crucial to estimate 8(n), as the next result shows:
Theorem 21.13. For all n 2: 2, we llRve n/C(n) ::; 8(n). Proof. Let C be any number with C > C(n). By definition of C(n), there is a sequence of n-tuples of unitary matrices {Ui (m)} such that
,:~~,
{lit >-1
ui(m) 0 Ui(m/)11 . } ::;
(21. 7)
c.
min
Let ei (1 ::; i ::; n) be the canonical basis of e~. We define for any
nc
N
344
Introduction to Operator Space Theory
and we let Note that the presence of ei ensures that Eo has dimension n. Now fix 8 > 8(n). We will show that 8 2:: n/C. By defi.nition of 8(n), there is a sequence {Em} in OSn such that, for any E in OSn, there is an integer m such that dcb(E, Em) < 8, and in particular for any 0 there is an m with dcb(Eo, Em) < 8. Since the set of all 0 is continuous (Le., has the cardinality of the continuum), there is a continuous collection C of subsets of N for which the same m (fixed from now on) must be used, that is, such that for each 0 in C there is a map Vo: Eo --+ Em satisfying: (21.8) Now consider the continuous family (VO(Xi(O»)i~n of n-tuples of elements of Em. Since (Em)n is norm-separable, for any Tl > 0 there exists a continuous subcollection Cl C C such that, for all 0,0' in C1 , we have n
E Ilvo(Xi(O»
(21.9)
- vo,(xi(O'»11 < Tl·
i=l
A fortiori, C1 has cardinality> 1; hence we can find 0,0' in C1 such that 0' rt. 0, so that there is an integer m' E 0' with m' r:t o. This implies, by the definition of Xi(O)
IIExi(O) 0 ui(m')11
. = max min
{c,
sup mEO
lit
ui(m) 0
Ui(m')11
i=l
and, on the other hand, by (20.1), since m' EO', we have
Thus, by (21.8),'we have on one hand
and on the other hand
n~ Ilvn,lllcb IlL VO,(Xi(O'» 0 Ui(m'1!Lin ~ ilL VO'(Xi(O'» ui(m')II, ®
which gives by (21.9)
} min
~C
21. Local TI1eory of Operator Spaces. Nonseparability of OSn hence n ~ 8C + 1]. Finally, since 17 > 0 is arbitrary, we obtain n therefore n/C(n) ~ 8(n).
~
345 8C, and •
Remark. A simple modification of the preceding proof shows that the constant (3(n) ::::: n/C(n), introduced in Remark 20.11, satisfies (3(n) ~ 8(n). Thus, if we want to show that, for n large, OSn is "very" nonseparable, that is, that 8(n) is very big, we are reduced to showing that C(n) is "much smaller" than 11. Here is what we have seen in the preceding chapter. Theorem 21.14.
(i) For all n ::::: 3, C(n) < n; hence 8(n) > 1. (ii) Afore precisely, for all
11
:::::
2, C(n)
2vn=-I; hence 8(n) >
n/(2vn=-I)(::::: ,fii/2). Corollary 21.15. Fixn::::: 1. Then, for any 8 < 8(n), there is an uncountable family (Ei) in OSn such tllat (21.10)
Vi
I- j
In particular, tllis holds for some 8 > 1 for aJI n ::::: 3. Proof. This follows by a standard argument: Let (Ei) be a maximal collection in OSn satisfying (21.10). Let D = {Ed. By maximality we have dcb(E, D) ~ 8 for all E in OSn, and since 8 < 8(n), the set D cannot be countable. • Remark. Actually, "countable" can be replaced by "with cardinality less than the continuum" in the proofs, so we effectively obtain a continuous collection satisfying (21.10). Remark. By Theorem 21.3 (or Theorem 21.4) we have 8(n) ~ ,fii; hence the above lower bound for 8(n) is sharp, at least asymptotically. However, it might be that Corollary 21.15 itself can be considerably improved: Perhaps its conclusion is even true for all 8 < n. A result very close to this conjecture has been recently established in fORi]. We will now apply Theorem 21.13 and (21.6) to the special case X
C*(Foo). To simplify the notation, we set
Then, recalling Theorem 21.14, we find: Corollary 21.16. For any n ::::: 2 we have n/C(n) ~ sup{df(E) lEE OSn}.
=
Introduction to Operator Space Theory
346
In particular, for any n > 2 there is an E such that df (E) > 1. Problem. Find an explicit example of an E such that df (E) > 1. Remark. The class of all n-dimensional operator spaces E such that df(E) = 1 has surprisingly nice stability properties. This class is studied in detail in [Harl]. We will see in (22.4) that it is stable by duality. In addition, it is proved in [Harl] that this class is stable under the Haagerup tensor product and under complex interpolation. More precisely, we have (see [Harl] for details): (i) For any operator spaces El, E2 we have
and df (E I
®min
E2)
::;
df (E 1 )dsK.{E2 ).
(ii) For any compatible pair (Eo, E 1 ) of n-dimensional operator spaces we have df«Eo, E 1 )6) ::; df(Eo)I- 6df(Et)6.
In particular, df «Eo,E1 )6)
=1 if
df(Eo)
=
df(Et)
=
1.
These stability properties (in particular the last one) explain why all the natural examples belong to the class of operator spaces E such that df (E) = 1. Indeed, first observe that all exact operator spaces (for instance, K, min(e 2 ), e[O, 1]' ... ) are in this class by (22.2) below. Then the stability under duality shows that the class also contains, for example, SI, max(e 2 ), LIfO, 1]. Finally, using interpolation, we find that Sp, Lp, . .. also belong to the class for all 1 < P < 00. In particular, the operator Hilbert space OH is in this class, that is, we have
Exercises Exercise 21.1. Let (Pm) be a sequence of norms (or seminorms) on a finitedimensional vector space E. If (Pm) converges pointwise, then it automatically converges uniformly over any bounded subset of E. Exercise 21.2. Let Em (m ~ 1) and E be n~diQ1ensional Banach spaces. Show that the following assertions are equivalent: / (i) d(Em, E) -+ 1 when m -+ 00. (ii) For any basis {ei I 1 ::; i ::; n} in E there is, for each m, a basis {ei(m) 11 ::; i ::; n} in Em such that \/X=(Xi)i$n Ecn
"!~ooll~Xiei(m)11 =llfxieill· Em
1
E
21. Local TIleory of Operator Spaces. Nonseparability ofOSn
347
(ii)' For some basis, the same as (ii) holds. (iii) For any free ultrafilter U on N, we have
TIEm/U = E (isometrically).
Exercise 21.3. Let Em (m ~ 1) and E be n-dimensional operator spaces. Show that the following are equivalent. (i) dN(Em , E) ---> 1 for any N ~ 1 (i.e., Em tends weakly to E in OSn). (ii) For any basis (ei) in E, there is, for each Nand nI, a basis {ef"(m) 1 :::; i :::; m} in Em such that, for any n-tuple (ai)i
I
(ii)' For some basis (ei) the same as (ii) holds. (iii) For any free ultrafilter U on N, we have TIEm/U = E (completely isometrically) .
Exercise 21.4. Let {ei(m) 11 :::; i :::; n} and {e; 11:::; i:::; n} be bases in Em and E, respectively, such that, for any N and any n-tuple (a;)i~n in !lIN, we have = lim m--->oo I1/N(E",) I1/N(E)
liLa; ®ei(m)11
IILai ®eill
.
Let (~i (m)) and (~i) denote the biorthogonal basis of E:n and E* , respectively. Then show that
II" ~ ai ® ~i(m)11 = II" ~ill . ~ ai ®
lim
m--->oo
lII,v(E:,,)
lII,v(E*)
Hint: Recall Proposition 1.12.
Exercise 21.5. Let {ei(m) I 1 :::; i :::; n} and {ei I 1 :::; i :::; n} be as in Exercise 21.4. Let X be any exact operator space with C = dsdX). We then have for any (Xi)i:S;n in xn C- I
IlL ®eill : :; liminf IlL ®ei(m)11 : :; lim Slip IlL ®ei(m)11 :::; C ilL ® ei II· Xi
Xi
Xi
In particular, if dsdX)
= 1,
we have
Xi
Chapter 22. B(H)
Q9
B(H)
Recently, E. Kirchberg [Ki2, Ki5] revived interest in the study of nuclear pairs of C*-algebras, that is, pairs (A, B) such that A Q9min B = A Q9max B, or equivalently, such that there is only one C* - norm on the algebraic tensor product A Q9 B. Recall that this happens for all B iff A is nuclear (see Chapters 11 and 12). The C*-algebras C*(lFoo) and B(H) are typical examples of nonnuclear C* -algebras. Their nonnuclearity was first proved respectively in [Ta2] and [Wa1]. (However, this is now quite clear since every separable C*-algebra is a quotient of C* (IF 00) and a subalgebra of B( H). Since nuclear C* -algebras have nuclear quotients and exact subalgebras, it suffices to know the existence of a single nonexact C* -algebra, and the above Theorem 21.5 clearly guarantees that C*(lFoo) is not exact.) Therefore, Kirchberg's Theorem (already proved above as Theorem 13.1) came as a surprise: For any free group IF (and any H) we have C*(IF) Q9min B(H) = C*(IF) Q9max B(H). For a long time, the following question remained open: If a C* -algebra A satisfies A Q9min AOP = A Q9max AOP, (22.1) is A nuclear? In [Ki2], Kirchberg finally gave a counterexample, using Theorem 13.1 as a starting point. Recall the following generalization of Theorem 13.1 proved above as Theorem 16.2: Theorem 22.1. Let A, B be C* -algebras. If A has the LLP and B the WEP, then A Q9min B = A Q9max B.
In particular, if A has both the LLP and the WEP, then (22.1) Kirchberg also showed that (22.1) implies that A has the WEP. This was clarified by Haagerup [H4] (see Theorem 15.6), who showed that a C*-algebra A has the WEP iff for any finite sequence (ai) ih:~ we have
II 2: a i Q9aiIlA®mlnAop = II 2: a i ®d:IIA®maxAop. Thus, to produce a nonnuclear example satisfying (22.1), it would suffice to solve positively the following. Problem. Is there a nonnuclear C* -algebra with both the WEP and the LLP?
22. B(H) ® B(H)
349
Kirchberg's construction in [Ki2] (of a nonnuclear C* -algebra satisfying (22.1» comes very close, but unfortunately the preceding problem remains open. Nevertheless, as Kirchberg [Ki2] pointed out, there might be much simpler examples, and in that direction he raised in [Ki2] the following problems. Problem A: Does (22.1) holds if A = B(H)? Problem B: Does (22.1) holds if A = C*(IFoo)? Note that in these two cases A and AOP are isomorphic C* -algebras, so that Problem A (resp. B) can be reformulated simply as: Is there a unique C*-norm on B(H) ® B(H) (resp. C*(IFoo) ® C*(IFoo»? As we will see in the following, the nonseparability of OSn, discussed in the preceding chapter, gives a negative answer to Problem A. However, Problem B is still open. We now wish to study the finite-dimensional operator su bspaces of C* (IF 00)' in analogy with our study of the subspaces of IC in Chapter 17. Consider an n-dimensional subspace E C C*(lFoo) and a C*-embedding C*(IFoo) C B(1t). Then, for any u in E ® B(H), we have by Theorem 22.1 liuIIB(?-l)0lUBXB(H)
~
Ilullc*(lFx)0maxB(H) = Ilullc*(lFx)0lUiIlB(H) = Il u IIE0IUiIl B (H)
Thus the min-norm on E®B(H) coincides with the norm induced by B(1t)®rnax B(H). It turns out that this property characterizes subspaces of C*(IFoo). Recall that we denote
By Corollary 21.12, we already know that, for any n > 2, there are ndimensional operator spaces E such that df(E) > 1. Moreover, by (21.6) and Theorem 21.13, we know this number can actually be fairly large when n itself is large. Following [JP], we will now relate t.his to the ratio between the minimal and maximal norms on tensors of rank n in B(H) ® B(H).
Theorem 22.2. ([JP)) Let a ~ 0 be a constant and let X C B(1t) be all operator space. The following are equivalent. (i) df(E) ~ a for all finite-dimensional subspaces E C X. (ii) For any H and any operator space F C B(H), we llave \/u E E®F
(iii) The same as (ii) with H = £2 and F = B(£2)' Corollary 22.3. ([JPJ) Let E C B(1t) be an n-dimensional operator space. Then
350
Introduction to Operator Space Theory
Remark. Note that, using (17.5), this shows in particular that df(E) ~ dsdE)
(22.2)
for any finite-dimensional operator space E. In particular, by Theorem 21.4, this implies (22.3) \IE E OSn
Remark. It is observed in [JP] that, for a C* -algebra A, the condition df(A) = 1 ensures that any finite-dimensional subspace E c A is completely isometric to a subspace of C*(Foo). Note, however, that the algebra A need not embed as a C*-algebra into C*(Foo). For instance, let A be the Cuntz algebra (see §9.4). It is nuclear and hence a fortiori exact, so that df(A) = dsdA) = 1; however, A does not embed into C*(Foo), because C*(Foo) embeds into a direct sum of matrix algebras (see Exercise 13.1); hence left invertible elements in it are right invertible, and the latter property obviously fails in the Cuntz algebra. Nevertheless, by the Choi and Effros lifting theorem (cf., e.g., [Wa2, p. 53]; see also Remark 16.3(v)) there is a unital completely positive (and completely contractive) factorization of the identity of the Cuntz algebra (or any separable nuclear C*~algebra) through C*(Foo). The preceding theorem leads naturally to the following. Definition. Let El C B(Hl ), E2 c B(H2 ) be arbitrary operator spaces. We will denote by II II M the norm induced on El 0 E2 by B (H 1) 0 max B (H2) and by El 0M E2 its completion with respect to this norm. Clearly, El 0111 E2 can be viewed as an operator space embedded into B(Hl ) 0 max B(H2). It can be checked easily, using the extension property of c.b. maps into B(H) (Corollary 1.7), together with (11.5) and (11.7) (see [JP] for more details) that II IL,\[ and El 0111 E2 do not depend on the particular choices of complete embeddings El C B(Ht}, E2 c B(H2). Moreover, we have Lemma 22.4. Let F l , F2 be two operator spaces. (i) Consider c.b. maps Ul: El ---* Fl and U2: E2 ---* F2. Then u10u2 defines ac.b. map from E 1 0 ME 2 to F1 0 M F2 witl] IIUl 0 u21IcB(E 1 ®A1E2 ,F1 ®A1F2 ~ Ilulli cb II U211 cb ' (ii) IfUl andu2 arecompleteisometries, then u10u2: E 1 0 ME 2 ---* F 1 0 is also a complete isometry.
M
F2
Remark. When E l , E2 are C*-algebras, El 0~E2 can be identified with a C*-subalgebra of B(Ht} 0 max B(H2 ), so that thi!f tensor product 01\/ makes sense in both categories of operator spaces and C* -algebras. The next result analyzes more closely the significance of IIuII1I1 < 1 for U E E 0 F. It turns out to be closely connected to the factorizations of the associated linear operator U: F* ---* E through a subspace of C*(lFoo ). This can be considered as analogous to Theorem 17.1 or Theorem 17.2 but with C*(lFoo ) in the place of /C.
22. B(H) ® B(H)
351
Proposition 22.5. Let E, F be operator spaces; let 11 E E ®' F, and let U: F* --+ E be tlle associated finite rank linear operator. Consider a finitedimensional subspace S C C*(IFoo) and a factorization ofU oftlle form U = ba witll bounded linear maps a: F* --+ S and b: S --+ E, wllere a: F* --+ S is weak-* continuous. Tllen
wllere tlle infimum runs over all sucll factorizations of U. Proof. Ass~lme E C B(H) and F C B(K) with H, K Hilbert. It clearly suffices to prove this in the case when E and F are both finite-dimensionaL Assume U factorized as above with Iiallcblibl/cb < 1. Then, by Kirchberg's Theorem 13.1, the min- and max-norms are equal on C*(IF",,) ®B(K). Hence, by Lemma 22.4(ii), we have isometrically S®minF = S®AI F, so that if is the element of S®minF associated to a, we have I/al/AI = Iialicb and u = (b®IF )(a). Therefore, by Lemma 22.4(i), we have IluliAl ~ Ilbl/cbllallAl ~ Iiallcbl/blicb < 1. Thus we obtain l/ullAl ~ inf{llallcbllbllcb}.
a
We now turn to the converse inequality. Assume l/ullAl < 1. Let G be a large enough free group so that B(H) is a qnotient of C*(G), and let q: C*(G) --+ B(H) be the quotient *-homomorphism. By the exactness of the maximal tensor product (see Exercise 11.1), if we view u as sitting ill B(H) ® B(K), there is a lifting ti E C*(G) ® B(K) of u with litili max < 1. A fortiori litili min < 1. By Lemma 17.2, since u E B(H) ® F, there mllst actually exist a lifting ti in C*(G) ® F with I/till min < 1. Let S C C*(G) be a finite-dimensional subspace such that ti E S ® F. Note that since S is separable (say), there is a subgroup G 1 C G isomorphic to IF"" such that S C C*(G 1 ) ~ C*(IFoo). Let a: F* --+ S be the linear map associated to ti, and let b be the restriction of q to S. Since ti lifts u, we have U = ba, Ilbllcb ~ Ilqllcb ~ 1, and Iialicb = litili min ~ litil/max < 1. Thus we conclude inf{llallcbllbllcb} ~ IlulIAl. • Applying this result to U
= IE,
we obtain
Corollary 22.6. Let E be a finite-dimensional operator space. Let iE E E ® E* be tlle tensor associated to the identity on E. Then
In particular, we have (22.4) The next result was pointed out to me by N. Ozawa.
352
Introduction to Operator Space Theory
Corollary 22.7. For any finite-dimensional operator space E, tllere is a subspace of E c C* (IF 00) and an isomorpllism u: E ~ E such that
In particular, E satisfies df(E) = 1 iff E is completely isometric to a subspace ofC*(lFoo). • Proof. Indeed, assume E c B(H), E* c B(K) and let t E E* 0 B(H) be the tensor representing the inclusion E C B(H). By the preceding, we have df(E) = IItIlB(K)®maxB(H)' By the proof of Proposition 22.5 (and with the same notation), df(E) coincides with the infimum of IITIIE*®lUlnC*(G) over all T in E* 0 C*(G) lifting t. By Lemma 2.4.5, this infimum is actually attained: There is a T with IITllmin = df(E). But now T defines a linear map T1: E ~ C*(G) with IITlllcb = df(E). Letting E = T1(E), and with u = Tl viewed as acting from E to E, we find Ilulicb = df(E) and Ilu-11lcb ::; IIqllcb ::; 1.
•
Remark. (Due to N. Ozawa) Using Arveson's ideas in [Ar4], it is possible to show more generally that, if a separable operator space X is such that df(E) = 1 for every finite-dimensional subspace E C X and if, in addition, X admits a net of completely contractive finite rank maps tending pointwise to the identity, then X embeds completely isometrically into C*(lF oo ). This follows from Exercise 2.4.2. In particular, OR embeds completely isometrically into C*(lFoo ). Unfortunately, however, we cannot describe the embedding more explicitly. We can now relate the nonseparability of OSn with the possible C* -norms on B(H) 0 B(H), as follows (we denote by rk(u) the rank of u). Proposition 22.8. Let H = £2. For any n ::::: 1 we define
Then .\(n)
= sup{df(E) lEE OSn}
and~~n)::::: 8(n).
(22.5)
Proof. Any u in B(H) 0 B(H) with rank at mostn can be viewed as an element of E 0 B(H) for some n-dimensional subspace of B(H). Then (22.5) follows immediately from Corollary 22.3 and (21.6). • We can now exploit the results of the previous chapter to settle Problem A.
22. B(H) 0 B(H)
353
Theorem 22.9. lVe have B(H) 0 m in B(H) =I- B(H) 0
max
B(H).
(22.6)
More precisely: (i) A(n) :::; "fii for all n. (ii) A(n) > 1 for all n > 2. (iii) A(n) ~ n/(2Jn=l) ~ "fii/2 for all n ~ 3.
Proof. (i) follows from (22.3) and (22.5), and (ii) and (iii) follow from Theorem 21.14 and (22.5). Of course (ii) implies (22.6) since the equality can only • be isometric (cf. Proposition 1.5). Remark. Kirchberg proved in [Ki2] that if his conjecture 16.12 is correct, then, for any C* -algebra A, for any E > 0, and for any finite-dimensional operator subspace E c A*, there is a subspace E c SI with deb(E, E) < 1 +E. In particular, it would follow (by Corollary 2.13.3) that the metric subspace en c OSn formed by the n-dimensional subspaces of duals (or preduals) of C* -algebras would be deb-separable for any n. Thus, to disprove conjecture 16.12, in analogy with Theorem 22.9, one could have hoped to show that en is not separable. Unfortunately, this totally fails; indeed, en is deb-separable ([JUd]), and actually it is even compact ([Ozl])!
Chapter 23. Completely Isomorphic C* -Algebras It is natural to wonder whether two C* -algebras that are completely isomorphic must be isomorphic (as C*-algebras). The answer is negative. For instance, Christensen and Sinclair (see [CS5]) (extending previous remarks of Haagerup and Lindenstrauss) proved that any injective (Le. hyperfinite [Col]) von Neumann algebra with separable predual is completely isomorphic either to C~ (n EN), Coo or B(H). So B(H) and, for example, EBn lIIn provide a negative answer to the above question. (See also [RW] and [Blo] for related results.) In a different direction, A. Arias [AI] proved that the reduced C*-algebras C~(Fn) associated to the free groups with n generators are mutually completely isomorphic, that is, CHFn) ~ C~(H) as operator spaces for any n, k. Similarly, the von Neumann algebras are all mutually completely isomorphic. However, it is known from K-theory that C~(Fn) is not C*-isomorphic to C~(Fk) when n -:/= k (see [PVI-2]). The analogous nonisomorphism in the von Neumann case remains a major outstanding open problem. These negative results suggest that completely isomorphic C* -algebras might be quite different. Nevertheless, several basic structural properties of C* -algebras, such as nuclearity, injectivity (in the von Neumann case), the weak expectation property (in short, \VEP), and exactness, are preserved under complete isomorphisms. The rest of this chapter is devoted to this. We introduced nuclearity in Chapter 11. It is known (see Theorem 11.5) that a C* -algebra A is nuclear iff the identity on A is approximable by a net (Ui) of finite rank maps, of the form A..!~Mn;~A such that SUPiEI Iladlcbllbillcb :::; 1. A priori, this property does not seem stable under complete isomorphism. However (d. [PI]), it turns out that A is nuclear iff there is a constant C for which the identity of A is approximable by a net as above with SUPiElliaillcbllbilicb :::; C. The latter is clearly invariant under complete isomorphisms. A similar situation reappears for injective von Neumann algebras. A von Neumann algebra M C B(H) is called injective if there is a completely contractive and completely positive projection P: B(H) ---4 M. (Actually, by a result of Tomiyama [Tol], in this situation, any contractive P is automatically completely both contractive and positive.) It was proved in [P3, PI] and independently in [CS3] that At is injective as soon as there is a constant C and a projection P: B(H) ---4 M with IlPlicb :::; Q, "This was extended with more general von Neuma~n algebras in the place of B(H) in [P7, CS4, H4]. A similar remark is valid for the class of C* -algebras with WEP (Le., those A for which the canonical inclusion iA: A ---4 A** factors completely positively and (completely) contractively through B(H)j see Chapter 15). Recapitulating, all these statements give us
23. Completely Isomorpllic C* -Algebras
355
Theorem 23.1. Let A, B be two C* -algebras that are completely isomorphic. Then:
(i) If A is nuclear, so is B. (ii) If A, B are von Neumann algebras, and if A is injecti\Te, so is B. (iii) If A has the WEP, so does B. (iv) If A is exact, so is B. Proof. (i), (ii), and (iii) are immediate consequences of the preceding facts. Concerning (iv), note that A is exact iff there is a constant C such that, for all finite-dimensional subspaces E c A, we have dsK{E) ::; C (see, e.g., Corollary 17.5). From this (iv) follows, but it was already a consequence of Kirchberg's • earlier work. Remark. Concerning the injectivity of a von Neumann algebra 111 c B(H), it is unclear whether Af must be injective if there merely exists a bounded projection from B(H) onto .M. For some results in this direction see [P13 , HP2]. As observed in [P13] in the C*-case, for any operator space Xc B(H) the projection constant
"\(X) = inf{IIPIlI P: B(H)
-+
X, projection onto X}
is invariant under completely isometric isomorphism. (The proof uses the extension property of c.b. maps into B(H); see Corollary 1.7.) In particular, "\(X) does not depend on the particular embedding Xc B(H). Analogously, if Y is another operator space, we have ..\( X) ::; ..\(Y)dcb( X, Y). Let All, l'vf2 be two von Neumann algebras and let Af101112 denote their von Neumann algebraic tensor product. It is proved in [P13] that
Similar supermultiplicative estimates can be proved for certain constants related to the approximation property; see [SSl]. Of course similar estimates hold for the com.pletely bounded projection constant in B(H) defined as ..\cb(X) = inf{llPllcb}, and ..\cb(X) ~ "\(X), but not much is known about the relationship between these two constants even when X is a C* -algebra.
Chapter 24. Injective and Projective Operator Spaces This chapter is mainly a survey without proof. In general, injective spaces are those that satisfy an extension property while projective spaces satisfy a lifting property. The extension property corresponds to the following diagram: Y
U
s
'\.
~
X
where S is a subspace of Y and u is a "morphism" that we wish to extend to the whole of Y. When this is possible for any S, Y, and u with values in X, we say that X has the extension property or is injective. This property is interesting in various categories. See, for example, [MN, p. 170J for Banach lattices, [CE3J for operator systems, [Li, Lac, Bou5J for Banach spaces, and [Col J or [KRJ for von Neumann algebras. In all these cases, injective objects playa fundamental role. Thus it is not surprising that they should also be of interest in the operator space category. There, of course, X, Yare operator spaces, S C Y is a closed subspace, and u is c.b. We will say that an operator space X has the extension property (or Y -4 X so is injective) if any c.b. map u as above admits a c.b. extension that the resulting diagram commutes:
u:
Y
u S
~
X
u
It is easy to see that when this holds there is a constant A such that can always be found with lIuli c b ~ Allullcb. We then say that X is A-injective. For example, by Corollary 1.7, we know that B(H) is I-injective for any H. For Banach spaces, £00(1) is I-injective for any set I, and any Banach space embeds isometrically into £00 (I) for some I; therefore, in the Banach category, A-injective spaces are just the A-complemented subspaces of £00(1). Since B(H) is also "universal" for operator spaces (see §2.12), the analogous statement for operator spaces is immediate, as follows.
Proposition 24.1. Let A 2:: 1. The following prop~ties of an operator space X are equivalent: -"-
(i) X is A-injective. (ii) For any completely isometric embedding X c Y into an operator space Y, there is a c.b. projection P: Y - 4 X with IIPlicb ~ A. (iii) There is a completely isometric embedding Xc B(H) and a c.b. projection P: B(H) - 4 X with IlPlicb ~ A.
24. Injective and Projective Operator Spaces
357
(iv) TIle identity on X admits, for a suitable H, a factorization tIlrough B(H) of the form X ~B(H)~X with IIvllcbllwllcb :::; A. Proof. (i) ~ (ii): Let u onto X, whence (ii). (ii)
~
(iii) and (iii)
= Ix. ~
Any extension
u:
Y ---- X must be a projection
(iv) are obvious.
(iv) ~ (i): Consider u: S ---- X and let T = vu: S ---- B(H). Since T is B(H)-valued, by Corollary 1.7, T admits an extension T: Y ---- B(H) with IITllcb = IITllcb. Then the operator = wT: Y ---- X extends u and satisfies Ilulicb :::; IIwllcbllTllcb = IlwllcbllTlicb = Ilwllcbllvullcb :::; Ilwllcbllvllcbllullcb :::; Allullcb. •
u
The particular. case A = 1 is really special. In the (real or complex) Banach category, a space X is I-injective iff X is isometric to the space of (real- or complex-valued) continuous functions on an ext rem ally disconnected compact (also called Stonean) space (cf., e.g., [Lac, p. 92]). If, moreover, X is a dual space, then this holds iff X is isometric to Loo(O, f.1) for some measure space (0, f.1). In the operator space framework, the I-injective objects were characterized as follows by Ruan ([Ru2]), using some important previous work by Hamana [Ham].
Theorem 24.2. ([Ru2]) An operator space X is l-injective iff tlwre is a l-injective C* -algebra A and projections p, q ill A sucll that
X
~
pAq (completely isometrically).
Remark. Roger Smith observed (unpublished) that, if X is finite-dimensional, we can choose A finite-dimensional too. Remark. Note that the row (resp. column) space R (resp. C) is clearly 1injective. This corresponds to A = B(£2) with p = ell and q = I (resp. p = I and q = ell) in Theorem 24.2. By [PI3], a reflexive operator space can be A-injective for some A only if the underlying Banach space is isomorphic to a Hilbert space. Theorem 24.2 reduces the classification of I-injective operator spaces to that of I-injective C*-algebras. For dual spaces, the analogous result is as follows.
Theorem 24.3. ([EOR]) Let X be an operator space tllat is dual as a Banach space. Tllen X is l-injective iff there is an injective von Neumanll algebra Af and a projection p ill M such that
x
~
pM(l - p) (completely isometrically).
358
Introduction to Operator Space Theory
!l.foreover, when this holds we can make sure that the preceding isomorphism is also a weak-* isomorphism, in SUell a way that X* = pM*(1 - p) appears as the operator space predual of X. The preceding result is closely connected to the theory of "triple operator systems," for which we refer the reader in particular to [NRl, NR2]. The interest of the preceding theorem is that injective von Neumann algebras seem much better understood ([Col]) than injective C*-algebras. In any case, the preceding two characterizations are valid only for A = 1, and the classification of A-injective operator spaces for A > 1 seems to be much more delicate, just as it is for Banach spaces. For the latter spaces, the theory of the so-called .cp-spaces, especially for p = 1 and p = 00, plays a vital role; it is thus natural to develop the analogous theory for operator spaces. This was recently started in the papers [ERI2, JNRX, JR, JOR]. In another direction, it is natural to investigate injectivity in the separable context, by restricting the extension property discussed at the beginning of this chapter to separable spaces Y and X. A (separable) Banach space X with this property is called sepambly injective. It is a classical result proved in 1941 by Sobczyk that X = Co is separably injective, and the problem whether this is the only (up to isomorphism) separable infinite-dimensional Banach space with this property remained open for a long time, until M. Zippin finally proved in the late 1970s that it is indeed so (see the references in [OR]). Of course, the space IC of all compact operators on £2 is the obvious operator space analog of Co. Another possible analog is the direct sum in the co-sense of the spaces {Mn In:::: I}, which we will denote by K(o). Thus, one would expect these to satisfy the analogous "separable injectivity" in the c.b. context. Unfortunately, however, Kirchberg [Ki2] showed that, in either case, it is not so: Thus, there is a separable operator space Y and a c.b. map u: S --+ IC, defined on a subspace S c Y that does not admit any c.b. extension. Nevertheless, Haskell Rosenthal [R] discovered that the separable (c.b.) extension property holds if the space IC is replaced by either
( EB
LRn)
n>l -
Co
Moreover, he proved that, if Y is restricted to be locally reflexive (in addition to being separable), then 1C(0) satisfies a closely related complementation property, which he called the CSCP: A separable locally reflexive operator space X has the CSCP if, whenever X is completely isomorphic ally embedded in a locally reflexive, separable superspace Y, there is a c.b. projection P: Y --+ X (this corresponds to the extension property when u: S --+ X is a complete isomorphism onto X). Subsequently, with several co-authors ([OR,
24. Injective and Projective Operator Spaces
359
AR]) he proved that K itself has this property. Ozawa also found another proof of this.
Remark. It still remains an open problem (even if Y is locally reflexive) whether, when S C Y with Y separable, any c.b. map u: S --+ K admits a bounded extension to the whole of Y. Actually, it is an old open problem whether, for any embedding K C A of K as an ideal in a separable C* -algebra A, there is a bounded projection P: A --+ K. See Proposition 18.14 for more on this. \Ve note here in passing that Ozawa [Oz3] exhibited an example of a C*algebra A C B(H) and a bounded linear map u: A --+ B(£2) that does not admit any bounded extension to the whole of B(H).
Remark. We should mention also here the related paper [Bio], where it is proved that if A is a I-injective operator system on a separable Hilbert space, and if P is a c. b. projection on A, then either the range of P or that of 1 - P is completely isomorphic to A. We now turn to projective spaces, or equivalently, to spaces satisfying a lifting property. For instance, in the Banach setting, the basic example is the space £1: If X = £1, then any contractive map u: X --+ Y / S into a quotient X --+ Y with lIuli < 1 +€. (In general, space admits, for any € > 0, a lifting we cannot take c = 0.) Contrary to what one would expect at first glance, the analog of this phenomenon for operator spaces is not the space X = SI of all trace class operators, but instead the direct sum £1 ( {Sf I n ~ I}) of the family of all the finite-dimensional versions of S1' 1'Iore generally:
u:
Definition 24.4. Let A ~ 1. An operator space X is called A-projective if, for any € > 0, any c.b. map u: X --+ Y/S into a quotient operator space (here Y is any operator space and S C Y any closed subspace) admits a lifting u: X --+ Y with lIulicb :::; (A + €)llullcb' Examples. (i) For any (finite) integer n, Sf is I-projective. Indeed, by (2.3.2) and (2.4.1)', CB(Sf, Y/S) = Mn(Y/S) = Mn(Y)/Mn(S) (isometrically) and lHn(Y) = CB(Sf, Y); hence any 11 in the open unit ball of CB(Si1 , Y/S) admits a lifting in the open unit ball of CB(Sf, Y), which means that Sf is I-projective. (ii) More generally, if p,q are (orthogonal) projections in M n , then X = I x E Sf} is also 1-projective (since x --+ pxq is a completely contractive projection from Sf onto X). {pxq
(iii) Any direct sum in the £1-sense of a family {Xi liE I} of A-projective spaces is again A-projective. This is easy to see using the defining property of £l-direct sums (see §2.6): Indeed we have isometrically
CB(il({Xi liE I}), Y)
=
ioo{CB(Xi' Y) liE I},
Introduction to Operator Space Theory
360
from which our assertion follows. (iv) In particular, let {ni liE I} be a family of integers with ni ~ 1. Let Pi, qi be projections in Mn;. Let Ti = {PiXqi I x E Sr;}. Then the space £1 ({Ti liE I}) is I-projective. Conversely, any finite-dimensional I-projective space is of this form (this can be deduced from the remark after Theorem 24.2). The projective counterpart to Proposition 24.1 is the following simple observation ([B2J). Proposition 24.5. Let A ~ 1. The follow.ing properties of an operator space X are equivalent.
(i) X is A-projective. (ii) For any c > 0, there is a space Z of the form Z = £1 ({ Sri liE I}) as above such that the identity on X factorizes through Z as follows:
with
IIvllcbllwllcb < A + c.
Proof. (i) ~ (ii): By (2.12.2), any space X is completely isometric to a quotient Z/S for a suitable Z = £1({Sr i liE I}). Let u: Z --+ X be the quotient map. Fix c > o. If X is assumed A-projective, there is a lifting u: X --+ Z with lIuli cb < A + c. Then Ix = uu provides the factorization in (ii). (ii) ~ (i): Let Z be as in (ii). Since we already know that Z is I-projective, it is easy to deduce from (ii) that X is A-projective. • Remark. Note that (ii) above implies that X is completely isomorphic to a completely complemented subspace of Z. Since Z* = EBl\Jni is clearly injective, we immediately deduce: Corollary 24.6. 1£ X is A-projective, then X* is A-injf!ctive. I
While there are rather few projective Banach spaces, many more spaces satisfy the "local" version of projectivity (or, equivalently, a local form of the lifting property). The resulting class of Banach spaces is the class of £1 spaces (see [LiRJ) , which can be defined in many equivalent ways. One of these is X is £1 iff X** is isomorphic to a complemented subspace of an L 1-space. In sharp contrast, the operator space versions of the various definitions of £1-spaces (or, more generally, £p-spaces) lead to possibly distinct classes of operator spaces; see [ER12J. This difficulty is of course related to the lack of local reflexivity in general.
24. Injective and Projective Operator Spfl.ces
361
One of the possible variants is studied in [KyR] under the name of "operator local lifting property," but since we have already used this terminology for a different notion in Chapter 16; so we will change it: An operator space will be called A-locally projective if, for any map u: X --+ Y/ S, any c > 0, and any finite-dimensional su bspace E eX, the restriction of u to E admits a lifting E --+ Y with Ilulicb :::; (A + c)llullcb' It is proved in [KyR] that X has the A-LLP iff X* is A-injective. :More recently, in [EOR] the authors prove that this happens for A = 1 iff there is an injective von Neumann algebra R and a (self-adjoint) projection p in R such that
u:
X*
~
(1 - p)Rp
(completely isometrically).
It follows that X is I-locally projective iff there is a net of finite rank maps of the form X~S~i~X with identity on X.
lIo,;llcb, IIb;llcb:::; 1 that tend pointwise to the
In another direction, the results of [EOR] provide an extension to operator spaces of the classical work of Choi and Effros and Connes (see [CEI, CE2]) on nuclear C* -algebras. Recall that an operator space X is A-nuclear if there is a net of maps of the form X ~lIlni ~X with lIa;llcbllb; Ilcb :::; A that tends pointwise to the identity on X. l\Joreover, (see Theorem 11.6 above) a C*algebra A is I-nuclear iff A ** is injective (equivalently, is a I-injective operator space). The o.s. version of this result proved in [EOR] now reads like this:
Theorem 24.1. An operator spa.ce X is l-nuclear iff X is l-locally reflexive and l-WEP. Recall that X is 1-WEP if the canonical inclusion X pletely contractively through B(H).
--+
X** factors com-
PART III OPERATOR SPACES AND NON-SELF-ADJOINT OPERATOR ALGEBRAS
Chapter 25. Maximal Tensor Products and Free Products of Operator Algebras In the category of algebras (resp. unital algebras), the natural morphisms are of course algebra-homomorphisms (resp. unital ones). In this chapter, we work in the category of operator algebras (Le., closed subalgebras of B(H) for some H), and the natural morphisms there are the completely contractive homomorphisms. Actually, for the most part, we will work with unital operator algebras, and the morphisms will then be understood as unital completely contractive homomorphisms. When the context is sufficiently clear, we will use the term "morphism" and the reader will be supposed to know which category we are working in. We start by recalling the definition of the free product of a.lgebras: Let (Ai)iEI be a family of algebras (resp. unital algebras). We will denote by A (resp. A) their free product in the category of algebras (unital algebras). This object is characterized as the unique algebra (resp. unital algebra) A containing each Ai as a (resp. unital) subalgebra and such that, if we are given another object Band morphisms 'Pi: Ai -+ B (i E I), there is a unique morphism 'P: A -+ B such that 'PIA; = 'Pi for all i. If we now assume that (Ai)iEI is a family of operator algebras (resp. unital ones), then we can equip A (resp. A) with a (resp. unital) operator algebra structure in the following way. Let F be either A or A. Let C be the collection of all morphisms 1£: F-+ B(Hu) such that lIul A i Ilcb ::; 1 for all i in I. Let j: F -+ E9UEC B(Hu) be the embedding defined by j(x) = E9uEC u(x) for all x in F. Clearly j is a morphism and (by standard algebraic facts) it is injective. This allows us to equip F with the noncomplete operator algebra structure associated to j, and, after completion, we obtain an operator algebra (resp. a unital one), admitting F as a dense subalgebra. We will denote by *iEI Ai (resp. *iEI Ai) the resulting (resp. unital) operator algebra, which we call the free product of the family of (resp. unital) operator algebras (Ai)iEI. Let A be any of these two free products. Let fYi: Ai -+ A be the natural embedding. Then, for any family of morphisms Ui: Ai -+ B(H), there is a unique morphism u: A -+ B(H) such that UfYi = Ui for all i. We now turn to the maximal tensor product. This notion originaUy was introduced for C* -algebras. We already discussed this in detail in Chapter 11. In the more general (unital) operator algebra case, it was first considered in [PaP]. For simplicity we will restrict ourselves to a pair of (resp. unital) operator algebras AI, A 2. For any pair 'P = ('PI, 'P2) of morphisms 'Pi: Ai -+ B(H",) with values in a common space B(H",) we denote by 'PI . 'P2 the linear mapping from Al @ A2 to B(H",) that takes al @ a2 to 'PI(at}'P2(a2). If we assume, moreover, that 'PI and 'P2 have commuting ranges (Le., 'PI(at}'P2(a2) = ('P2(a2)'PI(al) for all al in At, a2 in A2)' then 'PI . 'P2 is a morphism from Al @ A2 into B(H",). Conversely, in the unital case, it is easy to see that any
Introduction to Operator Space Theory
366
morphism 1/;: Al ® A2 -> B(H) must be of the form 1/; = 1/;1 . 1/;2 for some pair (1/;1, 1/;2) of morphisms with commuting ranges. For any = L~ ® in Al ® A2 we define
x
Ilxll max
= sup{II'PI' 'P2(x)IIB(H.,)} = sup
{III:;
at aT
'Pl(aI)'P2(a;)IIB(H.,J,
(25.1) where the supremum runs over all pairs 'P = ('PI, 'P2) of (resp. unital) completely contractive morphisms 'Pi: Ai -> B(H",) with commuting ranges. We denote by Al ®max A2 the completion of Al ® A2 equipped with the norm II Ilmax. Clearly this is an operator algebra (resp. a unital one).
Remark 25.1. (i) It is easy to see that if 'Pi: A; -> Bi are morphisms, then 'PI ® 'P2 extends to a (contractive) morphism from Al ®max A2 to Bl ®max B 2. (ii) Since the morphisms are different in each of the three categories of operator algebras, unital operator algebras, and unital C* -algebras, the definition of the maximal tensor product leads to three different notions. Fortunately, however, at least two of these notions "match" each other; that is, given two unital C* -algebras, their maximal tensor products as unital operator algebras and as C* -algebras coincide. Thus the various notions are simply extensions of each other, when a comparison makes sense. Indeed, if AI, A2 are unital C* -algebras, then any unital completely contractive morphism (Ji: Ai -> B(Hi) is "automatically" a *--homomorphism. Thus the two associated "max-norms" are equal.
Lemma 25.2. In the unital case, the mappings
from Al to Al ®max A2 and A2 to Al ®max A2 extend to a morphism
which is a complete metric surjection, that is, it induces a completely isometric isomorphism between the unital operator algebras Al *~b/ ker(q) and Al ®max A 2. Moreover, let :F denote the algebraic free product of Al and A 2. Then the restriction of q to:F defines a complete isometry between :F/ ker(q) n :F and Al ® A2 (where:F C Al *- A2 and Al ® A2 C Al ®max A2 are equipped with the induced operator space structures). Proof. This is essentially routine. By the universal property of the free product, q is uniquely defined and is completely contractive. Let Q = Al *A2/ ker(q) and let v: Q -> Al ®max A2 be the associated injective complete
25. Maximal Tensor Products and Free Products of Operator Algebras 367
contraction. Obviously Al @ A2 is included is the range of v. Therefore, 'I/J = viA\~M2 is a morphism into Q. Let O'i: Ai ----> Al * A2 (i = 1,2) be the inclusion into the free product. It is easy to check that we have
By definition of the maximal tensor product, since IlqO'illcb = 1, this implies that '¢ extends to a complete contraction on Al @max A 2 • In other words, v is onto and completely isometric. • We will see that, when one of the two algebras is the universal operator algebra OAu(E) associated to an operator space E, certain interesting identities appear, for instance, the following one. Lemma 25.3. Let E be an arbitrary operator space. Consider the linear mapping T: A@E@A ----+ OA,,(E) * A that takes a ® e@b (a, bE A,e E E) to the product a . e . b, where E is identified with a subspace of 0 Au (E). Then T extends to a completely isometric embedding from A @h E @h A into OAu(E) * A. Proof. Let x = E ai @ei @ bi be an element in A @E @ A. With the notation of Chapter 5, we have
where the supremum runs over aU possible complete contractions
(i
= 1,2) and 0':
E
---->
B(H)
into the same B(H). By the factorization of c.b. maps (cf. Theorem 1.6) we can assume that CPiO = V; '71"?(') Wi , where 1I"i: A ----> B(HiJ is a representation restricted to A and Vi, Wi are contractions (i = 1,2). Replacing 0'(.) by WI 0'(.) V2 and deleting VI and W 2 we find (25.2) where
1I"J, 11"2
are representations restricted to A and 1I001Icb an d
~
1. Let
~() _ (00 0 O'(e») . O'e-
368
Introduction to Operator Space Theory
Denote by Pi: HI EB H2 -+ Hi the projection onto the i-th coordinate. A simple calculation shows that
and hence we deduce from (25.2) that
where the supremum runs over all 11": A -+ B(H) completely contractive morphisms and all E -+ B(H) with Iialicb = 1. Since any such extends to a completely contractive morphism on OAu(E), we conclude that
a:
a
.
This shows that T is isometric. We leave it to the reader to complete the ~~
By a simple modification, we can prove: Lemma 25.4. Let Ell ... , E n - I be arbitrary operator spaces (n let A be a unital operator algebra. Consider the linear mapping
~
2), and
whicll takes al@el@a2@e2@" '@a n to the product aIeIa2e2 ... an in the free product. Then T is a complete isometry from A @h E 1 @h A @h E 2 @ .•. @h A into the free product OA u (E 1 ) * ... * OA u (En - 1 ) * A. Proof. We only give a hint: Given morphisms 1I"i: A -+ B(Hi) (i :::; n) and complete contractions O"i: Ei -+ B(Hi+t. Hi) we introduce
o 11"
=
(11"1
o The rest is as before.
0) 1I"n
and
a= 0
10"1
0
O"n-1
o
•
Lemma 25.5. Let E, A be respectively an operator space and a unital operator algebra. Let:F be the algebraic free product of 0 Au (E) and A. For any
25. Maximal Tensor Products and Free Products of Operator Algebras 369
z in
Tl1en any element x in F can be uniquely written as a finite sum where eac11 Xj is F j . ~Moreover, we have
Xo
+ Xl +
... + XN,
(25.3)
Proof. Clearly any element X in F can be written as above and (25.3) is obvious. The unicity follows from (25.3). • Consider an arbitrary operator space E and a unital operator algebra A. We will now introduce two (a priori distinct) o.s.s.'s on E 0 A, and we will show that they actually coincide. First, we note that E0A can be viewed as a linear subspace of OA tI (E)0Aj therefore we may equip E 0 A with the operator space structure induced by OAu(E) 0 max A. We denote by ~ the induced norm and by E 06 A the operator space obtained by completing E 0 A equipped with this structure. Let y be an element of E 0 A. We have clearly ~(y)
= sup Iia . 1I'(y) II ,
(25.4)
where the supremum runs over all pairs (a,1I') where a: E ~ B(H) is a complete contraction, 11': A ~ B(H) a morphism, and, moreover, a and 11' have commuting ranges. Indeed, by definition of OAu(E), any complete contraction a: E ~ 1I'(A)' is the restriction of a morphism &: OAu(E) ~
1I'(A)'. Remark. Note tl1at, by Remark 25.1, if A is a unital C* -algebra, the preceding definition ofthe norm ~ coincides with the one given in Theorem 12.1. We now introduce an a priori different structure on E0A, as follows. Note that any y in E0A can be written, for some integer N, as y = L~=l Xjj0aibj, where x E lvfN(E), ai E A, bj E A. (Indeed, since A is assumed unital, this is clear). We define (25.5)
Introduction to Operator Space Theory
370
where the infimum runs over all possible N and all possible representations of y as above. Let q: A ® E ® A -+ E ® A be the linear mapping that takes a ® e ® b to e ® abo It is easy to check that
Note in particular that 8 is a norm. We will denote by E ®/j A the completion of E ® A with respect to this norm 8. More generally, for any n and any y in AJn(E ® A), we define
After completion, we obtain (by Ruan's Theorem) an operator space structure on E ®/j A, and, from now on, we will consider E ®/j A as an operator space. Note that q extends to a complete contraction from A ®h E ®h A to E ®/j A (which we still denote abusively by q), which is a complete metric surjection from A ®h E ®h A onto E ®/j A. In other words, we have A ®h E ®h AI ker(q)
~
E ®/j A
(completely isometrically).
We will now essentially repeat Theorem 12.1, but in a more general setting and with a different proof.
Theorem 25.6. Let E, A be respectively an operator space and a unital operator algebra. Then for any y in E ® A we have ~(y)
= 8(y).
More precisely, we have E
®~
A
~
E ®/j A
(completely isometrically).
Proof. Let y E E ® A. First recall that ~(y) = IlylloAu(El*A. Let 1jJ: OAu(E) * A -+ OAu(E) ®max A be the complete surjection appearing in Lemma 25.2. We will use the notation in Lemma 25.5. By Lemma 25.2, / we have ~(y) < 1 iff there is an element x in :F such that IIxlloA,,(El*A < 1 and 1jJ(x) = y. We claim that x can be replaced by an element Xl in:FI = span(A· E· A). Indeed, we have (cf. Lemma 25.5) X = Xo + Xl + X2 + ... + XN and O'z(x) = Xo + ZXt + z2 X2 + ... + zN XN • Moreover, it is easy to check that
Therefore we have 1jJO'z(x)
= zy for all z, which implies
25. Maximal Tensor Products and Free Products of Operator Algebras 371 Moreover, denoting by m the normalized Lebesgue measure on the unit circle, we have by (25.3) Xl
=
J
az(x)zdm(z),
which ensures (by Jensen's inequality) that
x
Finally, by Lemma 25.3, Xl can be identified with an element in the unit ball of A 0h E 0h A. Thus, we conclude that fl(y) < 1 iff there is an element x in A 0 E 0 A such that q(x) = y and IlxIIA0"E0"A < 1. This proves that fl(y) = c5(y), so that E 0L:!. A ~ E 0 0 A isometrically. It is easy to modify the argument to prove" that this is a complete isometry. We leave the details to the reader. • Proposition 25.7. Consider an operator space F and a unital operator algebra A c B(1t). Let u: F ---> A be a finite rank linear map, and let U E F* 0 A denote tile associated tensor. Til en c5(u) < 1 iff u admits, for some N, a
factorization of the form F~Afw-!!.-4A wi tIl Iialicb < 1 and (3 of tIle form (3(eij) = aibj wi tIl ai,bj E A sucil tllat (25.6)
Moreover, if F is a "minimal" operator space, tilat is, if F = min(F) (completely isometrically), then c5(u) < 1 iffu admits, for some N, a factorization of the form F~e[!~A witiliialicb < 1 and (3 oftlle form (3(ei)'= aibi witil ai, bi E A SUell tilat (25.6) 1101ds. Proof. The first part is essentially obvious by (25.5) with E = F*. Now, if F = min(F), then any element a in the unit ball of MN(max(F*)) admits a special factorization as described in Theorem 3.1. Using this, one easily completes the proof. • Definition 25.8. Let F, A be an operator space lllld a unital operator algebra. A linear map u: F ---> A will be called c5-boundedly approximable if there is a constant C and a net Ui: F ----> A of finite rank maps tending poillt~dse to u and sucil tilat tile associated elements Ui in F* 0 A satisfy SUPiEI c5(u·d ::; C. We will denote by D( u) tile smallest constant C such tllat tllis holds. Theorem 25.9. Tile following properties of a unital operator algebra A are equivalent: (i) For any unital operator algebra B, B 0 m in A = B ®max A (isomorph ically). (ii) The identity of A is c5-boundedly approximable.
Introduction to Operator Space Theory
372
Moreover, if A is a C* -algebra, this holds iff A is nuclear and, in the C* -case, the identity in (i) is automatically completely isometric. Proof. We first claim that, for any x in B @ A and any u: A rank with associated tensor E A * @ A, we have
-4
u
A of finite
(25.7) Indeed, let E = A* and let v: E -4 B be the finite rank map associated to x. Assume Ilxlimin = IIvll cb ::::: 1. Then v extends to a morphism v: OAu(E) -4 B, so that, by Remark 25.1.(i) we have lI(v@IA)[ulll max
:::::
lIulloA,,(E)0maxA = ~(u) = 8(u).
Since (v@IA)(u) = (v@IA)(u) = (IB@U)(X), we obtain (25.7) by homogeneity. Assume (ii) and let Ui be as in Definition 25.8, with U = IA' Note that, for any x in B @ A, (IB @ Ui)(X) -4 x in the largest Banach tensor norm (say); hence it is in the norm of B @max A. Then we have for any x in B @ A
which yields in the limit IIxll max ::::: Cllxllmin.
(25.8)
This shows that (ii) =? (i). Conversely, assume (i). By an easy direct sum argument, there is a constant C such that for all B and all x in B @ A we have (25.8). Consider then an arbitrary finite-dimensional subspace F c A and let iF: F - 4 A denote the inclusion map with associated tensor iF E F* @ A. Note that IliFllmin = IliFlicb = 1; hence (i) implies ~(iF) ::::: C. By Theorem 25.6 and' a simple extension argument, we obtain for any € > 0 a finite rank map u: A -4 A extending iF: F -4 A and such that lIuIIA.0.A < C(l + e). Letting F run over the directed net of finite-dimensional subspaces F c A we obtain a net as in Definition 25.8, whence (ii). Finally, if A is a C*-algebra, then any unital completely contractive homomorphism 1T: A -4 B(H) is a *homomorphism, and if A is nuclear, 1T(A) 'Is also and the min- and max-norm coincide on 1T(A), @ 1T(A). Therefore, for any unital operator algebra Band any morphism a: B - 4 1T(A)', we have for any x in B @A:
lIa . 1T(X) II
::::: II (a @1T)(x)II1T(A)'0max1T(A) = II (a @1T)(x)lImin ::::: Ilxllmin.
Hence we obtain IIxlimax ::; IIxlimin, which shows that (i) holds isometrically. Replacing A by Mn(A) (n 2: 1), we obtain a complete isometry. Conversely,
25. Maximal Tensor Products and Free Prodtlets of Operator Algebras 373 if a C* -algebra A satisfies (i), then a fortiori (i) holds whenever B is a C*algebra, which is the definition of nuclearity (cf. Chapter 11). • Remark 25.10. Let A be a unital operator algebra. The preceding proof shows that the smallest constant C such that, for any unital operator algebra B and any x in B ® A, we have Ilxll max ::; Cllxllmin, is equal to D(IA).
We will denote by D(A) this constant. In sharp contrast with the C*case, the constant D(A) can obviously be both finite and> 1 in certain cases (for instance, when A is finite-dimensional; see Remark 25.14). If A is a C*algebra, D(A) can only take the value 1 if it is finite. The converse is true (in a very strong sense), as shown by the next result. Theorem 25.11. Let A c B(H) be a uuital closed subalgebra. Theu D(A) = 1 implies that A is self-adjoint. Equivalently, D(A) = 1 iff A is a nuclear C*-subalgebra of B(H).
We will use the following elementary fact: Lemma 25.12. Fix el > 0 and e2 > o. Let ai, Pj E B(H) be SUdl that 1 = L~ aiPi, L~ aiai ::; 1 + el, L pj Pj ::; 1 + e2· Then we have
Proof. For any h in H with Ilhll = 1, we have 1 = (h,h) = L(p;h,aih); hence
2: Ilpi h -
ai hl1 2
=
(2: P; Pih, h) + (2: aia: hlh) -
::; el
2
+ e2·
Proof of Theorem 25.11. We have seen in Theorem 25.9 that D(A) = 1 if A is a nuclear C*-algebra. Conversely, assume D(A) = 1. Using the same argument and the same notation as in the proof of (i) ::::} (ii) in Theorem 25.9, we find that, for any finite-dimensional subspace F c A and any e > 0, there are, for some n, an element (eij) in A1n(F*) and a;,bj in A such that
and
n
'
X
=
L
eij(x)aibj.
i,j=l
[Indeed, this simply makes explicit the fact that c5(iF ) < (1
+ c)1/2.J
Introduction to Operator Space Theory
374
Let us now assume / E F, so that we have
where f3i
= 'L j eij(l)bj , "(j = 'Li eij(l)ai.
Note that 11'Lf3if3ill l / 2 ::; lI(eij(I)ILu"II'Lbjbjlll/2 < (1 +e)I/2, and similarly "(i'Y; < 1 + e. By Lemma 25.12, it follows that
II'L
I
Z)f3i - a;)*(f3i - an ::; e
and
Z)bj - "(j)*(b j
"(j) ::; e.
-
Note that f3i, "(j E A, so that the last two inequalities show that at and bj are "close" to being also in A. More precisely, for any x in F, we have n
x* =
I: eij(x)bja;.
ij=1 Assume Ilxll ::; 1. Then, if we set n
y=
I: eij(X)"{jf3i,
ij=1 we have by (1.12)
IIx* -
yll ::;
III:eij(X)(bj - "(j)a;1I
+ lII:eij(x)"{j(a;
::; 1II:(b; - "(j )(b; - "(j)* 111/2 +
III: aia;
-
f3dll
111/2 (1 + e)I/2
III: "(j"(;11 1/ 2 11I:(a; - f3i)*(a; - f3i)1I 1/
2
(1 + e)I/2
::; 2(e(1 + e»1/2.
Thus we proved dist(x*, A) ::; 2(e(1 +e»1/2, and since e > 0 is arbitrary, we conclude that x* belongs to A. Applying this to Fx = span[/, xl, where x runs over all elements of A, we conclude that A is self-adjoint. •
Remark. See [LeM4] for generalizations of the preceding statement. Remark 25.13. Actually, we only use the fact that (eij) defines a map e: F -+ Mn with ordinary norm::; 1. In other words, we only use max(F) instead of F or, equivalently, min(F*) instead of F*. Moreover, we only consider two-dimensional subs paces F. Thus the conclusion of Theorem 25.11 holds if B ®min A = B ®max A (isometrically) for any B of the form B = OAu(G), where G is any two-dimensional subspace of loo.
25. l\Jaximal Tensor Products and Free Products of Operator Algebras 375 Remark 25.14. It is easy to show (using an Auerbach basis; cf. [LiT, p. 16]) that, for any n-dimensional operator space E, any unital operator algebra A, and any x in E ® A, we have (25.9)
8(x) ::; nllxllmill' Therefore, if dim(A)
= n,
then we have D(A) ::; n.
Remark 25.15. The conclusion of the preceding statement can fail if D(A) > 1. Indeed, consider the two-dimensional algebra A C B(H) (dimH = 2) formed by all matrices of the form
(~
!) (a, b
E
C). Note that the ma-
. ( 00 01).IS a nonzero eI '111 A wIt . I1 zero square. S'111ce A'IS fi lllte. tnx ement dimensional, we have trivially D(A) < 00; but on the other hand, A is not isomorphic to a C* -algebra. (Indeed, if it were, since A is commutative, it would have to be isomorphic to l;}, but this is absurd since C~) does not contain any nonzero element with zero square.) In the remainder of this chapter, we wish to discuss several examples for which the following well-known fact will be useful. We first recall some notation. Let D be the open unit disc in C with boundary 8D. For any integer k ~ 1, we denote by A(Dk) the closure in C((8D)k) of the algebra of aU (analytic) polynomials on Ck equipped with the induced operator algebra structure. Thus A(Dk) is equipped with its minimal o.s. Note the completely isometric identities A(Dk) = A(D)®mill" '®miIlA(D) and C(8D)k) = C(8D)®mill" ·®millC(8D).
Proposition 25.16. Let u: A(D) ® ... ® A(D) ----> B(H) be a unitalllOmomorphism. Let Tl = u(z ® 1 ® ... ® 1), T2 = u(1 ® z ® 1 ® ... ® 1), ... , and Tk = u(1 ® ... ® 1 ® z). The following assertions are equivalent.
(i) u extends completely contractively to A(D) ®mill ... ®rnill A(D). (ii) There is a Hilbert space H with H cHand a *-llOmomorpllism 7r: C(8D) ®rnill ... ®mill C(8D) ----> B(H) such that 'if
f
E
A(D) ® ... ® A(D)
u(f)
=
PH7r(f)IH.
(iii) There is a Hilbert space H witll H cHand a k-tuple of mutually commuting unitaries (Ul , ••• , Uk) 011 H SUell that for allY polynomial P we have
Proof. (i)
=}
(ii) is an immediate consequence of Corollary 1.8.
376
Introduction to Operator Space Theory
(ii) =} (iii) is immediate: We simply set Uk = 7I"(Zk) , where Zk denotes the k-th coordinate function on (8D)k viewed as an element of C(8D) ® ... ® C(8D).
Finally, assume (iii). The mapping P --+ P(U1 , ••• , Uk) obviously extends to a *-homomorphism 71" on C(8D)®max" '®maxC(8D) = C(OD)®min" '®min C(8D). Restricting to the subspace A(D) ®min ... ®min A(D), we obtain (i) .
•
Example 25.17. We have completely isometrically A(D) ®min A(D)
= A(D) ®max A(D).
(25.10)
As observed in [PaP], this is essentially a reformulation of a famous dilation theorem due to Ando (see e.g.[P1OD: Any pair of commuting contractions T1 , T2 in B(H) admits a unitary dilation, that is, there is jj with H c jj and commuting unit aries U1 , U2 on jj such that, for any polynomial P, we have P(Tl' T 2 ) = PHP(Ul, U2 )IH'
Proof of (25.10). It suffices to show that, for any pair of mutually commuting completely contractive morphism 7I"j: A(D) --+ B(H), (j = 1,2), the morphism 71"1 • 71"2: A(D) ® A(D) --+ B(H) extends completely contractively to A(D) ®min A(D). Taking into account the preceding proposition, this follows • from Ando's dilation theorem. By a well-known example of Varopoulos (see e.g.[P1OD Ando's dilation theorem does not extend to three mutually commuting contractions. The following related example due to S. Parrott [Par2] is very important.
Example 25.18. ([ParD There is a contractive homomorphism u: A(D 3 ) --+ B(H) that is not completely contractive. Equivalently, there is a triple (Tl' T 2 , T 3 ) of commuting contractions such that, for any polynomial P in three variables, we have (25.11)
but for which the morphism
is not completely contractive. The operators Tb T2, T3 will be of the form
T. = 1
(0 00) aj
25. Maximal Tensor Products and Free Products of Operator Algebras 377 acting on H = K tBK as in Exercise 25.1, and aj, (j = 1,2,3) will be suitably chosen contractions on K. Note that T j Tk = 0 for all j, k, so (TI' T 2 , T 3 ) mutually commute. Let P( ZI , Z2, Z3) be a polynomial on D3. Let P = Po +PI +... be its decomposition into a sum of homogeneous polynomials. Note that Pd(TI , T 2 , T 3 ) = 0 for all d ~ 2, so that we have
where Po(z)
= Ao and PI(z) = L~AjZj. Fix (ZI,Z2,Z3) in D3. Let T
(L~ IAjl) -1 L~ AjTj .
=
By von Neumann's inequality (see (8.11) above) we
have
II Aol + (L AjZj) . Til =IIP(zIT, Z2 T , z3T)II :::; sup IP(ZIZ, Z2Z,
z3z)I :::; sup IPI.
zED
Hence, choosing Zj E aD such that AjZj
D3
= IAjl, we find
which establishes (25.11). We now claim that, for a suitable choice of unitary operators aj' (j = 1,2,3), the morphism u: A(D 3) -+ B(H) taking P to P(TI' T2, T 3) is not completely contractive on A(D 3) ~ A(D) ®min A(D) ®min A(D). Indeed, by Proposition 25.16, if Ilulicb = 1, then there is jj with H c jj and a triple (Ut, U2, U3) of mutually commuting unitaries on jj dilating (T1' T2, T3). In particular we have Tj = PHUjI H. By Exercise 25.1, if aj itself is unitary, we must have for any j, k 'V h
E
K
Therefore, if Ut, U2, U3 commute, then U3- IU I and U:;1U2 commute, and the preceding identity shows that a3" 1a1 and a3" 1a2 must commute, but it is very easy to produce examples of unitaries (at, a2, a3) for which this fails! Just take, for instance, a3 = I and choose for aI, a2 any pair of 2 x 2 unitary matrices that do not commute. Remark. The preceding example shows that three commuting contractions may fail to dilate to three commuting unitaries. Nevertheless, it is proved
Introduction to Operator Space Theory
378
in [GaR] that any k-tuple of contractions (Tj commute (Le., such that
),
(j = 1, ... , k) that cyclically
TIT2 ... Tk = T2 ... TkTl = T3 T4 ... TkTlT2 = ... = TkTl ... Tk-l) admits unitary dilations that cyclically commute. See toP] for a discussion of a tensor product, analogous to Q9J.t' but associated to cyclic commutation. The following variant of Parrott's example was kindly pointed out to us by C. Foias.
Example 25.19. There is an example offour contractions (Tj such that TiTj = TjTi 1 '$ i '$ 2, 3 '$ j '$ 4 but which cannot be dilated to four unitaries (Uj UiUj
= UjUi
1 '$ i '$ 2,
),
),
(j = 1, ... ,4)
(j = 1, ... ,4) satisfying
3 '$ j '$ 4.
Indeed, consider again H = [( EB [( and Tj E B(H) of the form T. J
=
(0 00). aj
Let (Uj ) be unit aries on Ii :J H such that Tj = PHUjI H ' If each of Ul, U2 commutes with each of U3, U4, then a fortiori U1 l U2 must commute with UilU4' By Exercise 25.1, this forces a 11 a2 to commute with a3" l a4' but here again it is very easy to produce unitary matrices (aj) for which this fails! Just take al = a3 = I and a2, a4 noncommuting unitaries. • Consider two operator spaces E, F. Recall (see Remark 8.12) that we have completely isometric embeddings
OAu(E) C
C~(E)
and
OAu(F) C
C~(F).
Taking the tensor product of these (unital) morphisms we obtain a (unital) morphism
Q9max
OAu(F)
~ C~(E) Q9max C~(F).
(25.12)
such that
1I
(25.13)
In the particular case E = F == C, we have OAu(E) = OAu(F) = A(D) (by Theorem 8.11), and (25.10) impl~ a fortiori that
In.
25. Maximal Tensor Products and Free Products oE Operator Algebras 379 Proposition 25.20. The natural embedding E®F induces a completely isometric embedding
c OAv(E)®rnaxOAu(F)
E ®,.. F c OAu(E) ®rnax OAv(F). More genera.lly, Ear any N-tuple oE operator spaces E I , ... ,EN we have a completely isometric embedding
By the preceding proposition, this implies: Proposition 25.21. IE E, F are maximal operator spaces, tlWIl the Ilorms induced on E ® F by C~(E) ®rnax C~(F) and OAu(F) ®max OAtt(F) are equivalent. Proof. Consider x in E ® F. Let u: E* ---> F be the associated map. Note that E*, F* are minimal operator spaces by Exercise 3.2. By Theorem 19.11, we have
Since it is not hard to check (see Exercise 5.7) to check that
Ilxli lL =
'Y2(U),
we obtain
The converse follows from (25.13). Let 1,1' be two arbitrary sets. Let E = max(£I(I» and F Then, by Theorem 8.8, we have a (unital) morphism
• = max(£I(I'».
such that
Ilwllcb = 1. Let
ei
be the canonical basis of £1(1). Note that, for any i in I, the inclusion
380
Introduction to Operator Space Theory
takes ei to a unitary Ui in C*(Fr), so that, for any i' in I', \II(e;®ei') = Ui®Ui , is also unitary in C*(Fr) ®max C*(Fr'). Example 25.22. Consider the case I = I' = {1,2}, so that E = F = max(RD. Then \II is not completely isometric. More precisely, the restriction of \II to the five-dimensional space S = 1 ® 1 + E ® 1 + 1 ® F is not completely isometric.
Proof. Let (Tj ), (1 ~ j ~ 4) be 4 contractions as in Example 25.19. Then let CPI: OAu(R~) --+ B(H) (resp. CP2: OAu(R~) --+ B(H)) be the unique completely contractive unital morphisms such that
(resp. CP2(el) = T3 , CP2(e2) = T4). Our (partial) commutation assumption on (Tj ) ensures that CPt,CP2 have commuting ranges; hence CPl' CP2 extends to a completely contractive morphism on OAu(E) ®max OAu(F). Now, if \II were completely isometric, we would be able to extend cP = CPI . CP2 to a completely contractive unital map cP on C*(F2)® max C*(F2)' By Corollary 1.8, this would imply that there is ii with H c ii and a representation 11': C*(F2) ®max C*(F2) --+ B(ii) such that V x E OAu(E) ® OAu(F)
In particular, setting Uj = 1I'(ej ® 1) and Uj+2 = 11'(1 ® ej), we would obtain four unitaries contradicting Example 25.19. This contradiction shows that \II is not completely isometric. By Proposition 13.6, this implies that the restriction \IIjq,\S) has c.b. norm> 1, or • equivalently, that \illS is not completely isometric. Example 25.23. Let A be the disc algebra A(D), let cP be a finite Blaschke product of degree n, and let I c A be the ideal generated by cP, that is, 1= {cpl I I E A}. By Corollary 6.4, we may consider the quotient All as an operator algebra. Let q: A --+ AI I be the quotient morphism. Let 11': AI I --+ B(H) be any morphism, and consider the composition 11' = 1I'q: A --+ B(H). Since A is generated by thesingle function z, 11' and 11' are entirely determined by the single contraction T = 7r(z) E B(H). Note that cp(T) = 7r(cp) = O. Conversely, any contraction T such that cp(T) = 0 determines uniquely a morphism 11' as above. Moreover, {1I'(AII)}' = {T}'. Let us denote by C(cp) the smallest constant C such that, for any T satisfying cp(T) = 0 and for any polynomial P(z) = E~ akzk with coefficients in {T}', we have (25.14)
25. :Maximal Tensor Products and Free Products of Operator Algebras 381 The following fact is essentially due to Bourgain [Bou4] but was observed by Daher (see [PlO, p. 90]): There is a numerical constant [( such that, for any Blaschke product rp of degree n, we have
C(
s [( Log(n + 1).
(25.15)
Daher observed that Bourgain's proof of a related result in [Bou4] actually establishes (25.15). Let q E A* Q9 AI I be the tensor associated to q. It is not difficult to see that C(
C(
=
c5(q).
Therefore, the second part of Proposition 25.7 implies a striking factorization of q (which is closely related to, but a bit different from, the one appearing in [Bou4]). We now consider the special case when rp = rpn where rpn(z) = zn, and we will show in this case that (25.15) is sharp, that is, there is a constant [(' > 0 such that \In? 1
/(' Log(n + 1) S C(
(25.16)
To verify this we will use Hardy's classical inequality (d., e.g., [Ka, p. 91]) concerning "analytic" measures on the unit circle 'Jl'. We denote by .1I1('Jl') the Banach space of aU complex measures on 'Jl'. Recall that AJ('Jl') = C('Jl')* isometrically. A measure It in .111 ('Jl') will be called analytic if J w j II( dw) = 0 for all j > o. With this terminology, Hardy's inequality (together with the F. and M. Riesz' Theorem) asserts that there is a constant /(1 > 0 such that any analytic measure 11 satisfies (25.17)
where the Fourier coefficients are defined as
Fix an integer n > 1. We set rp( z) = zn, and we consider the operator T E .1Ifn defined by Tel = 0 and Tei = ei-l for i = 1,2, ... , n. Clearly, Tn = 0, IITII = 1 and IITn-lll = 1. Assume that (25.14) holds. We claim that this implies that there is an analytic measure 11 such that Il/llillI(lI') S C( rpn) and "ji(k) = 1 for all k = 1, ... ,n - 1. Using (25.17), we immediately deduce that (1 + ... + lin - 1) ::; K1C(
382
Introduction to Operator Space Theory
Now let f be a function in the linear span in C('Jr) of the functions {zi j E Z, j ~ n - 1}. For any fixed z in 'Jr, we define
fAe) =
2:
I
f(j)z i en- 1- i .
i:$n-l By (25.14) applied with ak
= [(k)T n- 1- k (0 ~ k ~ n - 1) we have
112:;-1 [(k)Tn- 111 ~ C sup II2: [(k)Tn-l-kzkll.
12:;-1 f(k)1 =
n
Izl=1
-
1
0
Now observe that fz is a polynomial such that fz(T) and by von Neumann's inequality we have
Thus we obtain
2:o
l
n
-
1
f(k)1
= L~-1 f(k)T n- 1- k z k ,
~ Csup Ifz(~)I· ~ET
But it is easy to check by translation invariance on the circle that
whence finally
2: f(j)z-i i~n-l
em
By the Hahn-Banach Theorem, there is a linear form IL E C('Jr)* with norm ~ C such that fl(j) = 1 if 0 f j ~ 11, - 1 and fl(j) = 0 if j < O. Identifying IL with a (complex) measure on 'Jr, we obtain the above claim. •
Remark. The proof of the preceding estimate (25.16) establishes a conjecture formulated in the first edition of [PlOj. Moreover, the equality C(cp) = 8((j) together with (25.16) yields a minor improvement over Bourgain's estimate in [Bou4j; namely, if T E Mn is polynomially bounded with constant C, then there is an invertible matrix S E Mn such that IIS-ITSII $ 1 and satisfying IIS- 1 11I1SI1 $ K Log(n + 1)C2 •
25. Maximal Tensor Prodllcts and Free Products of Operator Algebras 383 Exercises Exercise 25.1. Let [( be a Hilbert space and let H = [( EB K. Let T B([( EB [() be the operator defined by the matrix
E
Let U be a unitary dilation of T on a Hilbert space ii with ii :J H. Show that, if a is isometric, then for all x in [( we have necessarily (cf. [SNF, Par2]):
Moreover, if a is unitary, we also have
Exercise 25.2. Let E 1 , E2 be a pair of operator spaces. Consider the symmetrized Haagerup tensor product E 1 @p. E2 defined before Theorem 5.17. Show that we have a canonical completely isometric embedding
Chapter 26. The Blecher-Paulsen Factorization. Infinite Haagerup Tensor Products We begin this chapter by a striking factorization theorem due to Blecher and Paulsen ([BP2]).
Theorem 26.1. Let A c B(H) be a unital closed subalgebra that is generated by a unital subset S of the closed unit ball of A so that, if A denotes the algebra generated by S, then A is dense in A. Let K 2:: 0 be a constant. The following are equivalent: (i) Any unital homomorphism u: A is c.b. and satisfies lIulicb ::; K. (ii)
-+
B(H) such that
sUPaES
Ilu(a)11 ::; 1
For any n and any x in Mn(A) with IlxI1ll1 n(A) < 1, tIl ere is, for some m, a factorization of the form x = aoD!a! ... Dma m , where ao, ... ,am are (possibly rectangular) matrices with scalar entries, SUell that Il7=0 Ilaj II < K and D!, ... ,Dm are diagonal (hence square) matrices with entries in S.
(iii) Same as (ii) but with each Di having one diagonal entry in S and all others equal to the unit of A. Proof. The implication (ii) =} (i) is easy and left as an exercise for the reader. Conversely, assume (i). For any x in Mn(A) let us denote
I!lxlil.
~ in!
{fi IIO;II} ,
where the infimum runs over all possible factorizations of x as in (ii) above. Note that IIxI11l1n(A) ::; IIlxlll n . It is rather easy to verify that the sequence of the norms (III Illn) satisfies Ruan's Axioms (R~) and (R~); hence, after completion, we obtain an operator space E such that E is the completion of A and IIlxlll n = IIxI11l1,,(E) for all x in Mn(A). (See §2.2). The very definition of the norms III Illn shows that the product of A extends to a completely contractive mapping from E ®h E to E and lilA liE = 1I1AIIA = 1, so that,'by Theorem 6.1, E is an operator algebra for this product. Hence there is a completely isometric unital homomorphism j: E -+ B(H). Obviously (using a trivial factorization of the 1 x 1 matrix (x» we have, by definition of III Illn'
'IxE
S
IIxllE = III(x)1I1t ::; 1.
Therefore, our assumption (i) implies /lj: A -+ B(H)lIcb :5 K.
26. The Blecher-Paulsen Factorization
385
In particular, this implies that, for any x in Afn(A), we have Illxlll n
= 1I(I®j)(x)IIM,,(B(H))::; [(llxIIM,,(A),
whence (ii). Finally, the equivalence of (ii) and (iii) is obvious, since each N x N diagonal matrix Di with entries in S can clearly be written as a product of N matrices of the form appearing in (iii). •
Definition. When the properties in Theorem 26.1 hold, we say that S [(completely generates A, and if [( = 1, we simply say that S completely generates A. The simplest example consists in taking S equal to the whole closed unit ball of A. Then we immediately obtain.
Corollary 26.2. Let A c B(H.) be a unital closed subalgebra and let [( :::: 0 be a constant. The following are equivalent:
(i) Any contractive unital homomorphism u: A ---> B(H) is c.b. and satisfies lIull c b ::; [{. (ii) For any n and any x in .lIIn(A) with IlxIIM,,(A) < 1, there is a factorization of tIle form x = QOD I Q I D 2 ... DmQ m , where Qo, QI, ... , Qm are (possibly rectangular) matrices with scalar entries and D I , ... , Dm are diagonal (hence square) matrices with entries in A satisfyiilg
(Note that m and the sizes of the matrices
Qi,
Di are arbitrary.)
Definition 26.3. 'Ve will say that a unital operator algebra A is "fulf' (perhaps "completely full" would be better!) if every contractive unitall1Ol110l110rphism u: A ---> B(H) is completely contractive. In other words, A is full iff it satisfies (i) in Corollary 26.2 with [( = 1, 01', equivalently, if it is completely generated by its unit ball. For example, every unital C* -algebra A satisfies this. In this case, every contractive unital homomorphism u: A ---> B(H) is automatically a *homomorphism (hence a fortiori a complete contraction). Indeed, first observe that an element x in A (or in B(H) is unitary iff :1: is invertible and both x and x-I are contractive. Thus, if we denote by U(A) the set of all unitary elements in A, we see that any unital homomorphism u: A ---> B(H) satisfying sup{llu(x)1I I x E U(A)} ::; 1 must take unitaries to unitaries and hence must be a *-homomorphism. In addition to C*-algebras, the disc algebra A(D) and the bidisc algebra A(D2) are both full. More precisely, A(D) (resp. A(D2» is completely generated by {I, z} (resp. {I, Zl, Z2} ). This follows, as explained in the following,
386
Introduction to Operator Space Theory
from two classical dilation theorems due to Sz.-Nagy and Ando (see [Pal] or [P1OD. However, by Example 25.18, A(Dn) is not full when n 2:: 3. Note that A(Dn) is a minimal operator space; hence, for n = 1 or n = 2, if a unital operator algebra is isometrically isomorphic to A(Dn ), it is automatically completely isometric to it. The classical Sz.-Nagy dilation theorem says that, given any contraction T on a Hilbert space H, there is a larger Hilbert space H containing H and a unitary operator U: H --+ H such that (26.1 ) Let A be the algebra of (analytic) polynomials on the open unit disc Dee, and let u: A --+ B(H) be the homomorphism defined by u(P) = P(T).
Then (26.1) implies (26.2) where 11': A --+ B(H) is defined by 1I'(P) = P(U). Note that (by the spectral functional calculus), since U is unitary (hence a fortiori normal), 11' extends to a C*-representation from C(8D) to B(H), and in particular 111I'1Icb = 1. Let A(D) C C(8D) be the closure of A in C(8D). Then,.(26.2) implies lIuIlCB(A(D),B(ll))
=
1.
Thus, Sz.-Nagy's dilation theorem shows that, in the disc algebra, the set S = {1, z} satisfies (i) in Theorem 26.1 with K = 1. Whence: Corollary 26.4. (fBP2J) Fix n 2:: 1. Let f: D --+ Mn be an analytic matrix valued function stich that sUPzED IIf(z)IIM .. < 1. Assume moreover that all the entries hj (z)\ are polynomials. Then there is, for some integer m, a factorization of tIle form
where aj, bj are scalar (possibly rectangular) matrices such that sUPzED lIaj + zbj II < 1 for all j ::; m. More precisely, we can obtain aj + zbj of the following form:
387
26. The Blecher-Paulsen Factorization
where Qj-J,Qj are scalar matrices SUcll tlJat lIaj-lll, Ilajll < 1 andwhereDj is a diagonal matrix of size (say) N j x N j of the form 1
o
1 Z
o
Z
with 1 appearing Pj times and z appearing qj times (Pj N j ).
~
0, qj
~
0, Pj
+ qj
=
In the case of the bidisc algebra A(D2), let A be the algebra of all polynomials P(ZI,Z2) on D2 and let u: A --+ B(H) be the unital homomorphism that takes P to P(T1 ,T2 ), where TI,T2 E B(H) are two commuting contractions. Then, ~mlo's dilation theorem (see e.g.[PlOJ) asserts that there is a Hilbert space H ::J H and two commuting unit aries U1 , U2 such that for any P we have P(T1 , T2) = PHP(Ul, U2)IH. As above, this shows that, in the bidisc algebra A(D2) C C(8D x 8D), the set S = {1,ZI,Z2} satisfies the property (i) in Theorem 26.1. Whence:
Corollary 26.5. Fix n ~ 1. Let f: D2 --+ !lIn be an analytic matrix valued function, with polynomial entries, such that sUPzI,z2ED IIf(ZI,Z2)1I,'I1" < 1. Then f admits for some Tn ~ 1 a factorization of tIle form
where TIllaj II < 1 and where each D j is a diagonal matrix of size (say) N j x N j with diagonal of the form
8~~2"~,,Z2J P.i
qj
Tj
Remark. If we wish, we can assume in the preceding statements that Nl = N2 = ... = N m = N, so that only Qo and Q m are rectangular matrices (of size n x Nand N x n) and all the other ones are N x N matrices. Remark. The preceding two corollaries are already significant for n = 1, that is, for ordinary complex-valued analytic functions! Problem. Surprisingly, the following seems to be still open: Let K(k) be the supremum of lIulicb over all possible contractive unital morphisms u: A(Dk) --+
388
Introduction to Operator Space Theory
B(H). By the Sz.-Nagy and Ando dilation theorems we have K(l) = 1 and K(2) = 1. By Example 25.18, we know that K(3) > 1 and a fortiori K(k) > 1 for any k 2:: 3. We suspect that Corollary 26.2 might be useful to show that K(3) = 00. However, although it is unlikely to be true, it is unknown whether K(3) is finite! Consider now a unital C* -algebra A and let S be the group formed of all the unit aries in A. Then (see Remark 26.3) any unital homomorphism u: A --+ B(H) such that sup{llu(x)1I1 XES} ~ 1 is a *-homomorphism and therefore has Ilulicb = 1. Whence, by Theorem 26.1: Corollary 26.6. Fix n 2:: 1. Let A be a unital C* -algebra. Consider x E Mn(A) with IlxIIMn(A) < 1. Then x admits a factorization of tIle form
n
where Clj are complex matrices (possibly rectangular) such that IIClj II < 1 and where D j are diagonal matrices with entries in the unitary group of A. Remark 26.7. Let G be a discrete group, and let reG be a subset containing the unit and generating G (so that any element of G is a finite product of elements of r). Let A = C*(G), let A = qG] c C*(G), and let S be the copy of r sitting naturally inside the unitary group of A. Then, it is easy to check that Theorem 26.1 applies in this situation so that, in the preceding statement, if x E Mn(A) and IlxIIM.. (A) < 1, we can find a factorization as above but with the D j having entries in S (Le., in r with the obvious identification).
In Theorem 26.1, we consider a subset S of the unit ball of A. Actually, in operator space theory, it is more natural to consider the unit ball of Mn(A) or of K ®min A. This leads to the following easy generalization of Theorem 26.1. Theorem 26.8. Let A be a unital operator algebra and let A c A be a dense unital subalgebra. We will identify as usual Mn(A) witl] a subset of M n+1(A) and with a subset of K ®min A. Let S be a subset of the unit ball of K ®min A. We assume that
ScK®A and, moreover, that the algebra generated by S contains Un> 1 Jl.,fn (A). Then, for any fixed constant K 2:: 0, the following are equivalent: (i) Any unital homomorphism u: A --+ B(H) such that SUPxES 1Ih:: ® u(x)1I ~ 1 is c.b. and satisfies lIull cb ~ K. (ii) For any n, any x in Mn(A) with II x IlMn(A) < 1 admits (for some m) a factorization of the form x = CloD1Cll ... DmCl m , where ClO, ... , Clm are in K ® 1 with nllClj II < K and where Db ... , Dm are elements of
389
26. The Blecher-Paulsen Factorization K0
m in
A represented by block diagonal matrices of the form
o o
IYN j (j)
with Yk(j) E S for all k and j. Proof. The proof is analogous to that of Theorem 26.1, so we merely out.line it. Assume (i). Note that any x in the algebra generated by S admits a factorization as in (ii). Thus, for any x in Un~11l1n(A), we can define the norm
IIlxlli
~ inf
{fi II<>'II} ,
where the infimum rl1ns over all possible factorizations as in (ii). By Ruan's Theorem and the Blecher-Ruan-Sinclair Theorem, there is a unital homomorphism u: A ---4 B(H) such that h0u defines an isometry from (Un>lllfn(A), III III) into K0 m in B (H). By definition of III III, if we denote by Pn~ K ---4111n the natural projection, for any x in S and for any n, we have III (Pn ® 1)( x) III ::; 1; hence IlPn 0 u(x)IIM,,(B(H» ::; 1. Therefore, IIh 0 u(x) II = limn--->oo IlPn 0 u(x)IIM,,(B(H)) ::; 1. Thus, by (i), this implies that lIullcB(A,B(H))::; K, or, equivalently, for any x in Un~l .l\fn(A) we have
IIlxlll = IIh 0 u(x) II ::; KlixIIK®minA, so that we conclude that (ii) holds. This shows (i)
=}
(ii). The converse is easy and left to the reader.
•
We will now show that the product factorization appearing in the preceding statements can be interpreted in terms of infinite Haagerup tensor products. We first define the latter. Let I be a totally ordered set and let (Ej)jEI be a collection of operator spaces, given together with a family (~j)jEI with ~j E E j and lI~jllEj = 1. For any finite subset J c I we can form the Haagerup tensor product
390
Introduction to Operator Space Theory
Note that this definition depends on the ordering of Jj that is, what we mean here is that, if J = {jl,h, ... ,jm} with jl < h < ... < jm (in the order of I), then EJ = EiI ®h ... ®h Ejm' For any finite set J' = {j~, ... ,j:n/} with j~ < ... < j:n, containing J, we define the linear mapping
which takes xiI ® ... ® Xj", to Yji ® ... ® Yj'rr,,' where Yj' = Xj' if j' E J and Yj' = (,jl if j' ~ J. Clearly (since we assume II(,j II = 1 for all j), iJ,J ' is a completely isometric embedding of E J onto E y subset of I containing J', we clearly have iJ',J"
0
iJ,J'
,
and if J" is another finite
= iJ,J'"
Therefore, we may unambiguously define the inductive limit of the system {EJ} equipped with the above embeddings. We will denote by
the resulting operator space. [More precisely, we introduce a vector space V containing each E J in such a way that the diagrams
all commute and such that V is the union of E J when J runs over all finite subsets of I. Then we obtain a norm on Mn(V) = UJ Mn(EJ) and we apply Ruan's Theorem. After completion V becomes an operator space denoted by (®jEIEj)h']
In the above situation, let 'H. be a Hilbert space and assume given, for each j in I, a complete contraction i j : E j --+ B('H.) such that ij(('j) = 1. For each finite subset J c I, say, J = {jl,'" ,jm} with jl < ... < jm, there is clearly a unique linear map from ®jEJ E j into B('H.) that takes xiI ® ... ® Xj", to iii (xii) ... ijm (Xjm)' By Corollary 5.4, this mapping extends to a complete contraction iJ: EJ --+ B('H.). Since ij(('j) = 1 for all j, for any finite subset J' containing J we have iJllEJ = iJ' Whence the next statement. Lemma 26.9. With the preceding notation, the system of maps {iJ 1 J c I, IJI < oo} extends unambiguously to 8 complete contraction from ( ®jEI E j ) h to B('H.). To illustrate this notion, let us return to the situation discussed in Theorem 26.1. Let us first show that the subset S naturally defines an
26. The Blecher-Paulsen Factorization operator space structure on its linear span. S. For any x in Mn(V) we define Ilxli n
391
Let V be the linear span of
= inf{lIaolilladl},
where the infimum runs over all possible factorizations of x of the form x = aODal, where ao, al are scalar rectangular matrices and where D is a diagonal matrix with entries in S. By Ruan's Theorem, these norms correspond to an operator space structure on V for which the inclusion V c A is completely contractive. Therefore, the inclusion map V c A extends to a complete contraction i: iT --+ A defined on the completion of V. Let us denote by E the resulting operator space, that is, E is iT equipped with the o.s.s. associated to the above norms II lin- We will form the infinite Haagerup tensor product (®jEI Ej)h in the case J = N (with its usual ordering), E j = E, i j = i, and = 1A for all j. (Recall that we assume 1A E S.) We will denote simply by Eoo the resulting operator space and by En = E 0h ... 0h E (n times) the finite tensor product considered as a subset of E oo , so that Un>l En is a dense linear subspace of Eoo. -
ej
By Lemma 26.9, the product mapping of A unambiguously defines an inductive system of complete contractions
which extend to a single complete contraction
which plays the role of an "infinite product map." 'vVe can now reformulate Theorem 26.1 with this notion.
Proposition 26.10. l-Vith the preceding notation, the hm properties in Theorem 26.1 imply that tIle completely contractive mapping Poo: Eoo --+ A is a complete surjection. l\fore precisely, it is surjective and the resulting isomorphism 0': Eoo/ ker(poo) --+ A satisfies
Remark 26.11. It is easy to check that the two properties in Theorem 26.1 are equivalent to the following one: (iii) Let £ = Un>l En C Eoo be equipped with the induced operator space structure and similarly for A C A. The restriction of Poo to £00 is a surjection from £ onto A that defines a completely contractive isomorphism £ / ker(poo} n £ --+ A
u:
such that lIu-1llcb :5 K. Similarly we have:
392
Introduction to Operator Space Theory
Proposition 26.12. In the situation of Corollary 26.2, let I = N, E = max(A), let i: E -+ A = A be the identity map, and let i j = i and ej = lA for all j. Then the infinite product map Poo: Eoo -+ A defines a completely contractive isomorphism a: Eoo/kerpoo -+ A such tllat lIa- 1 1lcb ::; I<. Thus any full operator algebra (in particular every unital C* -algebra A) is completely isometric to a quotient of an infinite Haagerup tensor product of copies of tlIe operator space max(A).
Let us now return to the disc and bidisc algebras A(D) and A(D2), as in Corollaries 26.4 and 26.5. Note that the linear span V of S = [1, z] (resp. S = [1, ZI, Z2]) in A(D) (resp. A(D2)) is isometric to .e~ (resp . .eD. By Theorem 2.7.2, the operator space E introduced above on the vector space V can be identified respectively with max(.ei) and max(.e1). Hence we obtain the following application of Proposition 26.10. Corollary 26.13. Tlle algebra A(D) (resp. A(D2)) is completely isometric to a quotient of all infinite Haagerup tensor product of copies ofmax(£i) (resp. max(.eD)· Remark. Actually, any full operator algebra A is completely isometric to an infinite Haagerup tensor product of copies of max(t'i2 »). Indeed, let 1= BA equipped with any total order, and let 3 be a countable disjoint union of copies of I simply ordered consecutively (we could use the lexicographical order on N x I). Thus an element j in 3 is simply an element of I = B A placed in one of the successive copies of I inside 3. We denote by (eo, el) the canonical basis of t'~2). Let E j = max(t'i2») and ej = eo for all j in 3. Then we have a mapping
P:
(®Ej) jE:J
-+
A
h
that is a complete metric surjection onto A. The latter is defined as follows: For any j in 3 corresponding to some a in I = B A we define Uj: max( t'~2») -+ A by uj(eo) = 1A and Uj(el) = a. Clearly the construction described in Lemma 26.9 yields a completely contractive map
P:
(®
jE:J
Ej
-+
)
A.
h
Note that any product ala2 ... an of elements in BA can be written as uh (eduh(ed ... uj" (ed by suitably choosing the placement jl < 12 < ... < jn of the elements ai, a2, ... , an in 3. The fact that P is a complete metric surjection then follows immediately from part (iii) in Theorem 26.1. Now let E be an arbitrary operator space. We can also apply Theorem 26.1 to the universal operator algebra 0 Au (E) associated to E, as defined in
26. The Blecher-Paulsen Factorization
393
Chapter 6. We will consider E as a subspace of of OAu(E) and will denote by e the unit element in OAu(E). Let E c OAu(E) be the operator space spanned bye and E, that is,
E = ICe + E. This space
E is the
unitization of E, already considered in Proposition 8.19.
ej
Then, we take again I = N, E j = E, and = e for aU j; we denote by (E)oo the infinite Haagerup tensor product associated to this data; and we let £ c (E)oo be the union of all the finite tensor products E Q9h ... r;;¢h E (n times) n ~ 1. We can then apply Theorem 26.1 as follows. Proposition 26.14. TIle product mapping from E Q9 E to OA.(E) extends to a complete contraction Poo from (E)oo onto 0 Au (E), which defines a completely isometric isomorpllism from (E)oo/ ker(poo) onto OA,,(E). More precisely, the restriction of Poo to £ defines a completely isometric isomorpllis111 between the (noncomplete) operator spaces £/ker(poo)n£
and T,,(E)cOA,,(E).
Proof. Let A = OA,,(E), A = Tu(E) and let S c K Q9min A be the unit ball of K Q9min E. Then the property (i) in Theorem 26.8 clearly holds with K = 1. Therefore, the proposition can be deduced from (ii) in Theorem 26.8 by a simple reformulation. • \Ve now return to the general setting of a family (Ej )jEI of operator spaces, given together with a family of complete contractions i j : E j ---4 B(1t) and a family (ej)jEI such that E E j and ij(ej) = 1 for all j in I. We will denote by A the (unital) subalgebra of B(1t) generated by UjEI ij(Ej ).
ej
We will say that the triple T = {{Ej I j E I}, {ej a "generating family of operator spaces."
Ij
E
I}, {i j I j
E
In is
Given such a generating family, we can introduce the enveloping unital operator algebra AT as follows: We consider the "collection" C T of all unital homomorphisms u: A ---4 B(Hu) such that IIuijllcb = 1 for all j in I, and we introduce the embedding
iT:
A
---4
E9 B(Hu) uEC,.
defined by jT(X) = E9UEC,. u(x). Note that, since the inclusion A c B(1t) belongs to C T, this map iT is injective; hence it defines a new operator space
394
Introduction to Operator Space Theory
(and operator algebra) structure on A. We will denote by AT the completion of A for the latter structure. Note that the mappings i j : E j --+ A C AT are completely contractive as mappings into AT! so that, by Lemma 26.9, we have a natural complete contraction from ( ®jEI Ej ) h into AT' But actually, we will need to enlarge the ordered set I: We will denote by I the totally ordered set that is the disjoint union of count ably many copies of I, ordered consecutively. Let {Ej I j E I} be the extended family where each E j (j E I) is repeated countably infinitely many times. Theorem 26.15. Let T be a generating family of operator spaces as above and let p: (®jEI E j ) h --+ AT be the natural completely contractive map, associated to the "duplicated" family {Ej I j E I}. Then p is a complete metric surjection, that is, it is surjective and it defines a completely isometric isomorphism from (®jEI E j ) hi ker(p) to AT' Moreover, if we denote byE the subspace of (®jEI E j ) h that is the union of the (algebraic) tensor products ®jEJ E j (with J running over all finite subsets of I), then the restriction ofp to £ induces a completely isometric isomorphism between the (noncomplete) operator spaces £ I ker(p) n £ and A viewed as a subspace of AT' Proof. A proof can be given using Theorem 26.8 applied to S = UjEI(h; 0 ij )(B1(.®m'n E j) and applied to A
= AT'
For the convenience of the reader, here is a direct argument (which in essence is the same): One first observes that p(£) = A, so that the induced map 0': £1 ker(p) n £ --+ A is a linear isomorphism. Note that the quotient space Q = £1 ker(p) n £ is equipped with a natural O.S.S. as a (noncomplete) quotient of a subspace of ( ®jEI E j ) h' Let B be the unital operator algebra obtained by equipping Q with the product x· y = a- 1 (a(x)a(y)). Then, after completion, B is a unital operator algebra, that is, we may view B as embedded into some B('H.). Indeed, this follows from Theorem 6.1, because the repetition of I infinitely many times in I allows us to show that this product map is a complete contraction from Q 0h Q to Q. But then 0'-1: A --+ B is a unital homomorphism such that a- 1i j coincides with the composition of the inclusion E j --+ £ foU
26. The Blecher-Paulsen Factorization
395
operator algebras naturally embeds into their free product Al * A 2 • \Ve will now see that the free product itself can be viewed as the quotient of an infinite Haagerup tensor product.
Corollary 26.16. Let AI, A2 be two unital operator algebras, and let Al *A 2 be their free product (as unital operator algebras). Let (Ej )JEN be the family of operator spaces defined by setting E j = Al if j is odd and E j = A2 if j is even. Then Al * A2 is naturally completely isometric to a quotient of (®jENEj)h·
Proof. We take I = {O, I}, EI = AI, Eo = A2 and take for i j the natural embedding into Al * A2 C B(1i). Then the algebra AT associated to this generating family actually coincides with At * A 2 , by definition of the latter. Thus Corollary 26.16 follows from Theorem 26.15. •
Chapter 27. Similarity Problems The following conjecture formulated in 1955 by R. Kadison is still open. It will serve as the motivation for all further developments in this chapter. Kadison's conjecture. Every bounded unital homomorphism u: A --+ B(H)
on a unital C* -algebra A is similar to a *-homomorphism, that is, there is an H --+ H such tllat
isomorphism~:
is a *-homomorphism. Note that u is only assumed to be a bounded Banach algebra morphism, and we conclude that after conjugation by ~ it becomes a C* -algebra morphism. More explicitly, our assumption are that u(xy) = u(x)u(y) V x, yEA and u(l) = 1, and the conclusion is that
VxEA When this conclusion holds, Kadison says that u is "orthogonalizable." Many partial results are known on this, mainly due to E. Christensen and U. Haagerup. In particular, this conjecture is known whenever A is a nuclear C*-algebra or when A = B(H) (or, more generally, when A has no tracial states) and also when A is a IIl-factor with property r. We will return on these examples later. Moreover, Haagerup [H6] (see also [Chr3]) proved that the conjecture is correct for all cyclic homomorphisms u: A --+ B(H). Nevertheless, the problem remains open in full generality, and, as shown by the next statement, it is really a problem on c.b. maps.
Theorem 27.1. ((H6J) Let u: A --+ B(H) be a unital homomorpllism on a unital C* -algebra. Then u is similar to a *-homomorphism iff u is c. b. and moreover lIuli c b = iF{IIClllll~lI},
where the infimum (which is actually attained) runs over all isomorphisms ~ for which ~u(-)~-l is a *-homomorphism. Thus, Kadison's conjecture boils down to the implication ?
lIuli < 00 => lIulicb < 00 for unital homomorphisms into B(H). Equivalently, we want to know whether complete boundedness is automatic for bounded homomorphisms. The above conjecture is closely connected to the following well-known question.
27. Similarity Problems
397
Derivation problem. Let A be a C* -algebra. Given a *-homomorpllism rr: A ----; B(H7r)' a bounded linear map 8: A ----; B(H7r) is called a rr-derivation if V a,b E A 8(ab) = rr(a)8(b) + 8(a)rr(b). lVe say that 8 is rr-inner, or simply "inner", if tIlere is T in B(H7r) stu'l] that
EA
Va
Tl]e problem is to
SllOW
8(a) = rr(a)T - Trr(a).
tbat all rr-derivations are inner.
Once again this problem turns out to be a problem on c.b. maps, because of the following result due to E. Christensen.
Theorem 27.2. ((CIlr5]) Let 8: A ----; B(H7r) be a rr-derivation (as above) on a C* -algebra A. Tllen 8 is inller iff it is c. b. A/oreover, we llave
11811cb = inf{21lTIII T wbere we bave set 8T(a)
E
8 = 8T },
B(H7r ),
(27.1 )
= rr(a)T - Trr(a).
Proof. If 8 is inner, say, if 8 = 8T , we clearly have by the triangle inequality 11811cb = 118Tlicb :::; 211TII; hence 11811cb :::; inf{21lTII I 8 = 8T}. The nontrivial direction is the converse for which we will use Theorem 27.1. Indeed, given a rr-derivation 8: A ----; B(H), the formula
u(a) = (rr(oa)
8(a») rr(a)
obviously defines a homomorphism into B(H EB H) ~ Af2(B(H». !\'1oreover, if 8 is c.b., u must also be c.b.; hence, by Theorem 27.1, u is similar to a *-homomorphism p: A ----; B(H EB H), that is, for some invertible~: H EB H ----; H EB H we have u(x) = ~p(x)~-l. Then let a = ~C so that u(x)a = ~p(x)C for all x in A. Since p(x) = p(x*)*, we have u(x)a = (u(x*)a)*, and hence
VXEA
u(x)a = au(x*)*.
In matricial notation, this becomes
8(X») rr(x)
(au
a 12 ) ( rr(x)
ah
a22
8(x*)*
which implies (consider the (2,2) and (1,2) entries only)
rr(x)a22
= a22 rr (x)
and
rr(x)a12
+ 8(x)a22 = a12rr(x).
(27.2)
Introduction to Operator Space Theory
398
Since a 2': 0 is invertible, :3 8 > 0 such that (ay, y) 2': 811yll2 for all y in H EB H; hence the same holds for .:122 on Hand a22 must be invertible. Moreover, a22 commutes with 1I'(x) so that (27.2) implies
with T = a12a2l, which shows that 8 is inner. Unfortunately, this method only seems to yield T with IITII ~ 411811cb. In order to replace this factor 4 by the correct factor 1/2, we will need more work. First we will reduce to the von Neumann algebra case. Assume that 8: A -> B(H7r) is c.b. Then, by a routine argument, 11' and 8 both extend to normal maps 1f: A ** -> B (H7r) and 8: A ** -> B (H7r), so that 1f is a *- homomorphism and 8 is a 1f-derivation with the same c.b. norm. Note that 8 must vanish on the kernel of 1f (since the latter ideal admits approximate units). Thus, replacing A** by its image under 1f, we can reduce to the case when A is a von Neumann algebra and 11' is its natural inclusion into B(H). Moreover, by the first part of the proof, we can assume that 8 is inner. Thus it suffices to prove (27.1) when 8: AI -> B(H) is an inner derivation, say, 8 = 8T , on a von Neumann subalgebra 1If C B(H). Recall that 8T is defined by 8T (x) = Tx - xT. Then we will prove 11811cb
= 2 inf{IITIlI 8 = 8T }
or, equivalently,
118Tllcb = 2 inf{IIT - Till
TE
JlI'},
(27.3)
where 111' denotes the commutant of 1\.1. We will denote by d(T,1\.1') the right side of (27.3). Here is a quick proof of (27.3) taken from [Chr5] (but the argument is similar to Arveson's one in [Ar3, p. 12]). We will prove that
d(T, 1\.1') ~ (1/2)118Tllcb'
(27.4)
First note that 118T llcb is equal to ~e norm of the operator I B (f 2 )08T acting on the von Neumann algebra B(f.2)01\.1 C B(f.2(H)) (Le., the one generated by the algebraic tensor product B(i2) 01\1)). Let B(H)* denote the predual of B(H). By Hahn-Banach, we have d(T,M')
= sup{I/(T)11 IE B(H)*,I 1. M', II/II = I}.
(27.5)
Consider IE B(H)* with I 1.. M' and 11/11 = 1. By the classical identification of B(H). with the space of trace class operators on H, there are unit vectors e,1] in f. 2 (H) such that "Ix E B(H)
J(x) = (/0 x)e, 1]).
(27.6)
27. Similarity Problems
399
Hence, for any x in !If', (1181 x)~ is orthogonal to 1]. Let p be the orthogonal projection on €2(H) onto the closure of {(1 181 x)~ 1x E !If'} c €2(H). Clearly p~ = ~ and prJ = o. l\loreover, since p commutes with (1181 Af'), we have p E B(€2)I8IM. By (27.6), we can write
If(T)1 = 1((1 181 T)p~, (1 - p)T}) 1 ~
11(1-p)(II8IT)pll
~
11(1181 T)p - p(I 181 T)II
=
1I[(1I81T)p-p(II8IT)]pll
= ~11(1I81T)(2p-1) - (2p-1)(II8IT)11 = ~11(11818T)(2p-1)11: hence note that 2p - 1 = p - (1 - p) is unitary)
IF(T)I ~ ~IIII8I8TII Taking the supremum over all possible The converse is obvious.
f
=
~118Tllcb.
and using (27.5), we obtain (27.4). •
As we just saw, there is a very natural way to construct homomorphisms using derivations. Indeed, given a 1f-derivation 8: A ---> B(H,,), the formula
, ( ) = (1f(a) u a
0
8(a)) 1f( a)
obviously defines a homomorphism into B(H" EB H,,) ::: lIh(B(Hrr)). I\:[ore generally, given two homomorphisms
on the same,operator algebra, and given a (1fI' 1f2)-derivation 8: A ---> B(H) (by this we mean \j a, bE A, 8(ab) = 1fI(a)8(b) + 8(a)1f2(b)), the formula
u(a) = (1f I (a)
o
8(a)) 1f2(a)
(27.7)
defines a homomorphism from A to B(H EB H) ::: M2(B(H)). In particular, this observation shows that, if a C* -algebra A satisfies Kadison's conjecture, then the derivation problem has an affirmative solution for A. Recently, E. Kirchberg [Ki9] showed that the converse also holds, so that Kadison's similarity problem for a C* -algebra A is actually equivalent to the derivation problem for A as formulated above. Actually, Theorem 27.1 also holds for non-self-adjoint operator algebras. This is a very important result due to Paulsen.
400
Introduction to Operator Space Tlleory
Theorem 27.3. ((Pa4J) Let A c B(1i) be a unital operator algebra. Let u: A ----> B(H) be a unital homomorphism, and let C be a constant. The following properties are equivalent:
(i) u is c.b. and Ilulicb :::; C. (ii) There is an isomorphism~: H ----> H witll II~II 11~-111 x ----> ~U(X)Cl is completely contractive.
:::;
C such that
(iii) There is a Hilbert space ii with H c ii, a *-homomorphism 7r: B(1i) --> B(ii) and an isomorphism~: H ----> H with II~II 1I~-111 :::; C such that
\i
0
E
A
Proof. The implications (iii) =} (ii) =} (i) are obvious. We will show (i) (iii). Assume (i). By the fundamental factorization Theorem 1.6, we can write u(x) = V7r(x)W, =}
where 7r: B(1i) ----> B(ii) is a representation (C*-sense) and where V: ii ----> H and W: H ----> ii satisfy IIVII IIWII :::; C. Let El = span[7r(A)W(H)f.~ that El is invariant under 7r(A) and contains W(H) (and actually is minimal with these properties). We may rewrite the above factorization trivially as
(27.8) . Let E2 = ker(l'IE1). We claim that E2 c El is also 7r(A)-invariant. The simplest way to see this is to'Observe that
\iaEA
(27.9)
Indeed, it suffices to check (27.8) on a vector k of the form k = 7r(x)Wh with h E H, x E A. We then have V7r(a)k = V7r(ax)Wh = u(ax)h = u(a)u(x)h = u(a)V7r(x)Wh = u(a)Vk, which establishes (27.9). From (27.9) it is clear that 7r(a}E2 C E2 for all a in A. We will work with the (so-called semi-invariant) subspace El 8 E2 = El n Ei- (as in the proof of Theorem 6.3). We start by theQQservation that, since u(l) = 1, we have VW = I, and hence V(Et} = H, so V is surjective. Passing to the quotient by ker(V) we obtain from (27.8)
\ixEA where S: El 8 E2 ----> H is an isomorphism with IISII :::; IIVII. Moreover, since PE1eE27r(x) vanishes on E2 (since E2 is invariant), we actually have
\ixEA
27. Similarity Problems Taking x
=
401
1, this gives
1 = 8PE18E2ll', which implies that PEl 8E211' is the inverse of the isomorphism 8, or, equivalently, 8- 1 = PEI8E2lV. Hence \fxEA
which shows that 8- 1 u(·)8 is a *-homomorphism and we have 11811 :s: IIVII, 118- 1 11 :s: IIWII; hence 11811118- 1 11 :s: Ilulicb' This gives us (iii), except that 8 is an isomorphism from El 8 E2 to H. But since El 8 E2 and H have the same (Hilbertian) dimension, they are isometric; hence there is a unitary U: El 8 E2 ----+ H such that 8 can be written as 8 = f,U. This gives us an isomorphism f,: H ----+ H with 11f,111If,- 1 11 :s: Ilulicb such that Clu(.)f, = U(8- 1u(.)8)U* is a *-homomorphism. This completes the proof that (i) ::::} (iii). •
Definition 27.4. We say that a lInital operator a.lgebra A C B(1t) has tIle similarity property, in short SP, if any bOllllded ullital llOmOmol'pllism u: A ----+ B(H) is alltomatically completely bOllnded. As we will see, this property is closely connected with a certain notion of "length" for A, as follows.
Definition 27.5. An operator algebra A C B(1i) is said to be of length :s: d if tIl ere is a constant ]( such tllat, for any n and any x in IIIn (A), there is an integer N = N(n,x) and scalar matrices Cto E lIfn,N(e), 01 E MN(C) , ... ,Ctd-l E lIfN (C) , Ctd E MN,n(C) togetller with diagonal matrices D 1 , ... , Dd illllfN(A) satisfying
(27.10)
We denote by leA) the smallest d for wllidl this holds, and ,,'e call it the length of A (so that A lIas length :s: d is indeed the same aB l(A) :s: d). Remarks. (i) Fix an integer n and consider x in lIfn(A). We denote d
II x ll(d) = inf{II i=O
d
IICtil1 II IIDill}, i=1
where the infimum runs over all possible representations x = Ct OD l (\:l ..• DdD:d of the form appearing in (27.10) above. By (27.11) below, this is a norm on Mn (A). Clearly
Introduction to Operator Space Theory
402
(ii) When A is unital (or has a contractive approximate unit, which is the case whenever A is a C*-algebra) we have
Indeed we can always insert a redundant diagonal matrix with coefficients all equal to the unit. Moreover, if A is a C*-algebra, the results of Chapter 26 (see Corollary 26.6) show that
(iii) Let E = max(A). Let Pd: E ®h ... ®h E --+ A be the product map. Clearly Pd is completely contractive (since it is completely contractive on A ®h ... ®h A). We claim that the norm II . lI(d) on l\fn(A) is equal to the quotient norm Mn{E ®h ... ®h E)/Mn(ker(Pd)). Indeed, consider x E Mn(A) and let Y E Mn(E ® ... ® E) be such that (IM.. ® Pd)(Y) = x. Then we have (27.11)
where the infimum runs over all possible such y. This follows immediately from (5.6) combined with Theorem 3.1. But, since Pd is completely contraftive on the completion E®h" '®hE, the equality (27.11) must remain (a foytiori) true when the infimum runs over all Y in Mn{E®h'" ®hE) such that(hr" ® Pd)(Y) = x. Proposition 27.6. If £(A) ::; d as in (27.10) above, tllen A has the similarity property (SP), and for any bounded homomorphism u: A --+ B{H) we have lIulicb ::; Kllull d . Proof. Let u: A --+ B(11.) be any bounded homomorphism. Consider x E Mn{A) factorized as in (27.10). We then have
(IM" ® u)(x) = ao(IMN ® U){Dl)'" (hlN ® u)(Dd)ad, and clearly for any i
= 1, ... , d
hence
lI(fM " ® u)(x) II
d
d
o
1
::; lIuli d II lIaili II IIDili ::; Kllulldllxll·
•
Curiously, the preceding sufficient condition, which a priori seems rather strong, is actually also necessary for A to satisfy the SP. This is the main result of [P17j (see also [PI8]).
27. Similarity Problems
403
Theorem 27.7. Let A be a unital operator algebra. The following assertions are equivalent.
(i) There is a nlllnber c > 1 sudl tlmt allY Illlita.J llOlllomorpllism u: A B(H) with /lull ::; c is completely bounded. (ii) A satisfies the similarity property.
~
(iii) Tllere a,re (\' ~ 0 and a constant I< SUdl that any bounded IIllital llOll1Omorphism u: A ~ B(H) satisfies Ilulicb ::; I
I::II<
V
u ullitalllOmolllorphism Illlllcb::; I
then we have d(A)
=
£(A),
(27.12)
and tIle infimum defining d(A) is attained when a
= R(A).
Remark. Let A he a unital C* -algebra. Assume that there is a constant a ~ 0 such that, given an arbitrary *-homomorphism 7r: A ~ B(H), any 7r-derivation 8: A ~ B(H) satisfies
11811cb ::; a11811·
(27.13)
Then, for any bounded unital homomorphism '11: A ~ B(H), we have Ilulicb ::; Ilull<>. This is proved by a modification of Kirchberg's argument in [Ki9J; see [P17J for details. By the preceding statement, this implies that d(A) ::; [aJ (integral part of a), which leads us to conjecture that (in the C*-case) the best possible a in (27.13) is "automatically" an integer. Proof of Theorem 27.7. We have already shown that (iv) =? (iii) in the preceding statement. lVIoreover, (iii) =? (ii) =? (i) are obvious. Thus for the equivalence it suffices to show (i) =? (iv). Assume (i). By an easy direct sum argument, we can show that there is a constant (3 such that any u as in (i) with Ilull ::; c must satisfy Ilulicb ::; (3. Then we select d so that (3
L c-
k
< 1/2.
(27.14)
k>d
We will show that £(A) ::; d. Equivalently, it suffices to show that there is a constant I<' (independent of n) such that for all x in 1I1n(A) we have (27.15) Consider x with II x /lM,,(A) < 1. We will show that x can be written as x
= x' + x"
with IIx'lI(d) ::; (d
+ 1)(3 and
IIx"IIM,,(A)
< 1/2.
Introduction to Operator Space Theory
404
By a well-known iteration argument, this implies (27.15) for some K'. Let S = {c- t BA} U {lA}. We will apply Theorem 24.1 to this set. We have for any unital homomorphism u: A --+ B(H) sup lIu(x)11 = 1 =}
lIuli c b ~ (3.
xES
Hence any x in Mn(A) with
where
Ilxll < 1 can be written in the form
nIIO:il1 < (3 and where each matrix Di is of the form
for suitable elements Xi in the unit ball of A. (Indeed, by rearranging the lines and columns of the O:i we can assume for simplicity that Di has this70r .) Then we denote for Z in C ZXi
D;(z)
~
(
0
1
Let x(Z)
= o:oDt(z)O:t ... Dm(z)O:m.
Developing the product we find x(z)
=
""m ~o
Z
k Cl.k.
Note that sup Ilx(z)llu.,(A) ~ (3. Izl:51
Hence, by Cauchy's formula (since Cl.k
= J e-iktx(eit)dt/27r), we have (27.16)
Then we decompose x as
x = x' +x"
27. Similarity Problems with x'
405
= Lg c-ktl k and
Note that our original choice of d in (27.14) guarantees with (27.16) that
Ilx"IIM,,(A) < 1/2. Thus it remains only to majorize Ilx/ll(d) as announced earlier. In order to do this we will "lift" x(z) up into the algebra OAu(E) studied in Chapters 6 and 8, with E = max(A). Let e be the unit in OAu(E). Let E be the unitization of E, that is, the span of e and E in OAu(E). For any i we denote
where Xi denotes Jl1n(E). We let
Xi
E
E viewed as an element of OAu(E). Thus Dj(z) E
X(Z) Then x(z)
E
E is
= O:ODl(Z)O:l'" Dm(z)O:m.
Mn(OAu(E», and if
Izi
~
1,
Ilx(z)IIM,,(OA,,(E» ~
II 1I00ili < /3.
Moreover, if we denote by fi: OAu(E) ---+ A the product map, then ll'd" Q9 P applied to x(z) gives us back x(z). Expanding x(z) as a polynomial in Z we find x(z) = zk3. k ,
L:
where 3. k
E
Mn(E.Q9· .. Q9 E) (k times). By Cauchy's formula again, we have
l13. k ll lll,,(OA u (E» < /3; but now, since E Q9h ... Q9h E is completely isometrically embedded into OAu(E) (see Proposition 6.6), 3. k can be identified with an element of JlIn (EQ9 '" Q9E) (k times) such that l13. k Il M "(E0""'0,,E) < /3. Since (1M" Q9 P)(X(z» = x(z), we have (1M" Q9 Pk)(3. k ) preceding remarks give us for any k ~ d
= tl k.
Hence the
406
Introduction to Operator Space Theory
Finally, we conclude as announced
• Corollary 27.S. If d(A) :::; d, then any linear map u: A ---> B(H) such that there are bounded linear maps ~i: A ---> B(Hi+l' Hi) (i = 1, ... d) witll H d+ 1 = H 1 = H satisfying
must be completely bounded and we have d
Ilullcb:::; KIT II~ill, i=l
where K is a constant independent of u. Proof. By the preceding result we have £(A) :::; d. Then this is easy to prove, just like Proposition 27.6 above. •
algebr~d
Remark. The analog of Theorem 27.7 for dual operator normal homomorphisms) is proved in [LeM7]. Examples. Obviously any finite-dimensional A satisfies d(A) = 1. Assume A infinite-dimensional. Then, if A is a nuclear C*-algebra, we have d(A) = 2 (due to Bunce and Christensen; see [Chr 1-2]). Moreover, if A is a C*algebra without tradal states, such as B(H), then d(A) :::; 3 (see [H6]), and we proved in [P17] that d(B(H» = 3 (see Exercise 5.10 for a direct proof that £(B(H)) :::; 3). Moreover, if Kl denotes the unitization of K, then for any C*-algebra B we have d(Kl ®min B) :::; 3. In [PIg] it is observed that d(A) ~ 3 if A is a Ill-factor and that d(A) = 3 for the hyperfinite one. More recently, in [Chr6] it is proved that d(A) = 3 for any fIt-factor with property r, improving on the previous successive estimates d(A) :::; 44 ([Chr4]) and d(A) :::; 5 ([PIg]). Finally, in [PI8] we prove that, for any integer d, there is a unital (non-self-adjoint) operator algebra Ad with length equal to d (Le., we have d(Ad) = d). However, it remains open whether there exist C*-algebras, or uniform algebras, or Q-algebras with arbitrarily large finite length. Exercise Exercise 27.1. Let f c A be a closed two-sided ideal in an operator algebra. Show that €(A/ I) :::; £(A). Then prove that £(A) :::; max{ £(f), £(A/ f)},
and that the equality holds (Le., £(f) :::; £(A» when A is a C*-algebra (use the fact that f admits a quasi-central approximate unit) or when A ~ f x (A/f).
Chapter 28. The Sz.-Nagy-Halmos Similarity Problem This chapter concentrates on the following question: How can we recognize when an operator T in B{H) is similar to a contraction? Of course T similar to a contraction means that there is an isomorphism H ---+ H such that Ile-ITell :0:; 1. This problem originates from a question of Sz.-Nagy [SNj, who observed that an obvious necessary condition is that T should be power bounded, that is, we should have
e:
sup II Tn II <
00.
n;::>:l
He asked whether the converse holds (and proved it for compact T), but Foguel [Foj quickly found a counterexample. Soon after that, Paul Halmos (and perhaps others, too) noticed that there was a stronger necessary condition, namely, T should be polynomially bounded, which means that there is a constant C such that for all polynomials P we have
IIP{T)II :0:; CllPllexll
(28.1)
where
IIPII"" =
sup IP{z)l·
zED
Here, of course, if P{z) = L~ akzk, we set P{T) = aoI + L~ akTk. In that case, we will say that T is polynomially C -bounded. This is indeed necessary: By von Neumann's classical inequality {see (8.11)) any contraction T must satisfy (28.1) with C = 1. Hence, if T = e-IOe with 11011 :0:; 1, we must have (28.1) with C ~ lIe- 1 11 lIeli. Foguel's example was shown to fail (28.1) ([Leb]). Halmos then asked in 1970 (see [Hal]) whether this condition is sufficient, that is, whether, conversely, every polynomially bounded operator is similar to a contraction. This question was recently answered negatively in [P20j. The original proof of the polynomial boundedness in [P20j, which was rather complicated, was simplified in [Kis1j and in [DPj. In [DPj, the polynomial bounded ness is deduced very simply from the "vectorial Nehari theorem," which we discussed in §9.1. We will follow the same approach. See [PlOj for more background on these questions. We start with a very useful criterion for similarity to a contraction due to Paulsen [Pa4]. This is nothing but Theorem 25.2 applied when A is the disc algebra. Note. Throughout this chapter, unless specified otherwise, we always equip the disc algebra A(D) with its minimal operator space structure, associated, for instance, to the embedding A(D) c C(8D).
408
Introduction to Operator Space Theory
Theorem 28.1. ([Pa4]) Consider an operator T in B(H) and fix a constant C 2:: O. The following are equivalent: (i) There is an isomorphism e: H ---+ H with Ilell lie-III ::; C such that lIe-ITell ::; 1. (ii) The homomorphism P ---+ P(T) defined on polynomials extends to a c.b. homomorplIism UT: A(D) ---+ B(H) with IIUTlicb ::; C. (iii) For any n and any n x n matrix (Pij ) with polynomial entries, we have
II(Pij(T»IIM,,(B(H» ::; C sup II(Pij(z»IIM". Izl::;1
(iv) There is a Hilbert space ii with ii :J H, a unitary operator U on and an isomorphism e: H ---+ H with lie II lie-III ::; C such that
ii,
Proof. The equivalence between (ii) and (iii) and (iv) => (i) are obvious. (ii) => (iv) follows from Theorem 27.3 and (i) => (ii) from Theorem 8.7. • Note. When C = 1 in the preceding statement, (i) just means that T is a contraction. In that case, the equivalence between (i) and (iv) is known as Sz.-Nagy's dilation theorem, saying that any contraction admits a unitary dilation. The next statement gives a counterexample to Halmos's question. Theorem 28.2. For any e > 0, there is a polynomially (1 + e)-bounded operator on £2 that is not similar to a contraction. More precisely, for any n there is an n x n matrix Te,n E J.[n that is polynomially (1 + e)-bounded but such that (28.2)
where 8> 0 is a number independent of (e, n). Notation. For any T in B(H), assumed similar to a contraction, we let
Sim(T) = inf{IICIlillell}, where the infimum runs over all the isomorphisms e: H ---+ H such that lIe-ITell ::; 1. Remarks. (i) If Te,n is as above, then T = EBn Te,n is polynomially (1 + e)bounded but is not similar to a contraction. Indeed, it is easy to check directly that the restriction of an arbitrary operator T in B(H) to an invariant subspace E c H satisfies Sim(l1E) ::; Sim(T). (This can also be seen from Theorem 28.1.)
28. The Sz.-Nagy-Halmos Similarity Problem
409
(ii) The lower estimate (28.2) should be compared with Bourgain's upper bound [Bou4], slighly improved in Example 25.23: For any c > 0, n ~ 1 and any polynomially (1 + c)-bounded T in Mn we have Sim(T) :-:; K(1
+ c)2Log(n + 1),
where K is a numerical constant. It would be nice to close the gap between these two estimates. In particular, in (28.2), can (Log(n + 1))1/2 be replaced by Log(n + 1)? The obstruction to "similar to a contraction" will come from the following. Lemma 28.3. Let (Bn) be a sequence in B(H) such that, for some f3 > 0, we have (28.3)
Let T, V, IF ill B(H) be such tllat Vn~O
If
S~PN-1/21IL~ Bn Q9 Bnll =
00
(in particular, this holds if Bn is a spin system or a sequence satisfying the CAR), tllen T is llot similar to a contraction. More precisely. we always have (28.4)
Proof. We use the implication (i) ::::} (iv) in Paulsen's criterion. (This is actually Sz.-Nagy's dilation theorem.) Let (Bn) be as in Lemma 28.3. Assume (iv) in Theorem 28.1. Then we have
hence there are operators VI, M'l such that
Then we can write
But, by the classical spectral theorem, the C* -algebra generated by the unitary (hence normal) operator U is commutative and can be identified with
410
Introduction to Operator Space Theory
C(cr), where cr is the spectrum of U and of course cr C {z Hence, by (28.3),
Eel Izl
which, by our assumption on (Bn), is impossible when N precise recapitulation yields (28.4).
-+ 00.
= I}.
A more •
Remark. Let E C A(D) c C(8D) be the closed linear subspace spanned by {z2n In;::: I}. The operator space interpretation of the preceding lemma is simply that the linear mapping v: E -+ B(H) defined by v(Z2") = Bn is not completely bounded. Indeed, the preceding argument shows that
The crucial step to prove Theorem 28.2 is the following.
)
Lemma 28.4. Let H = £2' Let (Bn) be any sequence in ]}{H) satisfying (28.3). Then, for any 0 < c ::; 1, there is a polynomially(l + c)-bounded operator T in B(£2) and operators V, W with IIVII IIWII ::; 4,&-1 SUell that
and actually also V k
Vn;:::O
Bn = VT 2" W
rt {2n In;::: O}
0 = VTkW.
To prove Lemma 28.4, we need to construct a special class of polynomially bounded operators. We first discuss operators of the form: S* T= (
o
(28.5)
where S: £2 -+ £2 is the shift and where r: £2 -+ £2 is a Hankel operator, which means that rs = s*r. Operators of the form (28.5) were first considered by Peller [Pel]' who explicitly proposed them as possible counterexamples for the Halmos similarity problem. Independently, these examples also appeared, together with analogous considerations, in unpublished work by C. Foias and the late J. P. Williams (see also [CCFW) and [CCl-2j). But the hope to find a counterexample among these operators was much reduced by Bourgain [Bou4) in 1985, and finally, in the summer of 1995, Alexandrov and Peller [APel finished it off by showing that such a T is polynomially bounded iff it is similar to a contraction, and, moreover, this happens iff the "symbol" of r, the function
28. The Sz.-Nagy-Halmos Similarity Problem
411
cp(t) = En>o rOneint, has its derivative cp' in BMO. However, all this applies only to Hankel matrices with scalar entries, that is, with r ij E
8:
be the "multivariate" shift operator defined by
8:
8=S0
I H, or, equivalently,
(ho, hI, ... ) ~ (0, ho, hI,·· .).
Let r: £2(H) ~ C2(H) be a Hankelian operator with associated matrix where rij E B(H). The Hankelian character of r means that
(rij ),
8*r = r8, or, equivalently, that r ij depends only on i {a(n) In E N} in B(H) such that
+ j,
that is, there is a sequence
r ij = a(i + j). Our examples will be of the form Tr
=
( s* 0
(28.6)
where r: £2(H) ~ £2(H) is a Hankel matrix with entries in B(H), so that Tr acts on £2(H) EEl £2(H). A simple calculation shows that, for any n ~ 1,
n_ (8*n0
nr~n-I).
sn
Tr -
,
hence for any polynomial P we have
P(T, ) r
= (P(8*) r P'L8 )) . 0
P(S)
(28.7)
Thus (as Peller observed [Pe2]) Tr is polynomially bounded iff there is a constant C such that for any polynomial P
IIrp'(8)11 ::; CllPlioo. Remark. In passing, note that the homomorphism u: P ~ P(Tr) is associated to a derivation just like in (27.7): Indeed, if we set 8(P) = r P'(8), then 8(PQ) = P(8*)8(Q) + 8(P)Q(8) for all polynomial P, Q. Let 1'n: £2 ~ £2 be the basic Hankel operator defined (for any n ~ 0) by 1'n =
L i+j=n
We will use the following elementary fact.
eij·
412
Introduction to Operator Space Tlleory
Sublemma 28.5. Let D be the unbounded diagonal (derivation-like) operator defined on the canonical basis of £2 by .
Let an
= T n"(2"D =
I:
eij(j/2 n ).
i+j=2"
(Note in passing that "(2" = an + a~.) We then have: (i) For any polynomial P and any n ~ 0
(ii)
I: a~an
~ 4/3.
n~O
Proof. (i) By the Hankelian property of "(2", we have "(2"P(S) = P(S*h2"; hence Thus it suffices to check that
DP(S) - P(S)D
= SP'(S),
which is entirely elementary. (Indeed, we may assume P(S) = sm; then SP'(S) = msm and (Dsm - smD)(ej) = mem+j = SP'(S)ej for all j ~ 0.)
(ii) It is easy to check that
and so that
L a~an is a diagonal operator on £2 with norm
III: a~anll
::;
s~p L
• 2"~i
T 2n i 2 ~
I:
2- 2n
= 4/3.
n~O
Theorem 28.6. Fix (3 ~ O. Let Bn in B(H) be such that
•
28. Tile Sz.-Nagy-Halmos Similarity Problem
413
Let f
= LTn'Y2" S ® Bn
E
B(£2(H».
(28.8)
n2':O
Then, for any polynomial P, we have
IIfP'("S) II =
II L
Tn 'Y2"SP'(S) ® Bnll ::::; 4,611P1ICX)·
(28.9)
Consequently, the operator Tr given by (28.6) is polynomially bounded. However, if
S~PN-l/21IL; Bn ® Bnll =
00
(in particular, if Bn is a Spill system or a sequence satis(ving the CAR), then Tr is not similar to a contraction.
Proof. The idea of the proof is to start with the case Bn = enl, which will turn out to be very easy using Sublemma 28.5. Then we observe that we have a bounded linear mapping v: B(£2) ---> B(H) with 111'11 ::::; ,6 such that
v(end = Bn· (Indeed, we simply take v = wP, where P: B(£2) ---> B(C 2) is the natural contractive projection onto span(end and where w(enl) = Bn.) Note that we have (28.10) and this is a Hankel operator on £2(H). Hence, if we let
and then we have
B = (Id ® v)(A); hence by Corollary 9.1.6 IIBII ::::; 111'11
IIAII : : ; ,6I1AII·
(28.11)
But now IIAII is very easy to estimate! Indeed, by Sublemma 28.5 we have IIAII = Ilz)anP(S) - P(S*)anJ ® enlll
::::; IILa n ®en1111IP(S)1I ::::; 211P1Ioo IlL a~an 111/2 ::::; 4/V3I1Plloo ::::; 411P11CX).
+ IIP(S*)IIIILan ®enlll
414
Introduction to Operator Space Theory
Thus, recalling (28.10) and (28.11), we obtain (28.9). This yields the polynomial boundedness of Tr as announced. Finally note that the 2 x 2 matrix of P(Tr ) admits r P'(S) as its (1 2)-entry and, in turn, if P(z) = LPkZk, the (0 O)-entry of rp'(S) is given by:
Thus we obviously have operators V, W with
IIVII, IIWII
s 1 such that for all
P (28.12) In particular, for all n 2: 0
Bn =VTr2" W; hence we deduce the second part of Theorem 28.6 from Lemma 28.3.
•
Remark. Let r be a Hankel operator with entries r ij ~'n B H). Assume that there is a finite-dimensional subspace K c H such tha either all the operators r ij have range in K, or all their adjoints have r ge in K (so that we can think of B(H) as replaced by either B(H, K) or B(K, H)). Then· (see [DP]) if an operator Tr of the form (28.6) is polynomially bounded, it is similar to a contraction (Le., there is no counterexample in this class). This extends the case dim(H) = 1 proved in [APe]. Proof of Lemma 28.4. We use r given by (28.7) and replace Tr by Tor with 0 = c/(4(3). Then by (28.9) and (28.7) we have
and using (28.12) (with Tor instead of Tr hence with oBn instead of Bn) we obtain Lemma 28.4 with IIVII IIWII 1/0. •
s
Proof of Theorem 28.2. Let (B 1 , ... , B N ) be a system satisfying the CAR (see Theorem 9.3.1) on a Hilbert space HN. Recall that we may assume dim(HN) = 2N. Moreover, by Exercise 3.7, we know that
Let r = L~=o 2-k'Y2kS I8l Bk, as above, acting on £2(HN), and let Tr be the associated operator. Let KN C £2(HN) be the subspace spanned by {eij I8l HN I 0 :$ i,j :$ 2N}. Note that KN is a reducing subspace for r. Moreover, since KN (resp. K N) is invariant for S* (resp. S), the compressions
28. The Sz.-Nagy-Halmos Similarity Problem P ---4 PKN P(S*)IKN and P if we let s = PKNSIKN and
then T is polynomially (1
---4
"y
415
PKN P(S)IKN are homomorphisms. Therefore, = PKNrIKN' and if we define (for 0' > 0)
+ 4a!3)-bounded,
and by (28.4) we have
(a/2)N 1 / 2 ~ ;3Sim(T). On the other hand, T acts on a Hilbert space of dimension 2dim(KN) 2(2N + 1)dim(HN) = 2(2N + 1)2N. Hence, taking a = c(4;3)-1 and (say) n = 2(2N + 1)2N, we find a polynomially (1 + c)-bounded n x n matrix T with c/(4;32)N 1/ 2 ~ Sim(T),
•
and since N ':::'. Log(n), this yields (28.2).
Let T E B(H) be polynomiaUy bounded, and let 11: A(D) ---4 B(H) be the homomorphism taking P to P(T). Let Un: lIfn(A(D)) ---4 lIfn(B(H)) be the homomorphism I IIf .. ® U. By Theorem 28.1, we know that T is similar to a contraction iff sUPn>l Ilunll < 00, and we are thus lead to try t.o compare lIuli and II Un II; but it tUlTIS Ollt that only a very crude estimate can hold, as shown by the next statement. Theorem 28.7. There is a numerical constant K llOmomorpllisIIl u: A(D) ---4 B(H), we have
SUell
that, for any lInital (28.13)
Conversely, there is a 8 > 0 SUdl that for allY 0 < c ~ 1 and allY n ~ 1 there is a unitalllOIIlomorphism u: A(D) ---4 B(H) with lIuli ~ (1 + c) such that (28.14) The upper estimate (28.13) is a rather simple consequence of the following result due to Bourgain [Bou3] (see [P21] for a simpler proof). Theorem 28.8. There is a constant K such tlmt, for any bounded linear map v: A(D) ---4 H (H Hilbert), we have
v Ii
E A(D)
(i ~ n)
(
L
IIv(/;) 112 )
1/2
~
Kllvll sup Izl=1
(L: IIi(z)1 2) 1/2 . (28.15)
Introduction to Operator Space Tl1eory
416
Furtl1ermore, for any linear map u: A(D)
-4
B(H) we l1ave (28.16)
Note that (28.16) follows immediately from (28.15) applied to v{f) = u{f)e with fixed in the unit ball of H.
e
Proof of Theorem 28.7. Let (Pij ) be an Mn(A(D». Let
n
x
n
matrix in the unit ball of
n
}j =
L
eij
® u(Pij ).
i=l
By (28.16) we have for each j
On the other hand,
L
eij
® u(Pij )
i,j
Hence we obtain Ilunll ::; Kfoliuli, which proves (28.13). We now turn to (28.14). We will use the following fact, which is easy to prove using random matrices and a concentration of measure argument (see Exercise 28.1). For some numerical constant {3, we can find for each n an n-tuple (Bb ... , Bn) in Afn satisfying (28.3) but such that (28.17) Again we set a with
= e(4{3)-1. Now, applying again the preceding construction n
r
=
2:
'Y2k
S ® [aBk],
k=l
we then find for the associated map u: P
-4
P(Tr )
28. The Sz.-Nagy-Ha.lmos Similarity Problem
417
hence, by (28.12) (applied with Bk replaced by oBd and by (28.17),
and Tr is polynomially (1
+ c)-bounded,
so
Ilull ::; 1 + c.
•
Exercise Exercise 28.1. Show that there is j3 > 0 sHeh that, for each n, there is an n-tuple (B 1 , ... , Bn) in !lIn such that
(28.18) and (28.19) (Actually, one can even show this with an n-tuple of unitary matrices in 1I1n.)
SOLUTIONS TO THE EXERCISES Exercise 1.1. Consider u: R - t C. Fix n 2:: 1, and let a E Mn(R)~ Mn Q9 R be the column matrix with entries in R defined by a = 2::7 ejl Qgelj' We have (IM" Q9 u)(a) = 2::7 ejl Q9 u(e1j). Hence, by Remark 1.13, we have
But now let (Uij) be the matrix associated to u, so that u( e1j)
=L
Uijeil·
i
(ft.1u,;I') 1/' ~ (L; lIu(el;)II') 1/' ~ IIL~ u(el;).u(e~ '" lIuliob. whence IlullHS ::; IIulicb' To prove the converse, consider a finite sequence Xj in lIfn. We can write:
L L UijXj
Q9 eil
j
M,,(C)
M,,(C)
2
) 1/2
lIf"
(~~ I ~XjX; ::; IIuIIHS ilL: Xj e1j II <
. )
U ijl2
t
J
J
Q9
1/2
III"
l\f,,(R)
'
and we conclude IIulicb ::; IIuIIHS. Exercise 1.2. We first recall that II Tni Rn : Rn - t Cnllcb = II Tni c" : Cn - t Rnllcb = y'n. Let H be any Hilbert space. Consider now x = (Xij) in Mn(B(H». We claim that the transposed matrix tx satisfies WxII ::; nlixli.
Solutions to the Exercises From this claim (taking B(H) it follows that IITnllcb ::; n.
= M m, m
~ 1
419
and using JlJm(J1fn) c:= Mn(Mm))
To check this claim, we identify Mn(B(H)) with Mn &JB(H), and we write
lI'xll
~ IlL eu I " (~ t, 0
Xji
t, f'
~ (~
XjiX;i
Hence, since II Tnlc:" : en
---+
0
eij
Xji
"Vii ,up
'f
Y
XjiX;i
'f'
Rnllcb ::; ,jii, we find 1/2
Wxll ::;
yn. ynsup
::; nllxll,
i
which proves our claim, and hence IITn Ilcb ::; n. To verify the equality, consider the element T = Lij eij &J eji E Mn &J Mn c:= Mn(lIfn). We have (IAI" &J Tn)(T) = Lij eij &J e;.j. Viewing T as an element of B(£'2 &J2 (2)' it is easy to verify that T(ej &J ei) = ei &J ej, hence IITIIMn(JIl,,) = IITIIB(f202f2) = 1. On the other hand, a similar analysis shows that n -1 L eij &J eij is the orthogonal projection onto n- 1/ 2 L ej &J ej on the space £2 &J2 £2; hence n
L
Thus we conclude ~hat
=n.
e;j &J eij
ij=1
Mn(Al,,)
IITnlicb ~ n.
Exercise 1.3. This follows immediately from the first part of Remark 1.13. Exercise 1.4. Assume F C B(K). Let (Ka)aEI be a directed net of finitedimensional subspaces of K such that UKa is dense in K. Then, for any Y in F, we have obviously
IlylI = sup aEI IIPK,.YIK" II· Let deC\!) = dim(Ka )· Using an orthonormal basis in K a , we may identify B(Ka) with Md(a)· Let Va: F ---+ lIfd(a) be the map defined by va(y) PK"YIK". Clearly Ilvalicb ::; 1. Moreover, we have
lIylI
= sup Ilva(Y)II· aEI
Introduction to Operator Space Theory
420
More generally (since we have
Ua C~(KoJ
is dense in C~(K)), for any (Yij) in Mn(F)
Hence, for any (Xij) in the unit ball of Mn(E), we can write
which implies lIulicb :$ SUPoEI IIvaullcb. The announced equality is now obvious (recalling IIvullcb :$ Ilvllcbllullcb). Exercise 1.5. (i) Let H = C~ ®2 K = C~(K) and let V: C~ ---+ H (resp. W: C~ ---+ H) be the linear map defined by Vei = ei ®x; (resp. Wej· = ej ®Yj). Let rr: Mn ---+ ]\In ®B(K) be the representation defined by rr(a) = a®I. Then (if, say, our scalar product is linear in the second variable) we have
V*rr(a)W = LaijV*rr(ei.i)W = Laijeij(xi,YV Hence u(·)
= V*rr(·)W, which implies lIullcb:$ 1IVIIIIWil :$ 1.
Moreover, if x
= y,
then V
= H' and u is completely positive.
(ii) The generalization is obvious. Replacing C2 (S) and C2 (T) by their direct sum, we are reduced to the case S = T. We then define H = C2(T)®2K and V, W: C2(T) ---+ H by Ve s = es ® x s , Wet = et ® Yt. and rr: B( C2(T)) ---+ B(H) by rr(a) = a ® I. Then the same identity holds and yields lIulicb :$ 1.
Exercise 2.1.1. Assume E
c B(H), Ilxll
Fe B(K). We can write
= sup{l(xs, t)I},
where s, t run over the unit ball of H ®2 K. By density, we may clearly restrict the sup to s, tin H ®K. Then, for some finite-dimensional subspace Hn C H, we have s, t E Hn ® K. Hence, with v associated to Hn as above,
this yields
and since the converse is obvious, we obtain (2.1.2).
Solutions to the Exercises
421
When considering a directed net of subspaces, we note that, if HOt C Hj3 , we have IlvOt(e)11 ::; Ilvj3(e)11 and III:VOt(ai) 0 bill::; IIL:Vj3(ai) 0 bill; thus the preceding argument leads to the two announced equalities. Exercise 2.1.2. By the preceding exercise we have Ilu(x)11 = SUPOt Ilveru(x) II for any x, and hence Ilull = sUPo Ilveruli. Applying this. for each n, to hI" 0 u: A!n(F) ---+ lIln(E) (with hI" 0 Ver in place of ver) we obtain 11111/" 0ull = sUPer 11111/" 0 verull, which implies Ilulicb = sUPo Ilvoulicb. Exercise 2.1.3. The equality IITII = max{IITlll, IIT2 11} is immediate from the definition of T. Then, for any x in lIln(E), we have
(I 0 u)(x)
~
(I 0 Ul)(X) EB (I 0 U2)(X)
(using£20H = [f20H1]EB[£20H2]). Hl:'nce 1110u(x)11 = max{II(I0Ui)(X)111 i = 1, 2}, whence lIullcb = max{llulllcb, Ilu21Icb}.
Exercise 2.1.4. Let bi = Xi 01 and ai = 10 Yi, so that biai = aibi = Xi 0 Yi· The first inequality then follows immediately from (1.11). Exchanging the roles of ai, bi we get the second one. The equality cases follow from (1.10). Exercise 2.2.1. Let £oo(S, IR) denote the space all bounded real-valued functions on S with its usual norm. In £oo(S, IR), the set F is disjoint from the set C_ = {cp E foo(S,IR) I supcp < o}. Hence, by the Hahn-Banach Theorem (we separate the convex set F and the convex open set C_), there is a nonzero ~ E £oo(S, IR)* such that ~(f) ~ 0 V f E F and ~(f) ::; 0 V f E C_. Let 1Il C £oo(S, IR)* be the cone of all finitely supported (nonnegative) measures on S viewed as fllnctionals on toc(S, IR). Since we have ~(f) ::; 0 V f E C_, ~ must be in the bipolar of fl.! for thl:' duality of the pair (£oo(S, IR), £oo(S, IR)*). Therefore, by the bipolar theorem, ~ is the limit for the topology a(£oo(S,IR)*,£oo(S,IR)) of a net of finitely supported (nonnegative) measures ~Ot on S. We have for any f in £oo(S, IR), ~er(f) ---+ ~(f), and this holds in particular if f = 1; thus (since ~ is nonzero) we may assume ~0(1) > O. Hence, if we set >"o(f) = ~er(f)/~er(1), we obtain the announced result. Exercise 2.2.2. First observe that, by the arithmetic/geometric mean inequality, we have for any a, b ~ 0
(ab)1/2 In particular we have
= inf{2- 1 (ta+ (b/t))). t>O
Introduction to Operator Space Theory
422
Let Si be the set of states on Bi (i inequality implies
IL~(X{,X~)I
:::; Tl
=
1,2) and let S
{h (Lx{x{*)
sup f=(h,h)ES
= 81
X
8 2 • The last
+ h (Lx~*x~)},
Moreover, since the right side does not change if we replace xJ by ZjxJ with Zj E
= LTI h(x{x{*) + Tl h(~*x~) -1~(x{,~)I. j
By Exercise 2.2.1, there is a net U of probability measures (>'o,) on 8 such that, for any F in F, we have
l~
J
F(gl,g2)d>'o(gI,g2) 2::
o.
We may as well assume that U is an ultrafilter. Then, if we~'
Ii = l~
J
gi d>'o(gl,g2) E 8 i
(in the weak-* topology (1(B;, B i )), we find that for any choice of (x{) and (x~) we have LT1h(x{x{*) + 2-1h(x~*~) -1~(x{,x~)I2:: 0; j
hence, in particular, \iXl E FI, \ix2 E F2
By the homogeneity of ~, this implies inf{Tl(th(Xlxi)
t>O
+ h(X2X2)/t)} 2::
1~(Xl,X2)1,
and hence we obtain the desired conclusion using our initial observation on the geometric/arithmetic mean inequality.
Exercise 2.2.3. The proof of this is exactly the same as the preceding one, except that F should now be defined as formed of all possible functions F: 8 --+ R. of the form F(h,h) = LTI h(x-{x-{*) + 2- 1h(~*x~) -1'¢(x-{,Xj,~)I, j
where (x j) is allowed to be an arbitrary family in the unit ball of (G, 0:).
Solutions to tIle Exercises
423
Exercise 2.3.1. Let G be an arbitrary operator space (for instance, G = 1IJN ). Let x = L: a; ® bi E G ® E, and let u: E* ---+ G be associated to x. Since (by definition) IIxIIG0min E" = lIullcb, it suffices to show that IIxllG0miu E = lIulicb. By (2.1.6) we have IIxllG0minE =
{IlL
sup
n,t'EB"
v(b i ) ® a i "
Ill" (C)
} ,
where Bn = {v: E ---+ lIfn 1 IIvllcb :::; I}. By definition of E*, Bn can be identified with the unit ball of lIfn(E*), and if y E lIfn(E*) is the element corresponding to v: E ---+ A/n, we have
Hence, we obtain as announced IIxllC0minE = sup{III ® u(y)lIllIyIlJlJ,,(E') :::; I} = lIulicb. n
This completes the proof.
Exercise 2.3.2. We have lIu*lIcb = sup{II(I ® U*)(y)IIM,,(E') 1 lIyIlJlJ,,(p) :::; 1,n 2 I}. But each y in the unit ball of Afn(F*) is in one-to-one correspondence with a map Vy: F ---+ Mn with IIvyllcb = lIyIlJlJ,,(F') :::; 1. Moreover, in the identification lIfn(E*) ~ CB(E, M n ), (I ® u*)(y) corresponds to the composition vyu, so that 11(1 ® u*)(y)1I = IIvyulicb. Therefore we have III ® u*: Mn(F*)
---+
Mn(E*)1I
= sup{lIvulicb
IlIv: F
---+
Mnllcb:::; I}.
By Exercise 1.4, if we take the supremum of both sides over all n 2 1, we obtain IIu*IIcb = IIulicb. Exercise 2.3.3. The key to this variant is Proposition 1.12. Let us denote lIuli n = IIIJlJ n ®UllJlJ,,(E)-+JIJ,,(Fdl. By Proposition 1.12, each y in the unit ball of lI/n (F*) is in one-to-one correspondence with a map Vy: F ---+ 1I[n with IIvyiln = IIyllJlJ,,(p,) :::; 1. Thus, as in the preceding solution, we have
and we obtain IIu*IIn:::; IIuli n . By iteration, we find II(u*)*IIn :::; IIu*IIn. But, by Exercise 2.3.1 we have
hence we also obtain IIuli n :5 IIu*IIn.
Exercise 2.3.4. The "only if" part is essentially obvious using (2.3.3) for u and u- 1 , so we turn to the "if" part. Assume that u*: F* ---+ E* is a complete
Introduction to Operator Space Theory
424
isomorphism. Then u**: E** ---+ F** is a complete isomorphism by the first part of the solution. By restricting u** to E, we see that u: E ---+ F must be at least a completely isomorphic embedding; but, since u* is injective, we know u must have a dense range, so actually u must be onto F. Then we conclude that u = ujg: E ---+ F is a complete isomorphism. The completely isometric case follows from (2.3.3).
Exercise 2.3.5. It suffices to show that Ix: Q9 u: IC Q9min R* ---+ IC Q9min C is isometric. Let (Xi) be a finitely supported sequence in /C. We claim that
Let v: R we have
IILxi
---+
IC be the map defined by v(·)
=
LXi~i(')' By definition of R*
Q9~illmin = Ilvllcb = IIIx: Q9vll = sup{II(Ix: Q9 v)(Y)lk®lIlirl'~: lyE IC Q9 R, lIyllK:®mi R-:::; I}. H
By density, we may restrict the preceding supremum to those Y for which only finitely many Yi are nonzero. Let Y = LYi Q9 eli. Then by (1.10) we have lIyll = IILYiY;1I 1/ 2 and (h:Q9v)(y) = LYiQ9v(eli) = LYiQ9Xi. Thus we find
By Exercise 2.1.4, this last supremum is = IILXixiI11/2. Thus, by (1.10) again, we have
hence, since eil = U(~i)' we obtain our claim, which proves that u is a complete isometry. The other identities follow by exactly the same arguments.
Exercise 2.4.1. (ii) ::::} (i): Assume (ii). Note that IIIG Q9 v: G Q9min Fl ---+ G Q9min Ellcb :::; C by (2.1.4), and (IG Q9 u)(IG Q9 v) = IG Q9 1Ft , To show that IG Q9 u is surjective, it suffices to show that any X in G Q9 F with IIxllmin < 1 admits a lifting on G Q9 E with Ilxlimin < C. But this is clear since we can assume x E G Q9 Fl and set x = (IG Q9 v)(x). (i) ::::} (ii): By a routine argument (using direct sums) one can show that there is a constant C such that, for any G, Ia Q9 u is C-surjective, that is, any x in G Q9min F with IIxll :::; 1 can be lifted to an x in G Q9min E with Ilxll :::; C. Let Fl be as in (ii). Let x E Fi Q9 Fl be the tensor associated to the identity on Fl. Note Ilxlimin = 1. Let x E Fi Q9min E with IIxll :::; C be a lifting of x, and let v: Fl ---+ E be the linear map associated to X. Clearly (by (2.3.2)) IIxlimin = IIvllcb, and since x lifts x, uv is equal to the inclusion Fl C F.
x
Solutions to tIle Exercises
425
By the extension property of B(H), one shows easily that (ii) implies (iii). Finally, (iii) => (i) is easy by restricting first to the algebraic tensor product (for any t in G Q9 F with IItll min < 1, let Xi = (I Q9 Vi)t E G Q9 E; then Ilxillmin < C and (1 Q9 U)Xi --+ t, so that (i) follows by the open mapping theorem.) Exercise 2.4.2. Let Vn be as in the hint. By our assumption on u, there is a c.c. v: E --+ A such that
'Vi
= 1, ... , n + 1.
For a suitable choice of 0 (to be specified below) we will let
Note that, since
we have By Lemma 2.4.4, for any given x and c > 0, Q: can be chosen large enough so that IIV n +1(x) - [O"a:vn(x) + (1 - O"a:)v(x)lIl < cj hence Ilqvn +l (x) -qv(x)1I < c, which implies Ilqvn+1 (x) -u(x)11 < c + Ilqv(x)u(x)lI. So we can choose Q: large enough so that 'Vi
= 1, ... , n + 1.
Moreover, we have
hence, for
Q:
large enough, by (2.4.6), we can ensure that
IIVn+1(x)-vn(x)1I < c+llq[v(x)-vn(x)lll < c+llqv(x)-u(x)II+llqvn(x)-u(x)ll· Thus, if we now make this last choice of Q: valid for any x in {x 1, ... , Xn }, and if we take c = 2- n - 2 , we obtain the announced estimates for V n +l. It follows that vn(x) is Cauchy for any x in {x}, X2, . .. }. Therefore v(x) = limn vn(x) exists and satisfies qv = u.
426
Introduction to Operator Space Tl1eory
Exercise 2.6.1. We can identify Mn(t'l({Ei liE I})*) with CB(t'l({Ei I i E I}, Mn). By the definition of the t'l-direct sum, the latter space can be identified with t'oo( {CB(Ei' Mn) liE I}) or, equivalently, with
and by (2.6.1) this last space is the same as Mn (EBiEIE;). Thus we conclude
Exercise 2.6.2. The two sides of the equality are isometric Banach spaces. By a density argument it is easy to reduce this to the case when I is finite, so we will assume I finite. The universal property of the t'l-direct sum immediately implies that
is a complete contraction. To show that J is a complete iSOInettJ(it suffices to show the following claim: For any complete contraction u: t'l({E; liE I}) -> Mn we also have
lIu: CO({Ei liE I})*
->
lIInllcb::; 1.
But such a u defines an element in the unit ball of Mn(t'l({E; liE I})*), which, by the preceding exercise, coincides with the unit ball of !lIn (EBiEIEi*). But we have (by (2.6.1» !lIn (EBiEIEi*) = EBiEIMn(Ei*); hence (by (2.4.1» = EBiEI Mn(Ei)**, and hence (by elementary Banach space theory recalling I is finite) = (EBiEIMn(Ei»** or, equivalently, by (2.6.1) again, (lIIn (EBiEIEi»**, which by (2.4.1) again is the same as Mn (EBiEIEi)**)' Since this last space can be identified (see (2.3.1)') with cb (EBiEIEi)* ,Mn), we finally obtain (EBiEIEi)* -> 1, which is equivalent to the above claim (note that since I is finite, EBiEI Ei = co( {Ei liE I})) .
Ilu:
Mnllcb ::;
Exercise 2.6.3. This is an immediate consequence of the preceding two exercises. Exercise 2.13.1. For any x with IIxll = 1, there is an i such that IIX-Xill ::; £; hence lIuxll ~ lIuxdl - Ilu(x - xi)1I ~ 1 - £' - £.
By homogeneity, this implies lIu- 1 : u(E) -> Ell ::; (1_£_£,)-1 arid a fortiori d(E, u(E» ::; lIull . (1 - £ - £')-1 ::; (1 - £ - £')-1 .. Exercise 2.13.2. Choose £ so that (1-£)-1 < 1 +8. Let {Xi Ii::; N} be an £-net in the unit sphere of E. Let ei E BE· be such that ei(Xi) = 1, and let u: E -> t'~ be the mapping defined by u(x) = ei(x)ei. Let E = u(E).
Ei"
427
Solutions to tIle Exercises Then u satisfies the assumption of the preceding exercise with d(E,E) ::; (1- E)-I::; 1 + 6.
= O. Hence
E'
Exercise 2.13.3. Consider iI, ... , In in A. Let E = {iI,"', In}. Fix E > O. Let {Uj I j ::; N} be a finite open covering of J( such that the oscillation of each J; on each Uj is < E. Let {'Pj} be a partition of unity subordinate to {Uj }, that is, 'Pj is continuous, SIlPP 'Pj C Uj , 'Pj 2: 0 and L'Pj == 1. For each j, we choose a point tj in Uj . Then we define v: A ----> et/o by v(.f) = L~ I(tj)ej and w: et/o ----> A by w(ej) = 'Pj. Clearly IIvll ::; 1 and IIwll ::; 1. Moreover, for any i = 1, ... , n, we have lIu(.fi) - Ii I = IIwv(.fi) - Ii II = SUPtEK
ILj 'Pj(t)(.fi(tj) - Ii(t))1 ::;
u associated to E and
E
E. Let a = (E,E) and let u'" be the map as above. Clearly this yields the desired net.
Exercise 3.1. First consider the special case when G v
v
v
=
et/o. Then E 0 G
=
.
et/o(E) and (E 0 G)** = et/o(E**) = E** 0 G. Recall that, for any subspace Y C X of a Banach space X, we have Y** C X**; more precisely, Y** can be identified with the a(X**, X*)-closure of Y in X**. In particular, this shows v -
v -
-
(E 0 G)** = E** 0 G for any subspace G c et/o. For the general case we use the fact that, for any E > 0, there is an integer N anj a subspace G c e~ that is (1 + E)-isomorphic to G (i.e., such that d(G, G) ::; 1 + E). Using G instead of G and applying the first part, we get the announced result up to E, and letting E ----> 0 we obtain the general case. (See e.g.[DJT, p. 178] for v
v
more details). Note that E** 0 G = B(E*, G) isometrically and E 0 G corresponds to the subspace of B(E*, G) formed of all the weak-* continuous v
maps. In particular, if X = E 0 G, any element in B x ** is the a(X**, X*)limit of a net in Bx. Therefore, for any v: E* ----> G with IIvll = 1, there is a net of weak-* continuous maps v",: E* ----> G with II v" I ::; 1 such that, for any ~ in E*, v",(~) ----> v(~) in the weak (= norm) topology of G.
Exercise 3.2. Let F be an arbitrary operator space. Consider:r = F 0 E*. Let u: E ----> F be the associated map. Then
IIxIlF®m,umax(E)* =
lIuIlCB(max(E).F) = lIullB(E.F) =
L
bi0ai E
IIxll F®E* v
This proves that (max(E))* = min(E*). To prove the other equality, we will use the principle of local reflexivity (see Exercise 3.1). For any x in F ®min max(E*) with associated map u: E ----> F, we have, by (2.1.6), IIxIlF®m,umax(E*) = sup IIIF 0 v(x)ll, where the supremum runs over all n and v: E* ----> Mn with IIvll ::; 1. By Exercise 3.1 we may restrict to v weak-* continuous. But then v represents an element t E Mn 159 E with
428
Introduction to Operator Space Theory
IItllllf.,l8imln min(E) = IIvll :::; 1. Thus we obtain IIxIIFl8imin max(E*) u)(t)1I with the sup over all such t. Equivalently, IIxIIFl8iminmax(E*)
= sup II(IM.,
0
= lIuIlCB(min(E),F),
which means (take F = Mn) that max(E*) = (min(E))*. Exercise 3.3. By the preceding exercise, if E is minimal (resp. maximal), then E** is also. Conversely, if E** is minimal, then, by Exercise 2.3.1, E (being a subspace of E**) is itself minimal. If E** is maximal, then, for any F and any u : E -4 F, we have (by (2.3.3)) lIulicb = lIu**lIcb = lIu**1I (there is equality because E** is maximal), and of course lIu**1I = lIull; thus we obtain lIulicb = lIuli and we conclude that E is maximal (indeed taking for u the identity map from E to max(E), we find lIulicb = 1). Exercise 3.4. Let Ei be a family of Banach spaces, and let E = EBiEI Ei be the direct sum in the too-sense. It is easy to check that, for any finitedimensional normed space G, we have isometrically: •
Now assume that {Ei liE J} is a family of operator spaces. Consider x in Mn (EBiEI Ei) and let {Xi liE I} be the corresponding family with Xi E Mn(Ei) for each i. By (2.6.2) we have
IIxll
= sup IIXiIlM.,(E;)' iEI
But now, if each Ei is assumed minimal, we have
hence, applying the preceding observation with G = M n , we find
Thus we conclude that E
= min(E).
Exercise 3.5. It suffices to show that, for any u: £1 ({ Ei liE J}) -4 B(H), we have lIulicb = lIuli. We can assume that u corresponds to a family {Ui liE I} with Ui: Ei -4 B(H). Then obviously lIuli = sup IIUill, and, by definition of the o.s.s. on £1({Ei liE I}), we also have lIulicb = SUpllUilicb. But, since each Ei is maximal, we have IIUili = IIUilicb for each i, and hence we obtain lIulicb = lIull·
Solutions to the Exercises
429
Exercise 3.6. The operators Ui ® Ui and Ui ® Ui are commuting self-adjoint unitaries. In particular, they generate a commutative C*-algebra and therefore are minimal (recall Proposition 1.lO(ii)). Exercise 3.7. The first question is treated in §9.3, to which we refer the reader. Let T = Ck ® Ck. For any subset 0: C {I, ... , n} with 0: = {ii, ... , id with il < i2 ... < ik we denote e a = Pit /\ ... /\ eiA.' and e", denotes the vacuum vector. Let Wd = Llal=d ea ® ea. Since Cke a = 0, if k E 0:, we have (Ci ® Ci)(Wd) = Llal=d+l.iEa e a ® e a ; hence T(Wd) = (d + I)Wd+l' We
L7
have clearly
Ilwdll
1/2 = ( ~) , and therefore we find
IITII~(d+l) (
d:l
)
1/2 (
~
) -1/2
;
hence
Finally, we choose if n is odd: d = (n - 1)/2, and if n is even: d = n/2.
Exercise 3.8. By Theorem 3.1, it is easy to show that for any separable Y1 C X there is a separable subspace Y2 with Y1 C Y2 C X such that for any u: Y2 --+ B(H) we have IlulY,llcb ~ Iluli. (Consider a dense set in the open unit ball of Mn(Yd and factorize each of them as in Theorem 3.1. Then, any space Y 2 containing all the diagonal entries appearing in these factorizations for all possible n has the desired property.) Now, by induction, we can build a sequence X = Y1 C Y2 C ... C Yn C ... X of separable subspaces such t.hat. for any u: Yn + 1 --+ B(H) we have lIuw.. llcb ~ lIull; then. taking X 2 = UYn , we obtain for any u: X 2 --+ B(H), Ilulicb = lIull, which means that X 2 is maximal, and of course it is separable. Exercise 5.1. Consider ~ in the unit. ball of (EI ®h E2)*' By Corollary 5.4, there are a Hilbert space H and complete contractions 0"1: El --+ B(H, C) and 0"2: E2 --+ B(C, H) such t.hat ~(Xl ® X2) = 0"1 (Xt}0"2(X2). For simplicity of not.ation we will assume H separable. Let Pn : H --+ H be the orthogonal projection onto the span of the first n vectors of a fixed orthonormal basis in H. Then let ~n(XI ® X2) = 0"1(Xt}Pn0"2(X2). Note that ~n(XI ® X2) --+ ~(XI ® X2) (Xi E E i , i = 1,2) when n --+ 00. l\loreover, we have ~n E Ei ®h E2 and II~nllh ~ 11~llh ~ 1. In addition, for any scalar sequence (ej) with lejl = 1 we have IIL~ ej(Pj - Pj-dll ::; 1; hence sUPn
IIL;=1 ej(~j - ~j-l)ll. • ::; 1. E ®I.E 1
2
If Ei @h E2 1> Co, then en = 6 + L~ ~j - ~j-l must norm-converge in Ei ®h E 2· Indeed, otherwise we find t5 > 0 and an increasing integer sequence
Introduction to Operator Space Theory
430
such that II{n; -{n;-lll > 8 for all j. But then, setting Xj = {n; -{n;-l' we have Ilxjll > 8 and sUPn,!c;!=1I1Ecjxjll < 00, which implies (see [LT1, p. 22]) that Ei Q9h E2 contains Co isomorphic ally. Thus, {j norm-converges to { in Ei Q9h E 2, which shows that { actually belongs to Ei Q9h E 2· nj
Exercise 5.2. We will show that C n Q9h Ck ~ Cnk' Let B = B(€2). It suffices to show that, for any Y in B Q9 C n Q9 Ck, we have I = II, where 1= lIyIIB0min(C"0,,Ck) and II = lIyIIB0m,nCnk' Assume Y = Eij Yij Qg e il Qgej1. Then I I =
II Eij YijYij 111/2.
On the other hand, by Corollary 5.9, I is equal to
11E z; Zj 111/2 over all possible ways to write Yij = Clearly we have liE YijYij II = IIEj z; (E i yiYi) Zjll, and
the infimum of liE YiYi11 1/ 2
YiZj with Yi, Zj in B. hence II :::; I. Conversely, we can write
eu
Q9
Yij
= (eli
Q9
I)
(~ek1
Q9
Yk j )
hence, with the obvious identifications, this gives us Yi.j
and
ilL
j
= YiZj with
1/2 zjZj II 1/2
= LYkjYkj
= II.
kj
Thus we obtain I:::; II. This proves that C n Q9h Ck ~ C nk completely isometrically. The fact that Rn Q9h Rk ~ Rnk can either be proved similarly or be deduced from the preceding by duality using (5.14). This completes the proof of (5.16), or equivalently, of (5.16)' when H, K are both finite-dimensional. But then the general case follows using finite-dimensional approximations. We skip the routine details.
Exercise 5.3. Assume E C B(H). We first show that (5.17) is isometric. Consider x in Mn(E). Then, by Theorem 5.1,
where the sup runs over all O'i with 1I00ilicb :::; 1. Let bi = 0'1 (eil), aj = 0'3(e1j)' Note that (see Remark 1.13) IIEbi bil1 1/ 2 :::; 1I00tlicb IIEeileilll1/2 :::; 1 and similarly liE ajaj 1/2 :::; 1. Hence, by Remark 1.13, we have
II
IILeil
®Xij
®e1jllh:::; IIxIIMn(E)'
Solutions to the Exercises
431
Conversely, we have
where the sup runs over hj,ki E H with L IIhj l1 2 ~ 1, L Ilki l1 2 ~ 1. Then, if we define 0"3(elj) = (hj)e E B(CC, H) and O"I(eit} = (k;}r E B(H,q, we have (see Exercise 1.1) 1I0011Ieb ~ 1, 1I0"311eb ~ 1, and (xjjh j , k i ) = 0"1(eit}xij0"3(elj); hence we obtain the announced isometric equality (5.17). Now, knowing that (5.17) is isometric for any E, we will show that it must be completely isomet.ric. Indeed, for any k ~ 1 we have isometrically
hence, by associativity and by (5.16),
Since this holds for all k, we conclude that (5.17) is completely isometric. Then (5.18) is immediate by density. Finally, let us restrict the identification C®hE®hR ~ K:®min E to the subspaces spanned respectively by C®E®el1 and by ell ® E ® R. By the injectivity of both tensor products involved, we obtain C®hE ~ C®minE (resp. E®hR ~ E®minR) completely isometrically. IVlore generally, for any Hilbert space H we have He®hE ~ He®minE. Thus, taking E = Ke with K another Hilbert space, we find He®hKe ~ He®mirJ<e. But now, recalling He = B(CC, H) and Kc = B(CC, K), we find
Thus we obtain another proof of (5.16)'. Exercise 5.4. Fix c: > O. By definition of the completion, we have x = L~ Xk with Xk E El ® E2 such that L~ Ilxkllh < (1 + c:)lIxllh. Then, for each k, we can write Xk as a finite sum Xk = LiEAA a(k, i) ® b(k, i) with IILiEAAa(k, i)a(k, ~ Ilxkllh(l + c:) and IILiEAAb(k, i)*b(k, ~ IIXkllh(l +c:). Finally, we may order each set Ak in any way we wish, and the resulting series Lk LiEAk a(k, i) ® b(k, i) does the job.
i)*11
i)II
Exercise 5.5. The first and third isomorphisms were already proved in the text respectively in (5.18) and (5.23). To prove the second one, note that by (5.15) we have a completely isometric embedding R ®h E* ®h C --+ (C ®h E ®h R)*. Thus it actually suffices to show that R ® E* ® C is dense in (C ®h E ®h R)*. This can be verified in the following manner. Consider an element x in C ®h E ®h R. Let x(n) be the natural projection of x to
432
Introduction to Operator Space Theory
C n ®h E ®h R. We will denote Pnx
increasing sequence nl <
n2 .•.
=
x(n). It is easy to check that for any
we have
Now let cp be an element of (C ®h E ®h R)* and let cp(n) = P~cp. By duality we have (1Icp(udI1 2 + Ilcp(n2) - cp(nl)11 2 + ... )1/2::; Ilcpll. This inequality clearly implies that {cp( n) I n ~ I} must be a Cauchy sequence in (C®hC®hR)*. Hence cp(n) converges in (C®hE®hR)* to a limit that must be equal to cp since it coincides with it on the dense subspace Un C n ®h E ®h R. Thus we have Ilcp - P~cpll --+ O. But we may argue in exactly the same way for the projections Qm from C ®h E ®h R onto C ® E ®h Rm. This also gives us IIcp - Q~cpll --+ O. Hence IIcp - Q~P~cpll is arbitrarily small when nand mare suitably large. Since Q~R~cp E Rn ® E* ® Cm, this completes the solution. Exercise 5.6. By the preceding exercise we have (K ®min E)* ~ K* ®" E*, and by (4.4) (K* ®" E*)* ~ CB(K*,E**). Hence, if E is finite-dimensional, by (2.3.2) this is ~ K** ®min E.
If A is a C*-algebra, Exercise 5.5 together with Theorems 2.5.2 and 4.1 gives us (K ®min A)** ~ K**®A**. Exercise 5.7. Assume
Then, using (1.12), it is easy 2 to check that Ilxlltt ::; Ilxllh ::; lIall (E II X i11 ) 1/2 (E IIYjIl2)1/2. Assume conversely that Ilxlltt < 1. By Theorem 5.17 we may as well assume that either Ilxllh < 1 or IItxllh < 1. Since the two cases are similar, we will check only the first case, so we assume Ilxlih < 1. Then, by the very definition of II . IIh' we can write ell ® x = Xl 0 X 2 with Xl E Ml,n(Ed, X 2 E M n ,1(E2 ) such that IIXlll < 1, IIX211 < 1. By Theorem 3.1, we can write for some integer N x
=
EajjXi
Xl=O:o·D·O:l
®
Yj.
X2=(30·~·(3l'
and
where 0:0,0:1, (30, (31 are scalar matrices and D, ~ are diagonal matrices with entries in E l , E 2 , respectively, such that 110:011 IIDlIlIO:l II < 1
II(3ollll~lIlI(3lll
and
< 1.
Here the product 0:0' D (resp. ~. (3d is a line (resp. column) matrix of length N with coefficients (Xlo .. " XN) (resp. (Ylo ... , YN)) such that E IIxj 112 < 1 (resp. E IIYjl12 < 1). Thus, if we define aij = (0:1(30)ij, we obtain n
X
= Xl 0
X2
=
L ij=l
aijXi
® Yj,
433
Solutions to the Exercises and IlalllllN(E IIXiI12)1/2(E the proof.
IIYjI12)1/2 <
1. By homogeneity, this completes
Exercise 5.8. Consider x E A ® B ® ... ® A ® B (k times). We will show that IlxiITu- ::; (2k -1)2k- 1111P2k(X)II· Assume 111P2k(X) II ::; 1. By Corollary 5.3, it suffices to show that, for any family (aI, a 2, . .. ,a 2k ) of complete contractions into the same B(H), we have
By Theorem 1.6, we may as well assume that a j is extended to Ai:B (we view A and B as sitting inside A*B) and that it is of the form Vj1rj(·)Wj for some representation 1rj: A*B ---> B(Hj ) and with IIVjII = IIWjl1 ::; 1. lVIoreover, replacing each 1rj by the single representation 1r = 1r1 EE> ••• EB 1r2k and suitably modifying Vj and Wj, we may assume that, if x is either in A or B, we have aj(x) = Vj1r(x)Wj . Let us denote by L(Vj) and R(Wj) the operator of left and right multiplication, respectively, by Vj and Wj , so that a j is equal to L(Vj)R(Wj)1r restricted to either A or B, depending on the parity of j. Thus we are reduced to showing II(L(VdR(Wd1r) ..... (L(V2k)R(W2k)1r)
(x)11 ::;
1.
After some obvious simplifications, it finally suffices to show that, for any contractions T 1, T 2, ... ,T2k-1, we have
We will now use our assumption that 111P2k(X)11 ::; 1. This implies that, for any pair 1r1, 1r2 of representations on the same Hilbert space, we have
Let U be an arbitrary unitary in B(H,,). Taking 1r1 = 1r and 1r2(-) = U1r(·)U*, we find 111r· L(U)1r· L(U*)1r· .... L(U)1r (x)11 ::; 1. Since this holds for any
1r,
we may apply this with
1r
replaced by
1r
EB 0 and
_ ( T .(1 - TT*)1/2) U-(1 _ T*T)1/2 T* (as in the proof of Lemma 5.14), and we obtain that for any T with we have 1I1r' L(T)1r. L(T*)1r . .... L(T)1r (x)11 ::; 1.
IITII ::; 1
434
Introduction to Operator Space Theory
Finally we will use the polarization formula to replace (T, T*, T, T*, . .. ) in the last inequality by (T1 , T2, T3 , •• •), with IITjl1 ::; 1. We define for Z = (Zj) E ']['2k-1 2k-1
T(z)
=
L (zjTj)':j . (2k - 1)-1,
j=1
where Cj = 1 if j is odd and Cj = * if j is even. Note that IIT(z)1I ::; 1, and if we denote by dz the normalized Haar measure on ']['2k-1, we have j[T(Z) ® T(z)* ® ... ® T(zV ® T(Z)]Z1Z2 ... Z2k-Idz .
= (2k -
1)2k- 1T 1 ® T2 ® ... ® T2k-f.
Therefore, a simple averaging argument gives us that (*) implies
and we conclude as announced that
Hence we have shown II tP 2klw2k II ::; (2k - 1)2k-l. Using operator coefficients instead of scalar ones, the same argument yields
This completes the case d
= 2k.
The case d
= 2k + 1 is similar.
Exercise 5.9. To show Ex ~ C n ) it suffices to show the following claim: For any all ... ,an in B(H) we have
II L ai ® xillmin ::; 1. We will show II L a;aill ::; 1. Note L ai ® Xi = Lik Wikai ® ei ® ek· By Corollary 5.9, II L ai ® Xi II min ::; 1 implies (assuming dim(H) = 00) that there are ai, (3k in the unit ball of B(H) such that Assume
hence ai = wikaj(3k for any k, which implies ai = ain-l Lk Wik(3k, or, equivalently, if we set iik = n- I / 2 (3k and 'Yi = n- 1 / 2 Lk Wikiik, we have ai = ani. But then this implies
Solutions to tIle Exercises
435
and since Iln-1wIIAl" :S 1, we find
and we conclude as announced II L ai ai II :S 1. In the converse direction we have a natural completely contractive map en ---+ e~ taking eil to ei; hence, if we set Xi = eil ® Lk Wikek, we can write II 2:ai®Xill B(H)0minE... :S II2:ai®Xill B(H)0"oi,,(C,,0/.f~) ' and by Exercise 5.3 II 2: ai ® X i ll B( H)0min (C" 0/, €~) = II2: ai ® X i ll B( H)0minC" 0mi"€~
=s~pll~ai®Wikeilll
. B(H)0m",C"
=
II~UiaiII1/2
Hence we obtain II Lai ® xillmin :S II Luiaill!/2. This completes the proof of the above claim. Thus we have proved Ex ~ en completely isometrically. The proof for Ey ~ Rn is essentially the same. The reader will note (ifthe entries of ware no longer assumed unimodular) that, if we let c! = Iln-l/2[w;kllIIAl" and C2 = SUPi.k IWik·l, then we find a completely isomorphic embedding u: en ---+ e~ ®h £~ taking ei! to Xi such that Exercise 5.10. To show that eij ---+ Zij defines a complete isometry from A1n to e~ ®h e~ ®h e~ it clearly suffices to show the following claim: For any au in B(H) (with say H = (2) we have
II2: ai j ®zijll min = II(aij)IIAI,,(B(H)). Assume II
LUi]
® zijllmin :S 1.
We will show that lI(aij)IIAI,,(B(H)) :S 1.
By Corollary 5.9, since II E Uij ® Zij II = II Lijk UijWikU'~jei ® ek ® ej II, our assumption II E aij ® zi] II :S 1 implies that we can write
where ai, (3klj are in the unit ball of B(H). Hence we have
Introduction to Operator Space Theory
436 which implies that
where D l , D 2 , D3 are the diagonal matrkes with entries respectively «(li), (13k), (-yj). Thus we obtain
Conversely, since we have complete contractions from C n to e~ taking eil to and from Rn to ~ taking eli to ei, we have a complete contraction from Cn ®h e~ ®h Rn to e~ ®h e~ ®h e~ taking eil ® ek ® elj to ei ® ek ® ej. Hence we have ei
L aijWikW~j ®
eil
® ek ® elj
ijk
But, by (5.17), Cn ®h e~ ®h Rn ::= Mn(e~) ::= Mn ®min e~, and hence the last norm is equal to sup "[aijWikW~jJijIIMn(B(H))' k
Now, since
IWikl
=
IW~jl
= 1, we have for each fixed k
hence we conclude
which completes the proof of our claim. Note that we have established the last assertion in the course of the proof since the matrices Wand W' clearly have the same properties as Wand W'. Exercise 6.1. Consider Ei as embedded into a C*-algebra Bi and each Bi as embedded into the free product B = Bl *B N. Then, since B ®min A is an operator algebra, its N -fold product map defines a complete contraction
*...
PN: (B ®min A) ®h'" ®h (B ®min A)
-4
B ®min A.
If we restrict PN to (El ®min A) ®h ... ®h (EN ®min A),
we obtain the first announced result, by Theorem 5.13.
Solutions to the Exercises
437
By embedding A into a unital operator algebra, we may assume that A is unital. Then we may restrict the preceding map to (EI ®rnin A) ®h (E2 ®min CI) ®h (E3 ®rnin A), and this yields the second assertion for the case N = 3.
Exercise 7.1. Let (T;}iEl be any orthonormal basis of OH(I). For any finitely supported family (a;};El in (say) B(£2) we have
IILai ®
2
T i l/ = =
IILai ®aill liLa; ®a;11 =
liLa; ®T;112,
and on the other hano
=
II(La; ®a;)*11 = liLa; ®a;11 = liLa; ®T;112.
Exercise 7.2. Let (T;}iEl be orthonormal in E
(); =
(~* ~;)
= OH(I).
Let
E lIh(E}.
Clearly, if E c B(1l}, then ()i E lIh(B(1l)} is self-adjoint. Using the preceding exercise it is easy to show that span[(); liE I] ~ E completely isometrically.
Exercise 7.3. If we apply the formula
IILiEJ T; ® Till satisfies t = t l / 2 ;
with a; = 1\, we obtain that the norm t = hence t = 0 or 1, and t = 0 is excluded if PI
=1=
o.
Exercise 7.4. Let U: E ........ E be a unitary operator, and let (Ti}iEl be any orthonormal basis of E. Let (); = U(T;). Clearly ()i is orthonormal; hence, by the proof of Theorem 7.1, we have
for any finitely supported family (ai) in K or B(£2)' This shows that U: E ........ E is a completely isometric isomorphism. In particular, IlUlicb = 11U1i. By the Russo-Dye Theorem, this remains true for any U in B(E), whence CB(E, E) ~ B(£2(I)) isometrically. However, the latter identity cannot be completely isometric because, if we fix e in the unit sphere of E and consider the subspace of CB(E, E) formed of all the maps of the form x ........ (e, x)h with h E E, then in CB(E, E) we
Introduction to Operator Space Theory
438
obtain E = OH(1) but in B(£2(1» we obtain £2(I)c (column Hilbert space), and it is easy to see that these are different operator spaces. Exercise 7.5. Let (TaJoEI be an orthonormal basis in E
= OH(1). Then
and hence
2:)Xij, Xke)eij ® ekl ijkl
2:
(Xij, Xke)eij ® ekl
ijkl
The last line is because lI(akl)IIJ'If,,(E) =
II(ake)IL'H,,(E)
for any E.
Exercise 7.6. Let (Tk) be an orthonormal basis in OH. Let hi Then
IlL
h; ®
= Lk hi(k)Tk'
x;II' ~ 11~?k ® ~ h;( k )x; II'
~ ~ ( ~ 14 (k)x; ) =
II
® (
Y
h;(k) x; )
~ (2: hi(k)hj(k») Xi ® Xj t,}
k
Exercise 7.7. We have IILeil ® Till = IILeil ® eillll/2. But L~ eil ® eil obviously has the same norm as L~ eil; hence IIL~ eil ® Till = IIL~ eilll l/2 = n 1/4. The other estimate is proved similarly. If we identify en with R~, then we find that I L eil ® Ti II is equal to the c. b. norm of the identity map j from Rn to OHn; hence this is = n 1/ 4 • We may then view v as the composition of a unitary operator on Rn (hence a complete isometry) followed by j. This yields IIvllcb = nl/4. The proof for u is similar.
Solutions to the Exercises
439
Exercise 7.S. Let E() = (E, EOP)(). Note that (by Theorem 2.7.4) (E())* = (E*)o. By Theorem 7.8, there is a factorization
with Ilvllcbllwllcb = ,Jii. Since E(O) ~ en ®min E and E(I) ~ en ®min EOP, we have E(fJ) ~ en ®min E() ~ eB(R n, E()). Hence, if (Xi) and (~i) are defined by Xi = w(eli) and v(·) = L~i(-)ei1' we find Ilwllcb = lI(xi)IIE«() and Ilvllcb = Ilv*llcb = 11(~i)IIE·«(). This gives us II(Xi)IIE«()II(~i)IIE'«() = ,Jii. In the case fJ = 1/2, Corollary 7.12 then yields
Xill~i: : ; II(xi)IIE(1/2) ilL ~i ® (ill~i~l : ; 1I(~i.)IIE·(1/2).
IILXi ®
Hence we recover dcb(E, OHn) ::; ,Jii as a by-product. Exercise 7.9. Let /1 be the harmonic measure of the point z = 1/2 in the strip S = {z Eel 0 < Re( z) < I}. Recall that IL is a probability measure on as such that f(I/2) = J f d/1 whenever f is a bounded harmonic function on S extended nontangentially to S. Obviously /1 can be written as /1 = 2-1(/10 + /1d, where /10 and /11 are probability measures supported respectively by
ao = {z I Re(z) = O}
and
a1 = {z I Re(z) = I}.
Let (Ao, Ad be a compatible pair of Banach spaces. We first need to describe (Ao, Ad1/2 as a quotient of a subspace of L 2(/10; Ao) EB L 2(/11; Ad. The classical argument for this is as follows. Let F(A o,A 1) be as defined in §2.7. We start by showing that, for any X in (Ao, A1h/2' we have
where the infimum runs over all f in F(A o, Ad such that f(I/2) = x. For a proof, see, e.g., [KPS, p. 224]. Then let E = L 2(/10;A o) EBoo L2(,L1,A1) and let FeE be the closure of the subspace {fIBII EB flBI I f E F(Ao, Ad}· The preceding equality shows that the mapping f ---+ f(I/2) defines a metric surjection Q: F ---+ (Ao, Ad1/2. We now consider the couple (Ao, AI) = (R, e), where we think of Rand e as operator space stuctures on the "same" underlying vector space, identified with £2. We introduce the operator space E = L2(/10; £2)r EB L 2(/11; £2)c. Let F and Q: F ---+ £2 be the same as before, so that if we assume f analytically extended inside S, we have Q(f) = f(I/2). We first claim that
IIQ: F
---+
OHlicb ::; 1.
440
Introduction to Operator Space Theory
To verify this, consider x in Mn(F) with Iix Ii Jlf" (F) :'S 1. We claim that Iix(1/2)lllIf n (OH) :'S 1. We may view x as a sequence (Xk) of .l\{n-valued functions on 8S extended analytically inside S, so that
Iixil U" (F)
~ max {II (J L x,x; d~o
rt", I (J L x,x, d~rlJ '
and by (7.3)'
Ilx(1/2)II~f,,(OH)
= II2:Xk(1/2) ® Xk(1/2)ILn = sup {Itr
(2: Xk(1/2) axk(1/2)*b) I} ,
where the supremum runs over all a, b ~ 0 in l\fn such that trlal 2 :'S 1 and trlW :'S 1. Fix a, b satisfying these conditions. Consider then the analytic function on S. Note that
F(1/2) = tr
(2: Xk(1/2) ax k(1/2)*b) = T1 (i, F dpo + fat F dP1) .
But for all z = it in 8 0 we have
F(it)
=
2: tr(b1-itxk(it)a2itxk( _it)*b 1- it ); k
hence, by Cauchy-Schwarz, for any z in 8 0
A similar verification shows that for any z in 8 1 we have
Thus we obtain by Cauchy-Schwarz
IF(1/2)1 = I! Fdpi :'S 2- 1
(!ao IFI dpo + !al IFI dP1 )
:'S 2- 1 {tr (b 2 !LXkXZ dPo) :'S IixIIMn (F) :'S 1, which proves our claim.
+tr (a 2 !LXZXk dPl ) }
Solutions to tIle Exercises
441
It is now easy to show that Q is actually a complete metric surjection, or, equivalently, that I ® Q: Mn(F) -+ l\;fn(OH) is a metric surjection for any n 2: 1. Indeed, consider x E Mn(OH) with I/xIIJlI,,(OH) < 1. Since Mn(OH) = (Mn(R), Mn(C)h/2 (isometrically) by Corollary 5.9, there is a bounded continuous analytic function J on S with values in Mn(R) + lIfn(C) such that 0:0 = sup{IIJ(z)I/M,,(R) I z E ao} < 1, 0:1 = sup{IIJ(z)IIJII,,(c) I z E ad < 1 and J(1/2) = x. Let us write J(z) = (!k(Z»k, where !k is an Mn-valued function on S. We have trivially
and
hence IIJIIM,,(F) < 1. Since clearly (I ® Q)(f) = x, this shows that I ® Q: lIfn(F) -+ Mn(OH) is a metric surjection. Thus we have completely isometrically 0 H ~ F 1ker( Q). Finally, since E ~ R EB C and FeE, this completes the solution. Exercise 8.1. Let G = IF 00' If C is separable, let {Uj I i 2: 1} be a dense sequence in its unitary group. Let 11': G -+ C be the unitary representation taking the i-th (free) generator gj to Uj. By the universal property of C*(G) (see (8.1)), 11' extends to a representation 7?: C*(G) -+ C, taking Ua(gi) to Ui. Clearly 7? has dense range; hence 7? is onto and C ~ C* (IF (0) 1I with 1= ker(7?). The proof of the nonseparable case is the same. Exercise 8.2. By Corollary 1.7, any complete contraction 17: E2 -+ B(Hv) is the restriction of a complete contraction El -+ B(Hv). Thus any representation C*(E2) -+ B(H) is the restriction of a representation C*(E1 ) -+ B(H). Similarly, any completely contractive morphism u: OA(E2 ) -+ B(H) extends completely contractively to OA(E1)' This shows that C*(E2) -+ C*(El) and OA(E2) -+ OA(E1 ) are completely isometric embeddings. The unital case is identical.
v:
Exercise 8.3. The representation 11' obviously has dense range; hence it is onto, and it is automatically a complete metric surjection (see Proposition 1.5). Since I/ullcb ::; 1, U defines a complete contraction
u:
OA(EJ)I ker(u)
-+
OA(EJ/ E2)'
Note that the composition El -+ OA(Et} -+ OA(Edl ker(u) is a complete contraction that vanishes on E 2 ; hence it defines a complete contraction
Introduction to Operator Space Theory
442
Ed E2 -+ OA(E1)1 ker(u), and hence it extends further to a completely contractive morphism w: OA(Ed E 2) -+ OA(Edl ker(u). But then it is easy to check that is the identity, so = u- 1 and is a complete isometry. The unital case is identical.
uw
w
u
Exercise 8.4. Let'x ='xc and U = Uc. (i) ::::} (ii): Since U ®'x -:::: 1 ®'x (by Proposition 8.1) for any x in qG), we have III:x(t)U(t) ® ,X(t) II = lII:x(t)'x(t)lI· Let T = E x(t)U(t) ® ,X(t). Assuming U(t) E B(H), pick h, k in the unit ball of H. We then have
hence lI:x(t)(U(t)h,k)(8t *i<"fa)l:::;
IITII,
and since (8t * fa, fa) -+ 1 for any t, we obtain IE x(t)(U(t)h, k)1 :::; IITII and hence IIEx(t)U(t)1I :::; IITII = IIEx(t)'x(t)ll. This establishes (i) ::::} (ii) since the converse inequality is trivial. (ii) ::::} (iii): Let E c G be any finite subset. If (ii) holds, since the trivial representation (identically equal to 1 on G) is included in the universal one, we have
lEI =
II~ U(t)11 = II~ 'x(t)ll,
and hence (iii) holds. (iii) ::::} (i): Let E c S be a finite subset containing the unit. Assuming (iii), there is a net (Yo) in the unit sphere of f2(G) such that IIIEI- 1 EtEE 'x(t)· Yal12 -+ 1. By the uniform convexity of f2(G) (see Exercise 20.2 and recall e E E) this implies 1I,X(t)Ya - Yal12 -+ 0 for any t in E. Since this holds for any finite subset E, we can rearrange the resulting nets to make sure that the last convergence holds for any t in S. Then, since S generates G, the same convergence holds for any t in G. This shows (using the proposed alternate definition of amenability) that (iii) implies (i). Exercise 11.1. Let p: B ®max A -+ (BII) ®max A denote the natural representation (obtained from B -+ BII after tensoring with the identity of A). Obviously, p vanishes on I ®max A. Hence, denoting by Q: B ®max A -+ (B ®max A)/(I ®max A) the quotient map, we have a factorization of p of the form p= pQ
B ®max A~(B ®max A)/(I ®max A)~(BII) ®max A, where
p is a
(contractive) representation.
Solutions to the Exercises
443
Consider any x in (B II) @ A of the form x = I: (3; @ a;. Let b; E B be any lifting of (3;. We define 1T(X) = Q (I: bi @ai). Clearly this definition is unambiguous (it depends only on x and not on the choice of representations). l\Ioreover, 1T is a *-homomorphism from (BII)@A to the quotient C*-algebra (B@maxA)/(I@maxA) (which can be realized inside some B(H)). Hence, by the definition of the max norm on (BII)@A, we must have 111T(X)11 ::; II:rllrnax. Moreover, p{1T(X)) = x for any x in BII@A, but, since (as we just saw) 1T is continuous, this remains true for all x in B II @rnax A. Therefore 1T must be injective, and, since it obviously has dense range, it defines an isomorphism from (BII)@maxA to (B@rnaxA)/(I@rnaxA). Then the identity p(1T(:Z:)) =:z: implies that p is the inverse of 1T, in particular, p must be injective, and therefore the kernel of Q coincides with the kernel of pQ = p, so the sequence is exact. To prove that last assertion, observe that the isometric isomorphism
1T: BII@max A
---->
(B
@max A)/(I@max A)
induces a linear isomorphism between (BII) @A and the image of B @ A in (B @max A)/(I @max A). Thus, for any x in (BII) @A with Ilxli max < 1, we can find in B @A with Ilxli max < 1 such that x E I @max A, or, equivalently, x E I @ A.
x
x
x
Exercise 11.2. This follows from the preceding exercise. Exercise 11.3. The completion of At @ A2 for the given norm is a C*algebra. Clearly at ;::: 0, a2 ;::: 0 imply at @ a2 ;::: 0; hence al ;::: 0, bt - at ;::: 0, and a2 ;::: 0 imply 0 ::; al @a2 ::; bl @a2 and hence Iial @a211 ::; Ilbt @a211· Then ajx*xal ::; IIxl12ajai gives us Ilajx*xal @a2a211 ::; IIxl1211aiat @a2a211 = IlxI12I1at@a2112, and hence Ilxal@a211 ::; Ilxllllal@a211. Applying this twice we obtain IIxat @ya211 ::; Ilxllllyllllal@a211; hence lI at@a211 2= Ilajat@a2a211::; lIaill IIa21111at @a211, which, after division, yields Iial @a211 ::; Ilatlllla211. Exercise 11.4. (i) Since 1T(A t @ A 2 ) is a *-algebra, both I( and 1(.1 are invariant subspaces and 1T(x)IK-L = 0 for any x in At @ A 2, whence the announced decomposition, and it clearly suffices to solve the problem for 1T(.) IK. (ii) Since we know by the preceding exercise that 1T(xc<@a2) and 1T(al @Y/J) are uniformly bounded nets, it suffices to show that they converge all a total subset of H = K, in particular on {1T(A t @A 2)H}, on which this convergence becomes obvious. vVe have
and
444
Introduction to Operator Space Theory
which shows that 7r1 and 7r2 actually are independent of the choice of {x a } and {y,a}. (iii) From the preceding two identities it is immediate that 7r1 and 7r2 are *-homomorphisms with commuting ranges.
Exercise 11.5 (i) Assume the given matrix positive. For any real t we have
hence P(t) = t 2 (py, y) + 2t(ax, y) + (qx, x) 2: 0, so the discriminant of P must be::; 0, and we obtain l(ax,y)1 2 ::; (py,y)(qx,x). Conversely, if this holds, P must keep the same sign, and it is 2: 0 for large t; hence we have P(1) 2: 0 and the given matrix is positive. (ii) is an obvious consequence of (i). (iii) Since u is c.p., the map U2 = h.I2 0 u: M 2(A) ~ M 2(B) is positive. If A is unital, we have II all ::; 1
~ (;* ~)
2: 0
~ (~~J) ~~~D 2: O.
Hence, since u(a*) = u(a)*, (ii) shows that I(u(a)x, y) 12 ::; (u(1 )x, x) (u(1 )y, y) and hence Ilu(a)1I2 ::; Ilu(1)1I2. Thus we obtain lIuli = lIu(1)1I. Similarly, 111M " 0 ull = IIhI,,§ u(1)11 ::; lIull, and hence Ilulicb = Ilull ~ Ilu(1)11· Let (aa) be as in (ii). Let A be the unitization of A and let ua : A ~ B be the map defined by ua(a) = u(a~j2aa~2). Then clearly Ua is c.p. and ua(a) ~ u(a) for any a in A. Hence Ilulicb ::; sUPa Ilualicb = sUPa Ilua (1)1I = sUPa Ilu(aa)ll. Thus we conclude lIulicb = lIull.
Exercise 11.6. For (i), it clearly suffices to show that, if 0'1: A1 ~ B(H) and 0'2: A2 ~ B(H) are representations with commuting ranges, there is an extension &1: Ai* ~ B(H) such that &1 and 0'2 still have commuting ranges. But this is clear: By the universal property of the bidual (as the universal enveloping von Neumann algebra), since 0'2(A 2)' is a von Neumann algebra, any representation 0'1: A1 ~ 0'2(A 2)' extends to a (normal) representation &1: Ai* ~ 7r2(A2)'. (ii) follows from (i) by iteration. (iii) Assume At, A2 unital for simplicity. Let 7ri: Ai ~ (A1 0 max A2)**' (i = 1,2) be the obvious inclusions (a1 ~ a1 0 1,a2 ~ 10 a2), and let 7fi : Ai* ~ (A1 0 max A2)** be their canonical extensions as normal representations. Obviously (7rt, 7r2) and therefore (7ft, 7f2) have commuting ranges; hence, if we let 7r = 7f1 . 7r2, we obtain a (contractive) representation
which clearly coincides with j when restricted to A1 0 A 2.
445
Sollltions to the Exercises
Exercise 11.7. (i) Consider tEA ® B, where B is an arbitrary C*-algebra. \Ve need to show Iltll max :s: IItllmin. By density, it clearly suffices to show this for any t in (Un An) ®B. In other words, we may assume t E An ®B for some Ct. But then, since An is nuclear, we have IItIlA®m",xB :s: IItIlA,,®m",x B = IItllmin. (ii) Let Ct run over the set of all the finite a-subalgebras of A. Let An = L oo (O,Ct,/1) and A = L oo (n,A,/1). Since An is finite-dimensional, it is nuclear; hence (i) shows that A is nuclear. By Gelfand's Theorem, any commutative von Neumann algebra can be identified with Loo(O, A, p) for some (O,A,p). (iii) If A is commutative, A** is a commutative von Neumann aIgebra; hence it is nuclear by (ii). Hence A** ®min B = A**®rnax B for any B. But since A ®rnax B c A** ®max B (isometrically) by the preceding exercise, we have A ®rnax B = A ®rnin B and A is nuclear. Exercise 11.8. Let A = AG and p = PG. For any finite subset E c G, we clearly have A(t)p(t)b e = be for any t; hence LtEE A(t)p(t)be = IElbe, and therefore lEI is
:s:
IlL IlL
A(t)p(t)11
IlL
:s:
tEE
:s:
A(t) ® p(t)11 C;(G)®m",xCf~(G)
tEE
A(t) ® U(t)11
tEE
C;(G)®nwxC*(G)
Hence, if II . IImin = II· IImax either on C~(G) ® C;(G) or on q(G) ® C*(G), we obtain
IEI:s:
IlL
A(t) ® U(t)11 . '
tEE
mill
and therefore, by Proposition 8.1, lEI = IILtEE A(t)II, so that G is amenable by Exercise 8.4. Since C;(G) :::' C~(G), this completes the solution. Exercise 12.1. Clearly u is c.p. when a = b. Otherwise, let v: A be the mapping defined by
( ) = (b.Tb*
v x
axb*
An elementary verification shows that
where
bxa* ) axa* .
-+
1I12 (A)
446
Introduction to Operator Space Theory
Clearly this shows that v is c.p.; hence by definition of the dec-norm (before (11.2)) we have where Vll(X) = bxb* and V22(X) = axa*. Hence we obtain
Applying this to the mapping x
---+
u(x)lIa//-l//b//- l we find
lIulidec ~ lIall IIbll· Exercise 12.2. Let a = (aI, a2, ... ,an), b = (b l , b2, ... ,b n ) be viewed as onerow matrices with entries in A. Then, for any x in !lIn, u(x) can be written as a matrix product:
u(x) = axb*. Thus it is clear that u is c.p. if a = b. When ai, bj are arbitrary, we may (again) introduce the mapping v: Mn ---+ M 2 (A) defined by
bxb* vex) = ( axb*
bxa* ) axa* .
Again we note
V(X)=t(~ ~)
t*,
where t = 21/2 (: : ) E .l\h2n(A), which shows (see Corollary 1.8) that v is c.p., so we obtain again lIulidec ~ max{lIvlllI, IIV2211} ~ max{lIbIl 2, lI a ll 2}
= max {IILbjbjll, IILaia:II}, and, by homogeneity, this yields
Exercise 13.1. Since we may restrict ourselves to families of finite-dimensional unitaries in (8.8), the restriction of a to the span of J and {U(9i) / i E J} is completely isometric. Therefore, {a)j1 is completely contractive, which implies, by Proposition 13.6, that II (a) -llicb ~ 1. Exercise 15.1. By Theorem 15.5, A** I8i m in C = A** I8i max C iff A** has the WEP. Since A** is a von Neumann algebra, this holds iff it is injective and hence, by the Connes-Choi-Effros Theorem (Theorem 11.6), iff A is nuclear.
Solutions to the Exercises
447
Exercise 15.2. It is obvious that any nuclear C*-algebra has this property by the "injectivity" of the minimal tensor product. Conversely, assume that A has the property under consideration. We will show that A ** is injective. We may assume A unital (see Remark 12.13). Let 0': A --+ B(H) be a unital embedding such that A** ~ O'(A)". We will apply the property under consideration to Bl = O'(A)' and B2 = B(H). Let 11': A Q9max Bl --+ B(H) be the unital C*-representation defined by 1I'(aQ9b) = O'(a)b. Since AQ9maxBl C A Q9max B(H), by Corollaries 1.7 and 1.8, 11' admits a completely positive and completely contractive extension T: A Q9max B(H) --+ B(H). Let us define
v x E B(H)
P(x) = T(1 Q9 x).
Note that P is the identity when restricted to B l . We claim that P (which is clearly completely contractive) is a projection from B(H) onto B l . Indeed, by Lemma 14.3, T must be bimodular with respect to A Q9 Bb so that for any a in A we have
T(1 Q9 x)O'(a) = T(a Q9 x) = O'(a)T(1 Q9 x); hence O'(a)P(x) = P(x)O'(a), so that P(x) E O'(A)' = B l . This proves our claim. Thus we conclude that O'(A)' is injective and, by the "well-known fact" O'(A)" = A**, must also be injective. Exercise 15.3. Let B be a WEP C* -algebra with an ideal I c B so that BII = M. Let q: B --+ BII be the quotient map, and let Bl = q-l(M) c B. Since M = Bd I, it clearly suffices to show that Bl is WEP, or, equivalently (by Theorem 15.5), it suffices to show that, if C = C*{lFoo), we have C Q9max Bl = C Q9min B l . For that purpose, consider t in C Q9 Bl with IItllmin < 1. Since B is WEP (by Theorem 15.5), we have Iltllc®lIlnx B < 1, hence II (I Q9 q)(t)IIC®mnxM < 1, and finally, by Proposition 11.8, II{I Q9 q)(t)llc®max M :::; 1. By Exercise 11.1, since ]I.[ = BdI, it follows that there is 0 in CQ9B l with II 011 max < 1 and such that (IQ9q)(O) = (IQ9q)(t). Therefore we have t = O+x with x E CQ9I. But now by the triangle inequality Ilxlllllin :::; Iltlllllin
+ 1I011lllin < 2,
and, by Corollary 11.9, we have C Q9l1lin -[ = C Q9l1lax I; hence IIxll lllElX = Ilxlllllin < 2. Finally, again by the triangle inequality, we conclude Iltll lllElX ::; 1I011lllax + Ilxll lllElX < 3. Thus we proved that the norms of C Q9I1lElX Bl and C Q9l1lin Bl are equivalent on C Q9 B l ; since these are C*-norms, they coincide. Exercise 15.4. By construction, B = t'oo{{M(n) 1 n ~ I}) is WEP if the algebras M(n) are all WEP. Hence M = Bjlu is QWEP. Since M has
448
Introduction to Operator Space Theory
a finite faithful trace, there is a completely positive contractive conditional expectation P from M onto !vI. Thus, by the preceding exercise !vI is QWEP.
Exercise 15.5. We first claim that for any finite subset E have (actually this follows from Theorem 8.2) lEI
c
G we always
= L A(t) ® A(t) I c~ (G)0m"xC~ (G)
tEE
Note that, for any x in C(G), we have
Hence the correspondence A(t) ---+ p(t) linearly extends to an isometric representation from C~(G) to C;(G). Therefore, for any finite subset E c G, we have
=
LA(t) ® A(t) tEE
L A(t) ® p(t)11 tEE
max
max
But now (see the solution to Exercise 11.8)
IlL
~
A(t) ® p(t)11
tEE
IlL A(t)p(t)11 tEE
max
~ lEI· B(t 2 (G»
Thus we obtain IIEtEE A(t) ® A(t)limax ~ lEI, which proves our claim (since the inverse inequality is trivial). Similarly, if we let 1If = VN(G), we have (by the same reasoning) lEI Theorem 15.6, if either
C~ (G)
lEI =
=
IIEtEE A(t) ® A(t)11
_. Therefore, by
1I10mRxlll
or 1\1 is WEP, we also have
IlL tEE
A(t) ® A(t)11
.' mm
and by Proposition 8.1 this implies lEI = IIEtEE A(t)ll. Then, by Exercise 8.4, we conclude that G is amenable.
Exercise 17.1. Let A C B(H) be a C*-algebra. We already know (by Theorem 17.1 and Theorem 15.5) that, if A is nuclear, then it is exact and' WEP. Conversely, by Proposition 15.3, if A has the WEP, then, for any C*algebra C, the natural morphism A
®max
C
--+
B(H)
®max
C
449
Solutions to the Exercises
is isometric. On the other hand, by Theorem 17.1, if A is exact, the same mapping is a contraction from A ®min C to B(H) ®max C. Therefore, for any t in A ® C we must have IItll max ::; IItllmin, so that we conclude A ®min C = A ®max C for any C, and hence A is nuclear. Exercise 18.1. Let X be any operator space and let E c X be a subspace. We will show that any complete contraction u: E -> A** admits a completely X -> A **. It clearly suffices to show contractive (c.c. in short) extension this when E is finite-dimensional. In that case, by the local reflexivity of A, there is a net of c.c. maps Ua: E -> A tending to u with respect to a(A**, A*). Let iA: A -> A** denote the canonical inclusion. Since A has the \VEP, iAlla admits a c.c. extension Va: X -> A**. Let U be any ultrafilter refining our net. We let u(x) = limu va(x) (with respect to cr(A**, A*)). Then u: X -> A** is a c.c. map, and for any e in E we have
u:
= limva(e) = lim l1a(e) = u(e).
u(e) Thus
u is the desired extension.
Exercise 18.2. Fix 0 < c: < (1 + A)-I. Let {Xj 11 ::; j ::; N} be an c:-net in the unit sphere of E, and let ej E E* be such that 1 = ej (x j) = II ej II (1 ::; j ::; N). Let a be large enough so that
We claim that 1 - (1
+ A)c < IIU~\,,,(E)II·
Indeed, for any x in the unit sphere of E, pick j such that then have
IIx -
xjll
< c:. We
hence, by homogeneity, 'r:/xE E
1111a(x)1I 2: (1 - (1
+ A)c:)lIxlI,
which proves our claim. Exercise 18.3. By construction we_have V(gxh) = gV(x)h for all g, h in G, from which it is easy to see that V has the required form. \Ve then have lIuli n = III®u: Mn(E) -> Mn(X) II = lLVII, and, by the convexity ofthe norm, denoting Vg,h(X) = V(gxh) we have IIVII ::; sup{IIVg,hll I g, hE G} = IIVII· Exercise 18.4. Consider u: E -> (X/Y)** with dim(E) < 00 and lIulicb ::; 1. Let V = ru: E -> X**. Since X is A-locally reflexive, there is a net Va: E -> X
Introduction to Operator Space Theory
450
suitably approximating v with IIv o ll c b ~ A. If we let U o = qvo.: E ---+ X/Y, we obtain a similar approximating net for u, showing that X/Y is A-locally reflexive. Now, if X is a C*-algebra and Y is a closed two-sided ideal, we have X** ~ y** $ (X/Y)**, and hence our assumption is valid in that case.
Exercise 18.5. Since C*(lFoo ) embeds in C*(1F 2 ), it suffices to show that the latter is not locally reflexive. But if it were, then, by Exercises 18.4 and 8.1, every separable C* -algebra would be locally reflexive. By Proposition 18.5, every C* -algebra would be locally reflexive, which is absurd (for instance, B(£2) is not locally reflexive because, by Exercise 18.1, its bidual would then be injective). Exercise 18.6. By (18.2), for any C*-algebra B and for any t in X** ® B** we have IIJ(t)II(X®minB)" ;::: IItllx**®min B", So we only need consider the reverse inequality. Assume Iltllx**®min B** ~ 1. By property C", there is a net (to.) in X ® B** with II tall min ~ A" that w*tends to t (here w* refers to the a«X ®min B)**, (X ®min B)*) topology). By property C ' , each to can be w*-approximated by a net (to.,{3) in X ® B with IIto ,{3l1min ~ N'N. Thus we conclude IItll(X®minB)** ~ NAil.
Exercise 18.7. Consider E (finite-dimensional) and u: E ---+ A** with lIuli cb ~ 1. We need to approximate u (pointwise in the a(A**, A*)-sense) by completely contractive maps from E to A. Replacing u by VoU: Jl.,fn(o) ---+ A**, we see that we may as well assume E = Jl.,fn for some n. But then, if A ** is injective and E = M n , by (11.5) and (12.7) we have E*®oA** = E*®minA** (isometrically). But now, by (12.5) and Exercise 11.6(iii), we find IIE* ®o A** ---+ (E*®oA)**1I ~ 1; hence a fortiori IIE*®minA** ---+ (E*®minA)**1I ~ 1, which shows that A is locally reflexive. Exercise 18.8. Recall A** first observe that (A**
®min
~
(A/I)** $J**. Let B be any C*-algebra. We
B**)/(J**
®min
= (A/I)**
B**)
®min
B**.
Since A has (C), we have IIA**
®min
B**
---4
(A
®min
B)**II ~ 1;
hence a fortiori IIA**
®min
B**
---4
«A/I)
®min
B)**II ~ 1.
But this last mapping clearly vanishes on J** ® B**; hence, passing to the quotient, we find II(A**
®min
B**)/(J**
®min
B**)
---+
«A/I)
®min
B)**II ~ 1,
but then, by our initial observation, this means that A/I has property (C).
Solutions to tIle Exercises
451
Exercise 18.9. Let ix: X ---+ X** be the canonical inclusion. Then (ix:)**: K** ---+ (K**)** is not equal to iJC ••• In short (iJC)"* =I- iJC •• ! Let B = K**. To show that A is locally reflexive we need to show that IliB ® I: B ®min A** ---+ (B®min A)**II ::; 1, but the "argument" only proves that II(iJC)** ®I: B®min A** ---+ (B®min A)**II ::; 1. Exercise 19.1. Consider s = 2: ai ® Xi in E ® B! and t = 2: bj ® Yj in F ® B2 with Ilsllmin < 1 and IItllmin < 1. We have by (2.3.5)
IL u(ai) ® Xiii, l
C B( F,G0u,;u Bl)
::; Ilulicb.
But note that
(v ® I B2 )(t) = il(s, t); hence we obtain Ilil(s, t)llmin ::; lIulicb. Exercise 19.2. We will apply (19.1) with ai = bi = ei (here {e;} is an orthonormal basis in £~) and with u equal to the identity on max( £~). Since dsdmin(£~)) = 1, this yields (using (10.23)) n ::; 4 dsdmax(£~))vn. Exercise 20.1. By (7.2) we have
hence, by Fell's absorption principle (Prop.lO.l),
which after squaring and dividing gives the desired inequality. Exercise 20.2. We have ~ 2:~ Ilxk - Af(x)11 2 = ~ 2:711xk112 - 111If(·7:)11 2, whence the first inequality. Therefore IllIf(x)11 2 > l-E implies ~(x) ::; 2.,fii£, and the last assertion follows. Exercise 21.1. Let p(x) = limm--+ooPm(x). Let (ei) be a basis of E. By our assumption, there is a constant C such that, for all m and all x = 2: Xiei E E, we have Pm(x) :-:; C2: IXil = Cllxlll. Let f! = {x E E Ilixll! = 2: IXil ::; I}. Note that IPm(x) - Pm(y)1 ::; Pm(x - y) ::; Cllx - YII! and Pm(O) = 0; hence, by Ascoli's Theorem, the sequence {Pm} is relatively compact in C(f!). Hence Pm must converge uniformly on f!, and, by homogeneity, Pm also converges uniformly over any bounded subset of E.
Introduction to Operator Space Theory
452
Exercise 21.2. Assume (i). Let (ei) be any basis in E. Let Um: E ---+ Em be an isomorphism such that lIumll = d(Em'E) and Ilu;;;,lll = 1. Let ei(m) = um(ei). Then (ii) is easy to check. Thus (i) => (ii). (ii)=>(ii)' is trivial and (ii)' =>(iii) is obvious. Assume (iii). Again let (ei) be any basis in E. Viewing ei as an element of TIEm/U, we may select a representative of its equivalent class in t'oo({Em}), denoted by (ei(m»m~l' We then obtain
Therefore we may argue as in the proof of Theorem 21.1 to show that the convergence (along U) I1I>iei(m)II ---+ IILx;eill is actually uniform over the unit ball of E. Then we obtain limu deEm, E) = 1, and, since this holds for any U, (i) follows immediately. Exercise 21.3. We will show (ii)' => (iii). Assume (ii)'. Then, for any N, let ef" E TIEm/U be associated to the sequence (ef" (m»m. For any (ai) in M N , we have II L aj 0 eilillIN(E) = II L ai 0 ef"IIl\JN(llEm/U)' Let (ft, ... ,ln) be a cluster point of the sequence {(ef, ... ,e;;) I N ~ I} in TIEm/U. Then we must have, for any N and for any (ai) in M N , II L ai 0 eillllIN(E) = II L ai 0 /i IIMN (nE m
/U)'
Hence the correspondence ei ---+ Ii is a completely isometric isomorphism from E to TIEm/U, so that (iii) holds. The other implications can be proved exactly as in the preceding exercise, so we omit the details. Exercise 21.4. Let Um: Em ---+ MN and u: E ---+ AfN be the operators associated respectively to L ai 0 ei(m) and L ai 0 ei' Recall (by (2.3.2»
Thus it suffices to show that Ilumlicb ---+ Ilulicb. For any b = (bi)i~n in (lIIN)n wedefinepm(b) = IILbi0ei(m)lImin andp(b) = IILbi0eillmin' Then, by Proposition 1.12, we have lIumllcb
= sup {IILai 0 bill I bE (MN)n,Pm(b) :::;
and lIulicb = sup
I},
{IIEai 0 bill IbE (MN)n,p(b) :::; I}.
By our assumption, Pm(b) ---+ pCb) for any bj hence, by Exercise 21.1, Pm(b) ---+ pCb) uniformly over the subset {b E (MN)n ,pCb) :::; I}. Let Cm = sup{IPm(b)p(b)11 pCb) :::; I}. We then have Cm ---+ 0 and (1 - cm)p(b) :::; Pm(b) :::; (1
+ cm)p(b).
Solutions to tlle Exercises Therefore, the preceding formulas for
lIu m li cb and
453
Ilulicb yield
and
Thus we conclude that Ilumlicb
-+
Ilulicb'
Exercise 21.5. Let F =span [XiJ. Let isomorphism. Our assumption implies
FC
MN and let v: F -+ F be an
hence limsupIILxi@ei(m)11 . ::; Ilv-IllcblimIILv(xi)@ei(m)11 . m~oo
min
nlln
::; IIv-IllcbIILv(Xi)@eillmin ::; Ilv-11lcbllvllcb IlL Xi @ ei Ilmin ' and similarly
Let
e=
dsdF). Taking the infimum ver all poss ible N and v we obtain
e- l IlL
Xi @ eill ::; liminf IlL Xi @ ei(m)11 ::; lim sup IILXi @ei(m)11 ::; e IILXi @eill.
This gives the first assertion, and, in particular, ife = 1, we obtain the second one. Exercise 25.1. Since U dilates T, we have for all x, y in K
In particular,
Hence, if a is isometric, we have
Introduction to Operator Space Theory
454
which forces U (~) = PH U (~). Therefore U (~) = (a~)' Clearly this implies (~) = U- I COx); so, if a is onto, we obtain the other equality. Exercise 25.2. This boils down to the observation that commuting pairs of complete contractions O'i: Ei -+ B(H) (i = 1,2) are in one-to-one correspondence with commuting pairs of completely contractive homomorphisms U( OA(Ei ) -+ B(H) (i = 1,2). Exercise 27.1. Let q: A -+ A/I be the quotient map. Consider Y in the open unit ball of Mn(A/I). Let x be a lifting of y in the open unit ball of Mn(A). If €(A) ~ d, we can factorize x as indicated in (27.10). Then, applying I ® q to each of D I , ... , D d , we obtain a similar factorization for y = (I ® q)(x). This shows that €(A/I) ~ d, and hence €(A/I) ~ €(A). Let us now assume €(A/ I) ~ d and €(I) ~ d. Consider x in the open unit ball of Mn(A). Let y = (I ® q)(x). Since we assume €(A/I) ~ d, for some constant K(A/I) we can factorize y as in (27.10), so that y = ooD I ... Ddo d with Di E J.IN(A/I) such that fIlloili ~ K(A/I) and IIDil1 < 1. Let ~i E MN(A) be liftings of Di such that lI~ill < 1, and let
Note that Ilx'll < K(A/I). Then (I®q)(x) = (I®q)(x') and therefore x-x' E Mn(I). Let x" = x-x'. By the triangle inequality, Ilx"111I1,,(I) < 1 +K(A/I); hence (since €(I) ~ d by assumption) we can factorize x" as in (27.10) with a constant K(I). This gives us IIx"lI(d) ~ K(I)(1 + K(A/ I)). Recall that, by (27.11), II . II(d) is subadditive. Thus we obtain for x = x' + x"
Ilxll(d)
~
Ilx'll(d) + IIx"lI(d)
~ K(A/I)
+ K(I)(1 + K(A/I)),
and hence we conclude that €(A) ~ d. Now assume that A is a C* -algebra so that I has a quasi-central approximate unit (0'0,) as in Lemma 2.4.4. Consider x in Mn(I) with IIxll < 1. If €(A) ~ d, then x can be factorized as x = ooD I ... DIOd as in (27.10) but with D I , ... , Dd diagonal matrices all with entries in A with· liD; II < 1, fI Iloill ~ K. We will use the following notation: For any 0' in A and any y = [Y;j] either in Mn(A) or in MN(A) we denote by 0' . Y the matrix [O'Y;j]. Then we observe that the properties of quasi-central approximate units (described before Lemma 2.4.4) guarantee that IIO'~ . x -
and
00(0'0 . DI)OI ... (0'0 .
IIx - O'~ . xII
-+
Dd)odll
-+
0
o.
But now O'o·DI, ... , O'o·Dd have their entries in I; hence (since IIx-O'~.xll(d) 0) we obtain €(I) ~ d.
-+
Solutions to the Exercises Finally, if A f(I) ::; f(A).
~
I x (AlI), of course we have I
Exercise 28.1. Let B
455 ~
A/(AII), and hence
= (B 1 , ••• , Bn) be an n-tuple in Mn. We define
V a,y,z E
f'2
We need to produce (B 1 , .•. , Bn) in Mn satisfying (28.19) and such that sup
!FB(a, y, z)1 ::; {3.
o,y,zED"
By a well-known fact (cf., e.g., [P8, pp. 49-50]) there is a subset An C Dn with cardinal IAnl ::; 52n such that sup
!FB(a, y, z)1 ::; 8
o,y,zED".
sup
!FB(a, y, z)l·
Ct,y,zEA".
The idea is to choose B 1 , ••• , Bn "at random" in 1.1n using Gaussian random matrices. Let {9k(i,j) I 1 ::; i,j,k < oo} be a collection of independent standard complex-valued Gaussian variables, that is, with mean zero and such that lEI9k(i,j)1 2 = 1. For any 1 ::; k ::; n, we define 9k as the n x n matrix with entries {n- 1 / 2 9k(i,j) 11 ::; i,j ::; n}. Then, for each a, y, z, the random variable n
X(a, y, z)
= L ak(9kY, z) k=1
has the same distribution as
Hence, assuming a, y, z all in D n , for any t > 0, we can write lP'{IX(a, y, z)1 > t}
= lP'{IY(a, y, z)1
> t} ::; P{ln- I / 291(1, 1)1> t}.
Therefore, there are positive numerical constants
/(1, /(2, ...
so that
456
Introduction to Operator Space Theory
which implies (recalllAnl ::; 5 2n )
IP'{
sup
IX(a, y, z)1 > t} ::;
[(15 6n exp( -[(2nt2)
::;
[(1
exp( -([(2t2
-
12)n).
Ct,y,zEA",
Thus, if we choose 0 so that (say) [(20 2 -
12 = 1,
the preceding estimate guarantees (by Borel-Cantelli) that
IX(a, y, z)1 ::; O.
limsup
sup
n-+oo
Q,y,zEAn.
Hence (by the known fact recalled above) lim sup
sup
n-+oo
Ct,y,zED n
IX(a,y,z)l::; 80.
On the other hand, we claim that
1~~~fn-lll~9k0gkt" ~ 1 Hence, we can certainly choose wand N(w) so that, if we set Bk = 9k(W), then for all n ~ N(w) we have (28.18) with (say) (3 = 80 + 1 and n/2 ::;
Ilk~1 Bk ® Bkll·
Replacing Bk by
v'2 Bk
we obtain the result as stated, and
it is a straightforward matter to adjust the constants to include the (finitely many) cases n < N(w). We will now prove the preceding claim. By definition of the min-norm, we have (see Prop. 2.9.1)
where the supremum runs over x, y in the unit ball of the n x n HilbertSchmidt matrices. Taking x and y both equal to n- 1 / 2 times the identity we obtain
But, by the strong law of large numbers, we have almost surely
n- 3
L
19k(i,j)1 2
---4
1,
l:5i,j,k:5n
and hence we obtain the announced result.
•
REFERENCES [AB] R. Archbold and C. Batty. C*-tensor norms and slice maps. J. London Math. Soc. 22 (1980) 127-138. [AI] A. Arias. Completely bounded isomorphisms of operator algebras. Pmc. Amer. Math. Soc. 124 (1996) 1091-110l. [A2] . A Hilbertian operator space without the OAP. Pmc. Amer. Math. Soc. 130 (2002) 2669-2677. AFJS] A. Arias, T. Figiel, W. Johnson, and G. Schechtman. Banach spaces which have the 2-summing property. Trans. Amer. Math. Soc. 347 (1995) 3835-3857. [AnI] C. Anantharaman-Delaroche. Classification des C* -algebres purement infinies nucleaires (d'apres E. Kirchberg). Seminaire Bourbaki, Vol. 1995/96. Asterisque No. 241 (1997), Exp. No. 805, 3, 7-27. [An2] C. Anantharaman-Delaroche. Amenability and exactness for dynamical systems and their C*-algebras. Preprint 2000 (http:j /arXiv.org/ pdf/math.OA/0005014). [AO] C. Akemann and P. Ostrand. Computing norms in group C* -algebras. Arner. J. Math. 98 (1976) 1015-1047. [AP] C. Anantharaman-Delaroche and C. Pop. Relative tensor products and infinite C*-algebras. Preprint, 1999. [APe] A. B. Aleksandrov and V. Peller. Hankel operators and similarity to a contraction. Internat. Math. Res. Notices no. 6 (1996) 263-275. [Arl] W. Arveson. Subalgebras of C*-algebras. Acta Math. 123 (1969) 141-224. Part II, Acta Math. 128 (1972), 271-308. [Ar2] . An Invitation to C* -Algebms. Springer-Verlag, 1976. [Ar3] . Ten Lectures on Opemtor Algebms. CBMS (Regional Conferences of the A.M.S.) 55, 1984. [Ar4] . Notes on extensions on C* -algebras. Duke Math. J. 44 (1977) 329-355. [AR] A. Arias and H. P. Rosenthal. !If -complete approximate identities in operator spaces. Studia Math. 141 (2000) 143-200. [Bl] D. Blecher. Tensor products of operator spaces II. Canadian J. Math. 44 (1992) 75-90. [B2] . The standard dual of an operator space. Pacific J. Math. 153 (1992) 15-30. [B3] . Generalizing Grothendieck's program. Function spaces, edited by K. Jarosz. Lecture Notes in Pure and Applied Math. vol. 136. Marcel Dekker, 1992. [B4] . A completely bounded characterization of operator algebras. Math. Ann. 303 (1995) 227-240.
458 [B5]
Introduction to Operator Space Theory Multipliers and dual operator algebras. J. Funct. Anal. 183 (2001) 498-525.
[Ba] S. Banach. Theorie des operations lineaires. Varsovie, 1932, Oeuvres de Stephan Banach, vol. 2, Acad. Polon. Sc., Editions scientifiques de Pologne, Varsovie, 1979. [BCL] K. Ball, E. Carlen, and E. Lieb. Sharp uniform convexity and smoothness inequalities for trace norms. Invent. Mat. 115 (1994) 463-482. [BD] F. Bonsall and J. Duncan. Complete Normed Algebras. SpringerVerlag, 1973. [Be] J. Bergh. On the relation between the two complex methods of interpolation. Indiana Univ. Math. J. 28 (1979) 775-777. [BeL] J. Bergh and J. Lofstrom. Interpolation Spaces. An Introduction. Springer-Verlag, 1976. [BF] M. Bozejko and G. Fendler. Herz-Schur multipliers and completely bounded multipliers of the Fourier algebra of a locally compact group. Boll. Unione Mat. Ital. (6) 3-A (1984) 297-302. [Bla1] B. Blackadar. Nonnuclear subalgebras of C*-algebras. J. Operator Theory 14 (1985) 347-350. [Bla2]
. K - Theory for Operator Algebras. Second edition. Mathematical Sciences Research Institute Publications, 5. Cambridge University Press, 1998.
[Blo] G. Blower. The Banach space B(f2) is primary. Bull. London Math. Soc. 22 (1990) 176-182. [BLM] D. Blecher and C. Le Merdy. On quotients of function algebras, and operator algebra structures on f p • J. Operator Theory 34 (1995) 315346. [BMP] D. Blecher, P. Muhly, and V. Paulsen. Categories of operator modules (Morita equivalence and projective modules). Mem. Amer. Math. Soc. 143 (2000), no. 681, viii+94 pp. [Bo1] M. Bozejko. Positive-definite kernels, length functions on groups and a noncommutative von Neumann inequality. Studia Math. 95 (1989) 107-118. [BoS1] M. Bozejko and R. Speicher. An example of a generalized Brownian motion. Commun. Math. Phys. 137 (1991) 519-531. [BoS2] . Completely positive maps on Coxeter groups, deformed commutation relations, and operator spaces. Math. Ann. 300 (1994) 97-120. [Bou1] J. Bourgain. Vector valued singular integrals and the Hl_BlIfO duality. Probability Theory and Harmonic Analysis, Chao-Woyczynski. pp. 1-19. Marcel Dekker, 1986.
References [Bou2] [Bou3] [Bou4] [Bou5] [BPI] [BP2]
[BR] [BRS] [BS] [Buel] [Buc2] [Bur] [BV]
[C] [Cal [CC1] [CC2] [CCFW] [CE1] [CE2]
459
Real isomorphic complex Banach spaces need not be complex isomorphic. Pmc. Amer. Math. Soc. 96 (1986) 221-226. . New Banach space properties of the disc algebra and H OO • Acta Math. 152 (1984) 1-48. . On the similarity problem for polynomially bounded operators on Hilbert space. Israel J. Math. 54 (1986) 227-24l. . New Classes of £P-Spaces. Lecture Notes in 1-1athematics, 889. Springer-Verlag, 1981. D. Blecher and V. Paulsen. Tensor products of operator spaces. J. Funct. Anal. 99 (1991) 262-292. . Explicit constructions of universal operator algebras and applictions to polynomial factorization. Pmc. Amer. Math. Soc. 112 (1991) 839-850. O. Bratelli and D. Robinson. Operator Algebras and Quantum Statistical Mechanics II. Springer-Verlag, 1981. D. Blecher, Z. J. Ruan, and A. Sinclair. A characterization of operator algebras. J. Funct. Anal. 89 (1990) 188-20l. D. Blecher and R. Smith. The dual of the Haagerup tensor product. J. London Math. Soc. 45 (1992) 126-144. A. Buchholz. Norm of convolution by operator-valued functions on free groups. Pmc. Amer. Math. Soc. 127 (1999) 1671-1682. A. Buchholz. Operator Khintchine inequality in non-commutative probability. Math. Ann. 319 (2001) 1-16. D. Burkholder. Distribution function inequalities for martingales. Ann. Pmb. 1 (1973) 19-42. H. Bercovici and D. Voiculescu. Free convolution of measures with unbounded support. Indiana Univ. Math. J. 42 (1993) 733-773. A. Calderon. Intermediate spaces and interpolation, the complex method. Studia Math. 24 (1964) 113-190. T. K. Carne. Not all H'-algebras are operator algebras. Pmc. Camb. Phil. Soc. 86 (1979) 243-249. J. F. Carlson and D. N. Clark. Projectivity and extensions of Hilbert modules over A(D N ). Michigan Math. J. 44 (1997) 365-373. . Cohomology and extensions of Hilbert modules. J. Funct. Anal. 128 (1995) 278-306. J. Carlson, D. Clark, C. Foias, and J. Williams. Projective Hilbert A(D)-modules. New York J. Math. 1 (1994) 26-38, electronic. M. D. Choi and E. Effros. Nuclear C*-algebras and the approximation property. Amer. J. Math. 100 (1978) 61-79. . Nuclear C*-algebras and injectivity: The general case. Indiana Univ. Math. J. 26 (1977) 443-446.
460 [CE3] [CE4] [CES]
[ChI] [Ch2] [Che] [Chr1] [Chr2] [Chr3] [Chr4] [Chr5] [Chr6] [ChS]
[CK] [CL]
[Col] [Co2] [Co3] [CoH]
Introduction to Operator Space Theory . Injectivity and operator spaces. J. Funct. Anal. 24 (1977) 156-209. . Separable nuclear C* -algebras and injectivity. Duke Math. J. 43 (1976) 309-322. E. Christensen, E. Effros, and A. Sinclair. Completely bounded multilinear maps and C*-algebraic cohomology. Invent. Math. 90 (1987) 279-296. M. D. Choi. A Schwarz inequality for positive linear maps on C*algebras. Illinois J. Math. 18 (1974) 565-574. M. D.Choi. A simple C*-algebra generated by two finite order unitaries. Canadian J. Math. 31 (1979) 887-890. L. Chen. An inequality for the multivariate normal distribution. J. Mult. Anal. 12 (1982) 306-315. E. Christensen. Extensions of derivations. J. Funct. Anal. 27 (1978) 234-247. _ _ _ . Extensions of derivations II. Math. Scand. 50 (1982) 111122. . On non self adjoint representations of operator algebras Amer. J. Math. 103 (1981) 817-834. . Similarities of I h factors with property f. J. Operator Theory 15 (1986) 281-288. . Perturbation of operator algebras II. Indiana Math. J. 26 (1977) 891-904. . Finite von Neumann algebra factors with property f. J. Funct. Anal. 186 (2001) 366-380. A. Chaterjee and R. Smith. The central Haagerup tensor product and maps between von Neumann algebras. J. Operator Theory 112 (1993) 97-120. E. Carlen and P. Kree. On martingale inequalities in noncommutative stochastic analysis. J. Funct. Anal. 158 (1998) 475-508. E. Carlen and E. Lieb. Optimal hypercontractivity for Fermi fields and related non-commutative integration inequalities. Comm. Math. Phys. 155 (1993) 27-46. A. Connes. Classification of injective factors, Cases I h, IIco , I II).., >. l=I. Ann. Math. 104 (1976) 73-116. . Non-Commutative Geometry. Academic Press, 1995. . A factor not anti-isomorphic to itself. Bull. London Math. Soc. 7 (1975), 171-174. M. Cowling and U. Haagerup. Completely bounded multipliers of the Fourier algebra of a simple Lie group of real rank one. Invent. Math. 96 (1989) 507-549.
References
461
[CS1] E. Christensen and A. Sinclair. Representations of completely bounded multilinear operators. J. Funct. Anal. 72 (1987) 151-181. [CS2] _ _ _ . A survey of completely bounded operators. Bull. London Math. Soc. 21 (1989) 417-448. [CS3] _ _ _ . On von Neumann algebras which are complemented subspaces of B(H). J. Funct. Anal. 122 (1994) 91-102. [CS4] _ _ _ . Module mappings into von Neumann algebras and injectivity. Proc. London Math. Soc. (3) 71 (1995) 618-640. [CS5] _ _ _ . Completely bounded isomorphisms of injective von Neumann algebras. Proc. Edinburgh Math. Soc. 32 (1989) 317-327. [Cu] J. Cuntz. Simple C*-algebras generated by isometries. Comm. Math. Phys. 57 (1977) 173-185. [Da1] K. Davidson. Nest Algebms. Wiley, 1988. [Da2] _ _ _ . C* -Algebms by Example. Fields Institute, Toronto, AMS publication, 1996. [DCH] J. de Canniere and U. Haagerup. l\,Iultipliers of the Fourier algebras of some simple Lie groups and their discrete subgroups. Amer. J. Math. 107 (1985) 455-500. [DHV] P. de la Harpe and A. Valette. La Propriete T de Kazhdan pour les Groupes Localement Compacts. Asterisque, Soc. Math. France 175 (1989). [Dil] J. Dixmier. Les Algebres d'Opemteurs dans l'Espace Hilbertien (Algebres de von Neumann). Gauthier-Villars, 1969. (In translation: von Neumann Algebms. North-Holland, 1981). [Di2] _ __ Les C* -algebres et leurs representations. Gauthier-Villars, 1969. [Di3] Formes lineaires sur un anneau d'operateurs. Bull. Soc. Math. Prance 81 (1953) 9--39. [Dixl] P. Dixon. Q-algebras. Unpublished Lecture Notes. Sheffield University, 1975. [Dix2] . Varieties of Banach algebras. Quart. J. Math. Oxford 27 (1976) 481-487. [DJT] J. Diestel, H. Jarchow, and A. Tonge. Absolutely Summing Opemtors. Cambridge University Press, 1995. [DP] K. Davidson and V. Paulsen. On polynomiaUy bounded operators. J. fUr die reine und angewandte Math. 487 (1997) 153-170. [E1] E. Effros. Aspects of non-commutative order. Notes for a lecture given at the Second US-Japan seminar on C* -algebras and applications to physics (April 1977) . [E2] . Advances in quantized functional analysis. Proceedings International Congress of Mathematicians, Berkeley, 1986, pp. 906916.
462
Introduction to Operator Space Theory
[EE] E. Effros and R. Exel. On Multilinear Double Commutant Theorems. Opemtor Algebms and Applications, Vol. 1, edited by D. Evans and M. Takesaki. London Math. Soc. Lecture Notes Series 135, pp. 81-94. [EH] E. Effros and U. Haagerup. Lifting problems and local reflexivity for C*-algebras. Duke Math. J. 52 (1985) 103-128. [EJR] E. Effros, M. Junge, and Z. J. Ruan. Integral mappings and the principle of local reflexivity for noncommutative L271-spaces. Ann. of Math. 151 (2000) 59-92. [EK] E. Effros and A. Kishimoto. Module maps and Hochschild-Johnson cohomology. Indiana Univ. Math. J. 36 (1987) 257-276. [EKR] E. Effros, J. Kraus, and Z. J. Ruan. On two quantized tensor products. Opemtor Algebms, Mathematical Physics and Low Dimensional Topology (Istanbul, 1991). A. K. Peters, 1993, pp. 125-145. [EL] E. Effros and C. Lance. Tensor products of operator algebras. Adv. Math. 25 (1977) 1-34. [EI] G. Elliott. On approximate finite dimensional von Neumann algebras I and II. Math. Scand. 39 (1976) 91-101, and Canad. Math. Bull. 21 (1978) 415-418. [En] P. Enflo. A counterexample to the approximation problem in Banach spaces. Acta Math. 130 (1973) 309-317. [EOR] E. Effros, N. Ozawa, and Z. J. Ruan. On injectivity and nuclearity for operator spaces. Duke Math. J. 110 (2001) 489-521. [ER1] E. Effros and Z.J. Ruan. On matricially normed spaces. Pacific J. Math. 132 (1988) 243-264. [ER2] . A new approach to operator spaces. Canadian Math. Bull. 34 (1991) 329-337. [ER3] . On the abstract characterization of operator spaces. Proc. Amer. Math. Soc. 119 (1993) 579-584. [ER4] . Self duality for the Haagerup tensor product and Hilbert space factorization. J. Funct. Anal. 100 (1991) 257-284. [ER5] . Recent development in operator spaces. Current Topics in Opemtor Algebms. Proceedings of the ICM-90 Satellite Conference Held in Nara (August 1990). World Sci. Publishing, 1991, pp. 146164. [ER6] . Mapping spaces and liftings for operator spaces. Proc. London Math. Soc. 69 (1994) 171-197. [ER7] . The Grothendieck-Pietsch and Dvoretzky-Rogers Theorems for operator spaces. J. Funct. Anal. 122 (1994) 428-450. [ER8] . On approximation properties for operator spaces. Int. J. Math. 1 (1990) 163-187.
References [ER9]
463
Representations of operator bimodules and their applications. J. Operator Theory 19 (1988) 137-157.
[ERIO] _ _ _ . Operator space tensor products and Hopf convolution algebras, J. Operator Theory. To appear. [ER11] _ _ _ . Operator Spaces. Oxford Univ. Press, 2000. [ER12]
. OCp-spaces, Contemporary Math. 228 (1998) 51-77.
[EvK] D. E. Evans and Y. Kawahigashi. Quantum Symmetries on Operator Algebras. Oxford University Press, 1998, xvi+829 pp. [EvL] D. Evans and J. Lewis. Dilations of irreversible evolutions in algebraic quantum theory. Commun. Dublin Inst. for Advanced Studies, Series . A (Theoretical Physics) 24 (1977). [EW] E. Effros and S. Winkler. Matrix convexity: operator analogues of the bipolar and Hahn-Banach theorems. J. Funct. Anal. 144 (1997) 117-152. [Fi] P. Fillmore. A User's Guide to Operator Algebras. eMS Series of Monographs and Advanced Texts. Wiley, 1996. [Fol] G.B. Folland. A Course in Abstract Harmonic Analysis. Studies in Advanced l\'Iathematics. eRe Press, 1995, x+276 pp. [FLM] T. Figiel, J. Lindenstrauss, and V. Milman. The dimensions of the spherical sections of convex bodies. Acta Math. 139 (1977) 53--94. [Fo] S. Foguel. A counterexample to a problem of Sz. Nagy. Pmc. Amer. Math. Soc. 15 (1964) 788-790. [FTP] A. Figa-Talamanca and M. Picardello. Harmonic Analysis on Free Gmups. Marcel Dekker, 1983. [GaR] D. Gal1par and A. Racz. An extension of a theorem of T. Ando. Michigan Math. J. 16 (1969) 377-380. [GH] L. Ge and D. Hadwin. Ultraproducts of C* -algebras. Preprint, 1999. [GHJ] F. Goodman, P. de la Harpe and V. F. R. Jones. Coxeter Graphs and Towers of Algebras. MSRI Publications, 14. Springer-Verlag, 1989. [GI] E. Gluskin. The diameter of the Minkowski compact urn is roughly equal to n. Funct. Anal. Appl. 15 (1981) 72-73. [Gr] A. Grothendieck. Resume de la theorie metrique des produits tensoriels topologiques. Boll. Soc. Mat. Baa-Paulo 8 (1956) 1-79. [Gro] M. Gromov. Random walk in random groups. January 2002.
Preprint, IHES,
[Gu1] A. Guichardet. Tensor products of C*-algebras. Dokl. Akad. Nauk. SSSR 160 (1965) 986-989. [Gu2]
. Algebres d'observables associees aux relations de commutation. Armand Colin, 1968.
464 [Gu3] [HI]
[H2]
[H3] [H4] [H5]
[H6] [H7] [H8] [Ha]
[Hal] [Ham] [Har1]
[Har2] [HiP] [HK]
[HP1]
Introduction to Operator Space Theory Symmetric Hilbert Spaces and Related Topics. Springer Lecture Notes 261, Springer-Verlag, 1972. U. Haagerup. Injectivity and decomposition of completely bounded maps. Opemtor Algebms and Their Connection with Topology and Ergodic Theory. Springer Lecture Notes in Math. 1132, SpringerVerlag, 1985, pp. 170-222. . An example of a non-nuclear C* -algebra which has the metric approximation property. Inventiones Math. 50 (1979) 279293. . Decomposition of completely bounded maps on operator algebras. Unpublished manuscript, Sept. 1980. . Self-polar forms, conditional expectations and the weak expectation property for C* -algebras. Unpublished manuscript, 1995. . A new proof of the equivalence of injectivity and hyperfiniteness for factors on a separable Hilbert space. J. Punct. Anal. 62 (1985) 160-201. . Solution of the similarity problem for cyclic representations of C*-algebras. Ann. Math. 118 (1983) 215-240. . The standard form of von Neumann algebras. Math. Scand. 37 (1975) 271-283. . Group C* -algebras without the completely bounded approximation property. Unpublished manuscript, :May 1986. L. Harris. Bounded Symmetric Domains in Infinite Dimensional Spaces. Springer Lecture Notes 364. Springer-Verlag, 1974 pp. 1340. P. Halmos. Ten problems in Hilbert space. Bull. Amer. Math. Soc. 75 (1970) 887-933. M. Hamana. Injective envelopes of C*-algebras. J. Math. Soc. Japan 31 (1979) 181-197. A. Harcharras. On some stability properties of the full C* -algebra associated to the free group Foo. Proc. Edinburgh Math. Soc. (2) 41 (1998) 93-116. . Fourier analysis, Schur multipliers on SP and non-commuta tive A(p)-sets. Studia Math. 137 (1999) 203-260. F. Hiai and D. Petz. The Semicircle Law, Free Random Variables and Entropy. Math. Surveys 77. Amer. Math. Soc., 2000. U. Haagerup and J. Kraus. Approximation properties for group C*algebras and group von Neumann algebras. Trans. Amer. Math. Soc. 344 (1994) 667-699. U. Haagerup and G. Pisier. Factorization of analytic functions with values in non-commutative Ll-spaces. Canadian J. Math. 41 (1989) 882-906.
References [HP2] [HT1] [HT2] HWW]
[J1] [J2] [JL1\I]
JNRX] [.Jo]
[JO] [Jol] [Jon] [JOR] [JP]
PR]
PS] [K] [KaJ
465
. Bounded linear operators between C* -algebras. Duke Math. J. 71 (1993) 889-925. U. Haagerup and S. Thorbjeirnsen. Random matrices and K-theory for exact C*-algebras. Documenta Math. 4 (1999) 341-450. . A new application of random matrices: Ext(C;(F2)) is not a group. (Preliminary version, August 30, 2002). P. Harmand, D. Werner, and W. Werner. 1If -Ideals in Banach Spaces and Banach Algebras. Lecture Notes in Mathematics 1547. SpringerVerlag, 1993, viii+387 pp. 1\1. Junge. Factorization Theory for Spaces of Operators. Habilitation thesis. Kiel University, 1996. . The projection constant of OHn and the little Grothendieck inequality. Preprint (2002). 1\1. Junge and C. Le Merdy. Factorization through matrix spaces for finite rank operators between C* -algebras. Duke Afath. J. 100 (1999) 299-319. M. Junge, N. Nielsen, Z. J. Ruan, and Q. XU. COCp-spaces-The local structure of non-commutative Lp-spaces. Adv. Math. To appear. 'V.B. Johnson. A complementary universal conjugate Banach space and its relation to the approximation problem. Israel J. Math. 13 (1972) 301-310 (1973). W. Johnson and T. Oikhberg. Separable lifting property and extensions of local reflexivity. Illinois J. Math. 45 (2001) 123-137. P. Jolissaint. A characterization of completely bounded multipliers of Fourier algebras. Colloquium Math. 63 (1992) 311-313. V. F. R. Jones. Subfactors and Knots. CBMS Regional Conference Series in Mathematics, 80. American Math. Soc., 1991. M. Junge, N. Ozawa, and Z. J. Rmin. OC(XJ spaces. Math. Ann. To appear. 1\1. Junge and G. Pisier. Bilinear forms on exact operator spaces and B(H) 0 B(H). Geometric and Functional Analysis (GAFA J.) 5 (1995) 329-363. M. Junge and Z. J. Ruan. Approximation properties for noncommutative Lp-spaces associated with discrete groups. Preprint (2001). To appear in Duke Math. J. V. Jones and V. S. Sunder. Introduction to Subfactors. Cambridge Univ. Press, 1997. H. Kesten. Symmetric random walks on groups. Trans. Amer. Math. Soc. 92 (1959) 336-354. Y. Katznelson. An Introduction to Harmonic Analysis. Wiley, 1968 (republished by Dover in 1976).
466
Introduction to Operator Space Theory
[Kad] R. Kadison. A generalized Schwarz inequality and algebraic invariants for operator algebras. Ann. of Math. 56 (1952) 494-503. [KaR] R. Kadison and J. Ringrose. Fundamentals of the Theory of Opemtor Algebms, Vol. II, Advanced Theory. Academic Press, 1986. [Kil] E. Kirchberg. On subalgebras of the CAR-algebra. J. Funct. Anal. 129 (1995) 35-63. [Ki2] . On non-semisplit extensions, tensor products and exactness of group C*-algebras. Invent. Math. 112 (1993) 449-489. [Ki3] . C*-nuclearity implies CPAP. Math. Nachr. 76 (1977) 203212. [Ki4] . Positive maps and C"-nuclear algebras. Proc. Inter. Conference on Opemtor Algebms, Ideals and Their Applications in Theoretical Physics (Leipzig 1977). Teubner Texte, 1978. [Ki5] . Commutants of unitaries in UHF algebras and functorial properties of exactness. J. Reine Angew. Math. 452 (1994) 39-77. [Ki6] . Personal communication. [Ki7] . On the matricial approximation property. Preprint. [Ki8] . Exact C* -algebras, Tensor products, and Classification of purely infinite algebras. Proceedings ICM 94, Zurich. Vol. 2, Birkhauser, 1995, pp. 943-954. [Ki9] . The derivation and the similarity problem are equivalent. J. Opemtor Theo'ry 36 (1996) 59-62. [KiP] E. Kirchberg and N.C. Phillips. Embedding of exact C*-algebras in the Cuntz algebra (')2. J. Reine Angew. Math. 525 (2000) 17-53. [KiW] E. Kirchberg and S. Wassermann. Exact groups and continuous bundles of C* -algebras. Math. Ann. 315 (1999) 169-203. [Kisl] S. Kislyakov. Operators that are (dis)similar to a contraction: Pisier's counterexample in terms of singular integrals. Zap. Nauchn. Semin. S.- Peterburg. Otdel. Mat. Inst. Steklov (POl\U). [Kis2] . Similarity problem for certain martingale uniform algebras. Preprint. [KiV] E. Kirchberg and G. Vaillant. On C*-algebras having linear, polynomial, and subexponential growth. Invent. Math. 108 (1992) 635-652. [KiW] E. Kirchberg and S. Wassermann. C* -algebras generated by operator systems. J. Funct. Anal. 155 (1998) 324-351. [Ko] O. Kouba. Interpolation of injective or projective tensor products of Banach spaces. J. Funct. Anal. 96 (1991) 38-61. [Ko] H. Konig. On the complex Grothendieck constant in the n-dimensional case. Proc. of the Strobl Conf. Austria 1989, edited by P. Mueller and W. Schachermayer. London Math. Soc. Lect. Notes 158, 1990, pp. 181-198.
References
467
[KPS] S. G. Krein, Yu. Petunin, and E. M. Semenov, Interpolation of Linear Operators. Translations of Mathematical Monographs 54. American Mathematical Society, 1982. [Kr] J. Kraus. The slice map problem and approximation properties. J. Funct. Anal. 102 (1991) 116-155. [KTJ1] H. Konig and N. Tomczak-Jaegermann. Bounds for projection constants and I-summing norms. Trans. Amer. Math. Soc. 320 (1990) 799-823. [KTJ2] . Norms of minimal projections. J. Funct. Anal. 119 (1994) 253-280. [Ku] W. Kuratowski. Topology Vol. 1. Academic Press, 1966 (new edition translated from the French). [Ky] S-H. Kye. Notes on Operator Algebras. Lecture Notes Series 7, Seoul National Univ., 1993. [KyR] S-H. Kye and Z-J. Ruan. On the local lifting property for operator spaces. J. Funct. Anal. 168 (1999) 355-379. [La1] C. Lance. On nuclear C*-algebras. J. Funct. Anal. 12 (1973) 157176. [La2] . Tensor products and nuclear C* -algebras. Operator Algebras and Applications. Amer. Math. Soc. Proc. Symposia Pure Math., 1982, Vol. 38, part 1, pp. 379-399. [Lac] H. E. Lacey. The Isometric Theory of Classical Banach Spaces. Springer-Verlag, 1974. [Leh1] F. Lehner. A characterization of the Leinert property. Proc. Amer. Math. Soc. 125 (1997) 3423-3431. [Leh2] . Free operators with operator coefficients. Colloq. Math. 74 (1997) 321-328. [Le] M. Leinert. Faltungsoperatoren auf gewissen diskreten Gruppen. Studia Math. 52 (1974) 149-158. [Leb] A. Lebow. A power bounded operator which is not polynomially bounded. Michigan Math. J. 15 (1968) 397-399. [LeM1] C. Le Merdy. Factorization of p-completely bounded multilinear maps. Pacific J. Math. 172 (1996) 187-213. [LeM2] . Representations of a quotient of a subalgebra of B(X). Math. Proc. Cambridge Phil. Soc. 119 (1996) 83-90. [LeM3] . On the duality of operator spaces. Canadian Math. Bull. 38 (1995) 334-346. [LeM4] . Self-adjointness criteria for operator algebras. Arch. Math. 74 (2000) 212-220. [LeM5] . A strong similarity property of nuclear C* -algebras. Rocky Mountain J. Math. 30 (2000) 279-292.
468
Introduction to Operator Space Theory
[LeM6] [LeM7] [Li] [LiR] [LPP] [LPS]
[LR] [LT1] [LT2] [LT3] [Lu] [LuP] [M] [Mal] [Ma2] [MaP]
[MaU] [Mat2] [Mi] [MN]
An operator space characterization of dual operator algebras. Amer. J. Math. 121 (1999) 55-63. . The weak* similarity property on dual operator algebras. Integr. Equ. Oper. Theory 37 (2000) 72-94. J. Lindenstrauss. Extension of Compact Opemtors. Mem. Amer. Math. Soc. No. 48, 1964, pp. 1-112. J. Lindenstrauss and H. P. Rosenthal. The Cp spaces. Ismel J. Math. 7 (1969) 325-349. F. Lust-Piquard and G. Pisier. Non-commutative Khintchine and Paley inequalities. Ark. for Mat. 29 (1991) 241-260. A. Lubotzky, R. Phillips, and P. Sarnak. Hecke operators and distributing points on S2, I. Comm. Pure and Applied Math. 39 (1986) 149-186. J. Lopez and K. Ross. Sidon Sets. Marcel Dekker, 1975. J. Lindenstrauss and L. Tzafriri. Classical Banach Spaces, Vol. l. Sequence Spaces. Springer-Verlag, 1976. . Classical Banach Spaces, Vol. II. Function Spaces. SpringerVerlag, 1979. Classical Banach Spaces. Springer Lecture Notes 338, Springer-Verlag, 1973. A. Lubotzky. Discrete Groups, Expanding Gmphs and Invariant Measures. Progress in Math. 125. Birkhiiuser, 1994. F. Lust-Piquard. Inegalites de Khintchine dans C p (1 < p < 00). C.R. Acad. Sci. Paris 303 (1986), 289-292. P. A. Meyer. Quantum Probability for Probabilists. Springer Lecture Notes 1538, Springer-Verlag, 1993. B. Magajna. Strong operator modules and the Haagerup tensor product. Proc. London Math. Soc. 74 (1997) 201-240. . The Haagerup norm on the tensor .product of operator modules. J. Funct. Anal. 129 (1995) 325-348. B. Maurey and G. Pisier. Series de variables aleatoires vectorielles independantes et proprietes geometriques des espaces de Banach. Studia Math. 58 (1976) 45-90. B. Mathes. A completely bounded view of Hilbert-Schmidt operators. Houston J. Math. 17 (1991) 404-418. . Characterizations of row and column Hilbert space. J. London Math. Soc. 50 (1994), no. 2, 199-208. K. Miyazaki. Some remarks on intermediate spaces. Bull. Kyushu Inst. Tech. 15 (1968) 1-23. P. Meyer-Nieberg. Banach Lattices. Universitext. Springer-Verlag, 1991.
References
469
[Mor] M. Morgenstern. Existence and explicit constructions of q+ 1 regular Ramanujan graphs for every prime power q. J. Gombin. Theory Ser. B 62 (1994) 44-62. [MP] B. Mathes and V. Paulsen. Operator ideals and operator spaces. Pmc. Amer. Math. Soc. 123 (1995) 1763-1772. [MS] V. Milman and G. Schechtman. Asymptotic Theory of Finite Dimensional Normed Spaces. Springer Lecture Notes 1200, Springer-Verlag, 1986. [Ne] E. Nelson. Notes on non-commutative integration. J. Funct. Anal. 15 (1974) 103-116. [Ni] N. Nikolskii. Treatise on the Shift Opemtor. Springer-Verlag, 1986. [NRl] M. Neal and B. Russo. Contractive projections and operator spaces. G. R. Acad. Sci. Paris 331 (2000) 873-878. [NR2] [01] [02] [03] [04] lOP] [OR] fORi] [Ozl]
[Oz2] [Oz3] [Oz4] [Oz5]
. A holomorphic characterization of ternary rings of operators. Preprint, 2001. To appear. T. Oikhberg. Direct sums of homogeneous operator spaces. J. London Math. Soc. 64 (2001) 144-160. . Geometry of operator spaces and products of orthogonal projections. Ph.D. Thesis, Texas A&M Univ., 1997. . Subspaces of maximal operator spaces. Integml Equations Opemtor Theory. To appear. . Completely complemented subspace problem. J. Opemtor Theory 43 (2000) 375-387. T. Oikhberg and G. Pisier. The "maximal" tensor product of two operator spaces. Pmc. Edinburgh Math. Soc. 42 (1999) 267-284. T. Oikhberg and H. P. Rosenthal. On certain extension properties for the space of compact operators. J. Funct. Anal. 179 (2001) 251-308. T. Oikhberg and E. Ricard. Operator spaces with few completely bounded maps. Preprint (2002). N. Ozawa. On the set of finite-dimensional subspaces of preduals of von Neumann algebras. G. R. Acad. Sci. Paris Sr. I Math. 331 (2000) 309-312. . Amenable actions and exactness for discrete groups. G. R. Acad. Sci. Paris Sr. I Math. 330 (2000) 691-695. . On the lifting property for universal C* -algebras of operator spaces. J. Opemtor Theory 46 (2001) 579-591. . A non-extendable bounded linear map between C* -algebras. Pmc. Edinburgh Math. Soc. 44 (2001) 241-248. . An application of expanders to B(f2) ~ B(f2). J. Funct. A nat. To appear.
470 [Oz6]
Introduction to Operator Space Theory Local Theory and Local Reflexivity for Operator Spaces. Ph.D. Thesis, Texas A&M Univ., May 2000.
[PI] G. Pisier. The operator Hilbert space OH, complex interpolation and tensor norms. Memoirs Amer. Math. Soc. 122,585 (1996) 1-103. [P2]
. Non-commutative vector valued Lp-spaces and completely p-summing maps. Soc. Math. France. Asterisque 247 (1998) 1-131.
[P3]
. Espace de Hilbert d'operateurs et interpolatiol1 complexe. C. R. Acad. Sci. Paris S. 1316 (H)93) 47-52.
. Factorization of Linear Operators and the Geometry of Banach Spaces. CBMS (Regional Conferences of the MvIS) 60, 1986; reprinted with corrections 1987. [P5] . Completely bounded maps between sets of Banach space operators. Indiana Univ. Math. J. 39 (1990) 251-277. [P4]
[P6]
. Exact operator spaces. Colloque sur les algebres d'operateurs. Recent Advances in Operator Algebras (Orleans 1992) Asterisque (Soc. Math. France) 232 (1995) 159-186.
[P7]
. Projections from a von Neumann algebra onto a subalgebra. Bull. Soc. Math. France 123 (1995) 139-153. [P8] . The Volume of Convex Bodies and Banach Space Geometry. Cambridge Univ. Press, 1989. [P9] . Dvoretzky's theorem for operator spaces and applications. Houston J. Math. 22 (1996) 399-416.
[PlO]
. Similarity Problems and Completely Bounded Maps, Second, Expanded Edition. Springer Lecture Notes 1618, Springer-Verlag, 2001. [Pl1] . A simple proof of a theorem of Kirchberg and related results on C* -norms. J. Operator Theory 35 (1996) 317-335. [P13]
. Remarks on complemented subspaces of von-Neumann algebras. Proc. Royal Soc. Edinburgh 121A (1992) 1-4.
[P14]
. Quadratic forms in unitary operators. Linear Algebra Appl. 267 (1997) 125-137.
Riesz Transforms: A Simpler Analytic Proof of P.-A. Meyer's Inequality. Seminaire de Probabilites XXII. Springer Lecture Notes 1321, Springer-Verlag, 1988. [P16] . Holomorphic semi-groups and the geometry of Banach spa ces. Ann. of Math. 115 (1982) 375-392. [P17] . The similarity degree of an operator algebra. St. Petersburg Math. J. 10 (1999) 103-146. [P18] . The similarity degree of an operator algebra II. Math. Zeit. 234 (2000) 53-81. [P15]
References
[P19] [P20]
[P21] [P22] [Pal] [Pa2] [Pa3] [Pa4] [Pa5] [PaP] [ParI] [Par2] [PaS] [PaSu] [Pat] [Pel] [Pe2]
[Ped] [Pes]
471
Remarks on the similarit.y degree of an operator algebra. Intern. J. Math. 12 (2001) 403-414. . A polynomially bounded operator on Hilbert space which is not similar to a contraction. J. Amer. Math. Soc. 10 (1997) 351-369. . A simple proof of a theorem of Jean Bourgain. Michigan Math. J. 39 (1992) 475-484. The Operator Hilbert Space OH and TYPE III von Neumann Algebras. Preprint (2002). V. Paulsen. Completely Bounded Maps and Dilations. Pitman Research Notes 146. Pit.man Longman (Wiley) 1986. . Representation of function algebras, Abstract operator spa ces and Banach space geometry. J. Funct. Anal. 109 (1992) 113-129. . Completely bounded maps on C* -algebras and invariant operator ranges. Proc. Amer. Math. Soc. 86 (1982) 91-96. . Completely bounded homomorphisms of operator algebras. Proc. Amer. Math. Soc. 92 (1984) 225-228. . The maximal operator space of a normed space. Proc. Edinburgh Math. Soc. 39 (1996) 309-323. V. Paulsen and S. Power. Tensor products of non-self-adjoint operator algebras. Rocky Mountain J. Math. 20 (1990) 331-350. S. Parrott. On a quotient norm and the Sz.-Nagy Foias lifting theorem. J. Funct. Anal. 30 (1978) 311-328. . Unitary dilations for commuting contractions. Pacific J. Math. 34 (1970) 481-490. V. Paulsen and R. Smith. Multilinear maps and tensor norms on operator systems. J. Funct. Anal. 73 (1987) 258-276. V. Paulsen and C- Y. Suen. Commutant representations of completely bounded maps. J. Opemtor Theory 13 (1985) 87-101. A. Paterson. Amenability. l\'lath. Surveys 29. Amer. Math. Soc., 1988. V. Peller. Estimates of functions of power bounded operators on Hilbert space J. Opemtor Theory 7 (1982) 341-372. V. Peller. Estimates of functions of Hilbert space operators, similarity to a contraction and related function algebras. Linear and Complex Analysis Problem Book, edited by Havin, Hruscev, and Nikolskii. Springer Lecture Notes 1043, Springer-Verlag, 1984, pp. 199-204. G. Pedersen. C* - Algebras and Their Automorphism Groups. Academic Press, 1979. V. Pestov. Operator spaces and residually finite-dimensional C*Algebras. J. Funct. Anal. 123 (1994) 308-317.
472
Introduction to Operator Space Theory
[PiS] G. Pisier and D. Shlyakhtenko. Grothendieck's theorem for operator spaces. Inventiones Math. 150 (2002) 185-217. [Pol] S. Popa. A short proof of injectivity implies hyperfiniteness for finite von Neumann algebras: J. Operator Theory 16 (1986) 261-272. [Po2] . On amenability in type I It factors. Operator Algebras and Applications, edited by D. Evans and M. Takesaki, volume 2, LMS Lecture Notes Series 136. Cambridge Univ. Press, 1988. [Po3] S. Popa. Classification of Subfactors and Their Endomorphisms. CBMS Regional Conference Series in Mathematics, 86. American Math. Soc., 1995. [Pop] C. Pop. Bimodules normes represent-abIes sur des espaces hilbertiens. Ph.D. Thesis, University of Orleans, 1999. [Pow] S. C. Power. Hankel Operators on Hilbert Space. Research Notes in Mathematics, 64. Pitman, 1982. [Pu1] G. Popescu. Von Neumann inequality for (B('H)nh. Math. Scand. 68 (1991) 292-304. [Pu2] . Universal operator algebras associated to contractive sequences of non-commuting operators. J. London Math. Soc. 58 (1998) 469-479. [Pu3] . Poisson transforms on some C* -algebras generated by isometries. J. Funct. Ana.l. 161 (1999) 27-61. [PV1] M. Pimsner and D. Voiculescu. Exact sequences for K-groups and Ext-groups of certain cross-products of C* -algebras. J. Operator Theory 4 (1980) 93-118. [PV2] . K-groups of reduced crossed products by free groups. J. Operator Theory 8 (1982) 131-156. [PX] G. Pisier and Q. XU. Inegalites de martingales non commutatives. C. R. Acad. Sci. Paris Sr. I Math. 323 (1996) 817-822. [R] H. P. Rosenthal. The complete separable exteI!sion property. J. Operator Theory 43 (2000) 329-374. [Ril] E. Ricard. Decomposition de HI, multiplicateurs de Schur et espaces d'operateurs. These, Universite Paris VI, 2001. [Ri2] . A tensor norm on Q-spaces. J. Operator Theory 48 (2002) 431-445. [Rie] M. Rieffel. Induced representations of C*-algebras. Adv. Math. 13 (1974) 176-257. [RLL] M. R~rdam, F. Larsen, and N. Laustsen. An Introduction to KTheory for C27*-Algebras. Cambridge Univ. Press, 2000. [RoJ A. R. Robertson. Injective matricial Hilbert spaces. Math. Proc. Cambridge Philos. Soc. 110 (1991) 183-190.
References
473
[RS] M. Rl'lrdam and E. Stl'lrmer. Classification of Nuclear C*-Algebms. Entropy in Operator Algebras. Springer-Verlag, 2002. [RW] A. R Robertson and S. Wassermann. Completely bounded isomorphisms of injective operator systems. Bull. London Math. Soc. 21 (1989) 285-290. [Ru1] Z. J. Ruan. Subspaces of C*-algebras. J. Funct. Anal. 76 (1988) 217--230. [Ru2]
. Injectivity of operator spaces. Trans. Amer. Math. Soc. 315 (1989) 89-104.
[Sa] S. Sakai. C*-Algebms and W*-Algebms. Springer-Verlag, 1974. [Sar] P. Sarnak. Some Applications of Modular Forms. Cambridge Univ. Press, 1990. [Se] I. Segal. A non-commutative extension of abstract integration. Ann. Math. 57 (1953) 401-457. ISh] y. Shalom. Bounded generation and Kazhdan's property (T). Inst. Hautes Etudes Sci. Publ. Math. No. 90 (1999), 145-168. [Sk] G. Skandalis. Algebres de von Neumann de groupes libres et probabilites non commutatives. Seminaire Bourbaki, Vol. 1992/93. Asterisque No. 216 (1993) 87-102. [Sm1] R R Smith. Completely contractive factorization of C* -algebras. J. Funct. Anal. 64 (1985) 330-337. [Sm2] [SN]
[SNF] [SSl] [SS2] [SS3] 1St] ISta] IStZl
. Completely bounded maps between C* -algebras. J. London Math. Soc. 27 (1983) 157-166. B. Sz.-Nagy. Completely continuous operators with uniformly bounded iterates. Publ. Math. Inst. Hungarian Acad. Sci. 4 (1959) 89-92. B. Sz.-Nagy and C. Foias. Harmonic Analysis of Opemtors on Hilbert Space. Akademiai Kiad6, 1970. A. M. Sinclair and RR Smith. The Haagerup invariant for operator algebras. Amer. J. Math. 117 (1995) 441-456. . Hochschild Cohomology of von Neumann Algebms. London Math. Soc. Lecture Notes Series. Cambridge Univ. Press, 1995. . Factorization of completely bounded bilinear operators and injectivity. J. Funct. Anal. 157 (1998) 62-87. W. Stinespring. Positive functions on C* -algebras. Proc. A mer. Math. Soc. 6 (1955), 211-216. J. D. Stafney. The spectrum of an operator on an interpolation space. Trans. Amer. Math. Soc. 144 (1969) 333-349. S. Stratila and L. Zsid6. Lectures on von Neumann Algebms. Editura Academiei, Bucharest; Abacus Press, Tunbridge Wells, 1979.
474
Introduction to Operator Space TlJeory
[Su] C-Y. Suen, Completely bounded maps on C*-algebras. Proc. Amer. Math. Soc. 93 (1985) 81-87. [Sun] V. S. Sunder. An Invitation to von Neumann Algebms. Universitext. Springer-Verlag, 1987. [Sz] A. Szankowski. B{H) does not have the approximation property. Acta Math. 147 (1981) 89-108. [T] S. Thorbj(2lrnsen. Mixed Moments of Voiculescu's Gaussian Random Matrices. Preprint, Odense Univ., 1999. [Tal] M. Takesaki. A note on the cross-norm of the direct product of C*algebras. Kodai Math. Sem. Rep. 10 (1958) 137-140. [Ta2] . On the cross-norm of the direct product of C* -algebras. Tohoku Math. J. 16 (1964) 111-122. [Ta3] . Theory of Opemtor Algebms /. Springer-Verlag, 1979. Banach-Mazur Distances and Finite[TJ1] N. Tomczak-Jaegermann. Dimensional Opemtor Ideals. Pitman-Longman (Wiley), 1989. [TJ2] N. Tomczak-Jaegermann. The moduli of convexity and smoothness and the Rademacher averages of trace class Sp. Studia Math. 50 (1974) 163-182. [To1] J. Tomiyama. On the projection of norm one in W* -algebras. Proc. Japan Acad. 33 (1957) 608-612. [T02] Tensor Products and Projections of Norm One in von Neumann Algebms. Lecture Notes, Univ. of Copenhagen, 1970. [T03] . Applications of Fubini type theorem to the tensor product of C*-algebras. Tohoku Math. J. 19 (1967) 213-226. [Tor] A. M. Torpe. Nuclear C* -Algebms and Injective von Neumann. Odense Univ. Preprint, 1981. [Tr] S. Trott. A pair of generators for the unimodular group. Canadian Math. Bull. 3 (1962) 245-252. [Va] A. Valette. An application of Ramanujan graphs to C* -algebra tensor products. Discrete Math. 167/168 (1997) 597-603. [VI] N. Varopoulos. Some remarks on Q-algebras. Ann. Inst. Fourier 22 (1972) 1-11. [V2] . A theorem on operator algebras. Math. Scand. 37 (1975) 173-182. [VDN] D. Voiculescu, K. Dykema, and A. Nica. Free Random Variables. CRM Monograph Series, Vol. 1, Amer. Math. Soc., 1992. [Vol] D. Voiculescu. Property T and approximation of operators. Bull. London Math. Soc. 22 (1990) 25-30. [V02] . Limit laws for random matrices and free products. Invent. Math. 104 (1991) 201-220.
475
References
[Wal] S. Wassermann. On tensor products of certain group C* -algebras. J. Funct. Anal. 23 (1976) 239-254. [Wa2]
. Exact C* -Algebms and Related Topics. Series, 19. Seoul National Univ., 1994.
[Wa3]
. On subquotients of UHF algebras. Math. Proc. Cambridge Phil. Soc. 115 (1994) 489-500.
Lecture Notes
. The slice map problem for C* -algebras. Proc. London Math. Soc. 32 (1976) 537-559. [Wa5] . Injective W* -algebras. Math. Proc. Cambridge Phil. Soc. 82 (1977) 39-47. [Wa4]
[Watl] F. Watbled. Interpolation complexe d'un espace de Banach et de son antidual. C. R. Acad. Sci. Paris Sr. I Math. 321 (1995) 1437-1440. [Wat2] F. Watbled. Complex interpolation of a Banach space with its dual. Math. Scand. 87 (2000) 200-210. [We] J. Wermer. Quotient algebras of uniform algebras. Symp. on function algebms and mtional approximation. Univ. of l\Hchigan, 1969. [WeO] N. Wegge-Olsen. K-Theory and C*-Algebms. Oxford Univ. Press, 1993. [Win] S. Winkler. M8;trix convexity. Ph.D. Thesis, UCLA. 1996. [Witl] G. Wittstock. Ein operatorwertigen Hahn-Banach Satz. J. Funct. Anal. 40 (1981), 127-150. [Wit2]
[WW] [Xu] [Y] [ZI] [Z2] [Z3]
. Extensions of completely bounded module morphisms. Proc. Conference on Opemtor Algebras and Group Representations. Neptum, Pitman, 1983. C. Webster and S. Winkler. The Krein-Milman theorem in operator convexity. Trans. Amer. Math. Soc. 351 (1999) 307-322. Q. XU. Interpolation of operator spaces. J. Funct. Anal. 139 (1996) 500-539. K. Ylinen. Representing completely bounded multilinear operators. Acta Math. Hungar. 56 (1990) 295-297. C. Zhang. Representations of operator spaces. Preprint, Univ. of Houston, 1993. Completely bounded Banach-Mazur distance. Proc. Edinburgh Math. Soc. (2) 40 (1997) 247-260. . Representation and geometry of operator spaces. Ph.D. Thesis, Univ. of Houston (1995).
SUBJECT INDEX Antisymmetric Fock space 80, 173 Approximation property 85, 232, 240, 250, 293-296, 301 Bidual47, 48, 232, 267, 273, 303-315 Blecher-Paulsen factorization 384 Column Hilbert space 21-23, 95, 96, 172,177,341 Completely bounded 18 Complex conjugate 63 Complex interpolation 52, 106-107, 135-141, 147, 178. 270, 271 G*-norm 227 Cuntz algebra 175 Decomposable map 230, 261 Direct sum 51 Duality 40 Dvoretzky's theorem 215 Exact 285-302, 321 Factorization 23, 92, 101, 317 Fermion 173-175 Fock space 173, 176, 205 Free group 155, 182, 183, 188, 214, 215, 325, 331 Free product 98-101, 160, 365 Gaussian random variable 145, 191, 203, 331, 455 Grothendieck's theorem 316 Group G* -algebra 148 Haagerup tensor product 86-108, 128, 390 Halmos similarity problem 407 Hankel matrix 165-171, 410, 414 Hilbertian operator space 122, 173 Homogeneous operator space 172, 216, 217 Infinite Haagerup tensor product 390 Injective 232, 267, 273, 355, 356-357 Intersection (of operator spaces) 55 Kazhdan's property T 328 Kirchberg's conjecture 283 Kirchberg's theorem 252, 261, 296-297 Local lifting property (LLP) 275
Local reflexivity 303-315 Lp-space 138, 191 I\laximal tensor product (of G* -algebra) 149, 227 I\Iaximal tensor product (of nonself-adjoint operator algebra) 366 Maximal operator space structure on a Banach space 71, 323 Minimal opE:'rator space structurE:' on a Banach space 71, 321, 323 Minimal tensor product 28 Multilinear factorization 92 Multiplier 27,151-153,181 Nuclear 231, 232, 250, 269, 273,296 OLLP 277 Operator Hilbert space 122-147 Opposite 64 Perturbation 68 Point of continuity 339 Projective 359 Projective tensor product 81, 181 Properties G, G', and Gil 309-313 Property T 328 Quantization 65 Quotient by a subspace 42 Quotient by an ideal 43 Quotient of WEP (QWEP) 274, 283 Rademacher functions 192 Ramanujan graph 327 Random matrix 203,215,331,332,455 Reduced G* -algebra 149. 183 Row Hilbert space 21-23, 95-96, 172, 177, 341 Ruan's theorem 35 Schur multiplier 27, 151-153 SE:'mi-circular syst.em 215 Similarity problem 396, 407 Spin system 76,80,173,321,413 Sum (of operator spaces) 55 SymmE:'trized Haagerup tensor product 102 Sz.-Nagy-Halmos similarity problem 407
478
Introduction to Operator Space Theory
Ultraproduct 59, 210 Unitization of an operator space 163 Universal 67 Vector-valued Lp-spaces 140, 180
Von Neumann algebra 47 Weak expectation property (WEP) 267, 283
NOTATION INDEX A(D) (disc algebra) 380, 407
A1 * A2 (free product) 98 A1 *A 2 (unital free product) 98
C*(G) (full C*-algebra) 148 C~(G) (reduced C*-algebra) 149 C* (E) (universal C* -algebra of E) 160 C~ (E) (universal unital C* -algebra of E) 160 Cn, C (column Hilbert spaces) 21 CB(E, F) (space of c.b. maps from E to F) 19 d(A) (similarity degree of A) 403 d(E, F) (Banach-1Vlazur distance) 334 deh(E, F) (c.b. version of Banach-l\lazur distance, or "c. b. distance") 20 df(E) (infimum of c.b. distance of E to a subspace of C* (IF 00)) 345 dsx(E) (infimum of c.b. distance of E to a subspace of X) 8 dsdE) (infimum of c.b. distance of E to a subspace of K) 286, 342 6(n) 343 6 (logarithm of d) 334 6eb (logarithm of deb) 335 $ (direct sum, "block diagonal" or "in £00 sense") 51 $1 (direct sum in £1 sense) .52 $p (direct sum in £p sense) 54 D(G) (diagonal operators on £2(G)) 297 Efj 182,336 E~ 183 E 0 F (algebraic tensor product) 1, 28 E 0 m in F (minimal tensor product) 1,28 E 0h F (Haagerup tensor product) 87 E 0 1, F (symmetrized Haagerup tensor product) 102 E 0/\ F (o.s. projective tensor product) 81 E (unitization) 163 (Eo, Edo (complex interpolation space) 53
E* 40 cI>(I), cI>n ("Fermionic" Hilbertian operator space) 173 H 02 I< (Hilbertian tensor product) 1
K2 £(A) (length of A) 401 >"a (left regular representation on G) 148
>"(n) 352 >"(X) 134, 355 f~ 78
ff
78 rnin(E), min(e~) 71, 77 rnax(E), max(£~) 71,77 JlI* (predual of AI) 47, 305 AI0N (t.ensor product in von Neumann sense) 49 A1n (n x n matrices with complex entries) 12 AIn (E) (n x n matrices with entries in E) 2, 18 M<.p (Schur multiplier) 151-153 1I.llcb (c.b. norm) 19 II.IIHS (Hilbert-Schmidt norm) 21
OAu(E) 113 OH, OHn (operator Hilbert space) 5,123
OSn (set of n-dimensional operator spaces) 335 (ultraproduct) 59 11'1 11'2 ("free product" of morphisms) 98 11'1 *11'2 ("free product" of unit.al morphisms) 98 R n , R (row Hilbert spaces) 21 R + C 186,194 RnC 184 ( ®jEI E j ) It (infinite Haagerup Tensor TIiEI E;JU
*
Product)
Sp (Schatten p-class) 140 Tr 411 Ua (universal representation on G)
148 UC*(G) 297
Printed in the United Kingdom by Lightning Source UK Ltd. 113367UKSOOOOIB/131
1111111111111111 111111 1III
9 780521 811651