Multiplier Operator Algebras and Applications

Similar documents
ORDERED INVOLUTIVE OPERATOR SPACES

Lecture 1 Operator spaces and their duality. David Blecher, University of Houston

METRIC CHARACTERIZATIONS OF ISOMETRIES AND OF UNITAL OPERATOR SPACES AND SYSTEMS

Multipliers, C -modules, and algebraic structure in spaces of Hilbert space operators

Acta Mathematica Academiae Paedagogicae Nyíregyháziensis 24 (2008), ISSN

Morita equivalence of dual operator algebras

arxiv: v1 [math.oa] 21 Aug 2018

arxiv:math/ v4 [math.oa] 28 Dec 2005

Operator Space Approximation Properties for Group C*-algebras

On Some Local Operator Space Properties

BERNARD RUSSO. University of California, Irvine

Lectures 1 and 2: Operator algebras, etc (basic theory and real positivity) David Blecher. December 2016

Preliminaries on von Neumann algebras and operator spaces. Magdalena Musat University of Copenhagen. Copenhagen, January 25, 2010

BERNARD RUSSO. University of California, Irvine

Tensor algebras and subproduct systems arising from stochastic matrices

Substrictly Cyclic Operators

BERNARD RUSSO. University of California, Irvine QUEEN MARY COLLEGE UNIVERSITY OF LONDON DEPARTMENT OF MATHEMATICS

ALGEBRA II: RINGS AND MODULES OVER LITTLE RINGS.

arxiv:math/ v1 [math.oa] 4 Jan 2007 JAN M. CAMERON

Lecture 5: Peak sets for operator algebras and quantum cardinals. David Blecher. December 2016

von Neumann algebras, II 1 factors, and their subfactors V.S. Sunder (IMSc, Chennai)

arxiv:math/ v1 [math.oa] 9 May 2005

Spectral Theory, with an Introduction to Operator Means. William L. Green

TRACIAL POSITIVE LINEAR MAPS OF C*-ALGEBRAS

Non commutative Khintchine inequalities and Grothendieck s theo

N. CHRISTOPHER PHILLIPS

Normality of adjointable module maps

I will view the Kadison Singer question in the following form. Let M be the von Neumann algebra direct sum of complex matrix algebras of all sizes,

COMPLEXIFICATIONS OF REAL OPERATOR SPACES

Norms of Elementary Operators

arxiv: v3 [math.oa] 17 Apr 2013

Spectral isometries into commutative Banach algebras

A COMMENT ON FREE GROUP FACTORS

ON C*-ALGEBRAS WHICH CANNOT BE DECOMPOSED INTO TENSOR PRODUCTS WITH BOTH FACTORS INFINITE-DIMENSIONAL

ON THE SIMILARITY OF CENTERED OPERATORS TO CONTRACTIONS. Srdjan Petrović

arxiv: v2 [math.oa] 9 Jan 2014

NOTES ON THE UNIQUE EXTENSION PROPERTY

Universität des Saarlandes. Fachrichtung 6.1 Mathematik

A 3 3 DILATION COUNTEREXAMPLE

A Survey of Lance s Weak Expectation Property, The Double Commutant Expectation Property, and The Operator System Local Lifting Property

Orthogonal Pure States in Operator Theory

Fréchet algebras of finite type

SEMICROSSED PRODUCTS OF THE DISK ALGEBRA

Non-separable AF-algebras

Boolean Inner-Product Spaces and Boolean Matrices

VANISHING OF SECOND COHOMOLOGY FOR TENSOR PRODUCTS OF TYPE II 1 VON NEUMANN ALGEBRAS

Injective semigroup-algebras

arxiv:math/ v1 [math.fa] 7 Feb 2007

ON CONNES AMENABILITY OF UPPER TRIANGULAR MATRIX ALGEBRAS

Linear Algebra. Preliminary Lecture Notes

Thomas Hoover and Alan Lambert

Linear Algebra. Preliminary Lecture Notes

A DECOMPOSITION OF E 0 -SEMIGROUPS

Chapter 2 Linear Transformations

S. DUTTA AND T. S. S. R. K. RAO

FURTHER STUDIES OF STRONGLY AMENABLE -REPRESENTATIONS OF LAU -ALGEBRAS

ABELIAN SELF-COMMUTATORS IN FINITE FACTORS

Spectrally Bounded Operators on Simple C*-Algebras, II

NOTES ON PRODUCT SYSTEMS

I teach myself... Hilbert spaces

Spectral theory for linear operators on L 1 or C(K) spaces

Patryk Pagacz. Characterization of strong stability of power-bounded operators. Uniwersytet Jagielloński

THE COMPLETELY BOUNDED APPROXIMATION PROPERTY FOR EXTENDED CUNTZ PIMSNER ALGEBRAS

Perron-Frobenius theorem

STABLY ISOMORPHIC DUAL OPERATOR ALGEBRAS

On Problems on von Neumann Algebras by R. Kadison, 1967

arxiv: v4 [math.fa] 19 Apr 2010

Means of unitaries, conjugations, and the Friedrichs operator

Introduction to Bases in Banach Spaces

ON NON-SELF-ADJOINT OPERATOR ALGEBRAS

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

THE CHOQUET BOUNDARY OF AN OPERATOR SYSTEM

Elementary linear algebra

Fiberwise two-sided multiplications on homogeneous C*-algebras

Math 121 Homework 5: Notes on Selected Problems

COMMON COMPLEMENTS OF TWO SUBSPACES OF A HILBERT SPACE

A brief introduction to trace class operators

Outline. Multipliers and the Fourier algebra. Centralisers continued. Centralisers. Matthew Daws. July 2009

FREE PROBABILITY THEORY

What is a Hilbert C -module? arxiv:math/ v1 [math.oa] 29 Dec 2002

RIGHT-LEFT SYMMETRY OF RIGHT NONSINGULAR RIGHT MAX-MIN CS PRIME RINGS

The Choquet boundary of an operator system

Math 210B. Artin Rees and completions

A Version of the Grothendieck Conjecture for p-adic Local Fields

Rational and H dilation

René Bartsch and Harry Poppe (Received 4 July, 2015)

INTERMEDIATE BIMODULES FOR CROSSED PRODUCTS OF VON NEUMANN ALGEBRAS

Two-sided multiplications and phantom line bundles

Stability of Adjointable Mappings in Hilbert

PRIME NON-COMMUTATIVE JB -ALGEBRAS

Boundedly complete weak-cauchy basic sequences in Banach spaces with the PCP

Topics in Representation Theory: Fourier Analysis and the Peter Weyl Theorem

BASIC VON NEUMANN ALGEBRA THEORY

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

ON k-subspaces OF L-VECTOR-SPACES. George M. Bergman

Where is matrix multiplication locally open?

FUNCTION BASES FOR TOPOLOGICAL VECTOR SPACES. Yılmaz Yılmaz

Universal Algebra for Logics

arxiv: v2 [math.oa] 21 Nov 2010

arxiv: v1 [math.oa] 23 Jul 2014

Transcription:

Multiplier Operator Algebras and Applications David P. Blecher* Department of Mathematics, University of Houston, Houston, TX 77204. Vrej Zarikian Department of Mathematics, University of Texas, Austin, TX 78712. Abstract The one-sided multipliers of an operator space X are a key to latent operator algebraic structure in X. We begin with a survey of these multipliers, together with several of the applications that they have had to operator algebras. We then describe several new results on one-sided multipliers, and new applications, mostly to one-sided M-ideals. Classification: Physical Sciences, Mathematics Corresponding Author: David P. Blecher, Department of Mathematics, University of Houston, Houston, TX 77204-3008. Phone (713)-743-3451, Fax (713)-743-3505, E-mail: dblecher@math.uh.edu Text pages: 9 Words in abstract: 55 Characters in paper: *D. P. B. was supported in part by a grant from the National Science Foundation. V.Z. was supported by a NSF VIGRE Postdoctoral Fellowship. 1

1 Introduction For every closed linear subspace X of B(H), for a Hilbert space H, the first author has introduced two subalgebras A l (X) M l (X) B(X), where B(X) is the space of bounded linear maps on X. The first, M l (X), is an operator algebra (we will define this term formally in the next section). The second, A l (X), is a C -algebra, and indeed is a W -algebra (that is, it is an abstract von Neumann algebra ), if X is weak* closed in B(H). The elements of these algebras are maps on X, called left multipliers of X. Of course there is a matching theory of right multipliers, but we shall avoid mentioning this since it is essentially identical (with appropriate modifications such as replacing left by right and column by row, and so on). By a one-sided multiplier we mean of course a left or a right multiplier. The following list gives the main areas where the theory of multiplier algebras is being applied currently. We will discuss each point in greater detail below. (1) Most importantly perhaps, the algebras A l (X) and M l (X) are key to latent operator algebraic structure in the operator space X. We discuss this in further detail in Sections 3 and 4. (2) These algebras allow one to define and develop a noncommutative M-ideal theory; this is treated in Sections 5 and 6. We recall for now that classical M-ideals are a basic tool for Banach spaces, and they have an extensive and useful theory [1]. We develop a one-sided, or noncommutative, generalization of M-ideals, which we call left and right M-ideals, and we show that this class includes examples that are of interest to operator algebraists. We then are able to extend much of the classical M-ideal theory to our setting. (3) For many key examples, the spaces A l (X) and M l (X) do consist precisely of the maps on X of most interest. For example, if X is a C -module (see for example [2, 3] for the definition and basic theory of these objects), then these spaces consist of the maps that C -algebraists are interested in (see Example 4.1 (3)). Many of our results on left multipliers and right M-ideals may then be viewed as generalizations of important facts about C -modules, and their dual variant W -modules. Thus the multiplier algebras facilitate a generalization of some C - and W -module techniques to operator spaces. We will not however emphasize this aspect in this note. (4) Left multipliers work admirably in settings involving dual spaces, unlike some earlier techniques. We will amplify this point at the end of Sections 4 and 6. The notion of left multipliers above, and the corresponding algebras A l (X) and M l (X), are best viewed as a sequence of equivalent definitions, some of which we will meet in Sections 3 and 4. No one is best for all purposes; each seems good for some purposes, and unsuitable for others. However one of these definitions appears to be most useful. Namely the unit ball of M l (X) may be described as consisting of the linear maps T : X X satisfying [ ] ] T (x) x y [ (1) y (the precise statement may be found in Theorem 3.2). Because the criterion in Eq. (1) is so simple, it is easy to see that left multipliers remain left multipliers under most of the basic functional analytic operations: subspaces, tensor products, duality, interpolation, and so on. This is extremely useful in our theory and applications. This note is addressed to a general reader with a knowledge of functional analysis; the background facts needed may be found in the first six chapters of [4] for example. 2 Operator spaces and operator algebras We begin by recalling a few basic definitions from operator space theory. 2

A concrete operator space is a linear subspace X of B(H), for a Hilbert space H. For such a subspace X we may view the space M m,n (X) of m n matrices with entries in X as the corresponding subspace of B(H (n), H (m) ). Thus this matrix space has a natural norm, for each m, n N. An abstract operator space is a vector space X with a norm n on M n (X), for all n N, which is completely isometric to a concrete operator space. The latter term, complete isometry, refers to a linear map T : X Y such that [T (x ij )] n = [x ij ] n for all n N, [x ij ] M n (X). A complete contraction is defined similarly, simply replacing = by in the last centered equation. There is also a celebrated theorem of Ruan, characterizing operator spaces in terms of two simple conditions on the matrix norms. The interested reader should consult [5] for this and for other facts listed in this section, or the texts [6, 7]. D.P.B. is also completing a text on related matters with Le Merdy (to appear, Oxford University Press). We saw above that if X is an operator space then so is M m,n (X). In particular we will reserve the symbols C n (X) and R n (X) for M n,1 (X) and M 1,n (X), respectively. Every operator space X has a dual object which is also an operator space. Namely, we write X for the dual Banach space of X, with the norm of a matrix [f ij ] M n (X ) given by [f ij ] n = sup{ [f ij (x kl )] : [x kl ] Ball(M n (X))}. We call X, equipped with this operator space structure, a dual operator space. An early result in operator space theory, due to D.P.B. and Paulsen, and independently to Effros and Ruan, states that the canonical map from X into X is a complete isometry. Every Banach space X may be made into an operator space in canonical ways. The simplest of these is called MIN(X); this is obtained by assigning the smallest matrix norms { n } n 2 with respect to which X becomes an operator space (note that we do not mention n = 1 here, this is just the usual norm on X). Another way of describing MIN(X) is to embed X isometrically into a commutative C -algebra C(K) (this may be done with K = Ball(X ) for example, using the Hahn-Banach theorem). Then MIN(X) is simply X endowed with the canonical matrix norms n inherited from the C -algebra M n (C(K)) = C(K, M n ). A very nice situation occurs when a given noncommutative/operator space theory, applied to MIN(X) spaces, is just the associated classical theory. We will see an illustration of this in Section 5; right M-ideals of MIN(X) turn out to be precisely the classical M-ideals of X. A concrete operator algebra is a subalgebra A of B(H). For simplicity in this exposition, we shall always assume that A also has an identity of norm 1. By an (abstract) operator algebra we will mean a unital algebra which is also an operator space, which is completely isometrically homomorphic to a concrete operator algebra. There is an operator algebraic version of Ruan s theorem, due to D.P.B., Ruan and Sinclair [8]: Theorem 2.1 Up to completely isometric homomorphism, the abstract operator algebras are precisely the unital algebras A, which are also operator spaces, satisfying 1 = 1 and ab n a n b n, for all n N and a, b M n (A). As noted later, this theorem may be proved easily from the theory of left multipliers. Operator algebras may be regarded as the noncommutative function algebras. Indeed note that every function algebra or uniform algebra is an operator algebra (being a subalgebra of a commutative C -algebra). A main tool for studying a general operator algebra A is Arveson s noncommutative Shilov boundary [9, 10, 11, 12]. This is, loosely speaking, the smallest C -algebra containing A: it is a quotient of any other C -algebra containing A. Thus for the disk algebra A = A(D), for example, the smallest C -algebra containing A is C(T), the continuous functions on the circle. More generally, for an operator space X, the noncommutative Shilov boundary may be viewed as a smallest C -module (or ternary ring of operators ) containing X completely isometrically. Because we are aiming at an elementary presentation of our subject we will not give a more precise definition of this boundary, referring the interested reader to [12, 13] for more details. However it is in the background, and will be mentioned from time to time in passing. 3

3 Multipliers and algebraic structure With the extra structure consisting of the additional matrix norms on an operator algebra, one might expect to not have to rely as heavily on other structure, such as the product. Indeed one may ask questions like the following: Can one recover the product on an operator algebra A from its norm (or matrix norms)? Or could one describe the right ideals of A without referring to the product? The answers to these kinds of questions are essentially in the affirmative, as we shall see in the sequel, the key being the operator space multipliers mentioned in the introduction. These objects were introduced by D.P.B. in 99 (see the last cited paper); and independently by W. Werner but in an ordered context (see [14] for example). Soon after the main results of [13] were found, D.P.B. and Paulsen established the alternative approach to one-sided multipliers that is presented in [15]. A year later D.P.B., Effros, and V.Z. found the very practical criterion Eq. (1) (see [16]). Each of these contributed to the growing list of equivalent definitions of one-sided multipliers. As we said in the introduction, for any operator space X we have two subalgebras A l (X) M l (X) B(X). These can be defined in terms of the noncommutative Shilov boundary [13] or in terms of the injective envelope [15]. However for the sake of a simpler exposition we will not emphasize these approaches, and instead we will begin with the following equivalent definition of M l (X): namely as the linear maps T : X X such that there exists some completely isometric embedding X B(K), for a Hilbert space K, and an operator a B(K) with T (x) = ax, x X. We define the multiplier norm of T to be the infimum of the quantities a for all K, a as above. Finally, the matrix norms on M l (X) are specified by viewing a matrix t = [T ij ] in M n (M l (X)) as a map L t from C n (X) = M n,1 (X) to itself, via the formula L t ([x i ]) = [ j T ijx j ]. It turns out that L t is also a left multiplier, and we define [T ij ] Mn(M l (X)) to be the multiplier norm of L t. Proposition 3.1 For any operator space X, M l (X) is an operator algebra. The last result may be proved in several ways. In fact it is more or less obvious from the equivalent definition of left multipliers in terms of the noncommutative Shilov boundary or injective envelope [13, 15]. Not surprisingly, one may also prove Proposition 3.1 using Theorem 2.1. However one of the attractive features of the left multiplier approach is that it automatically yields theorems such as Theorem 2.1 as corollaries, and moreover allows a relaxation of some of the hypotheses of such characterization results. This was D.P.B. s main motivation for the introduction of operator space multipliers [13]. Thus for example Theorem 2.1 follows in just a few lines by combining Proposition 3.1 with the following result (see for example the proof of Theorem 1.11 in [17] for details; this deduction was independently noticed by Paulsen). Theorem 3.2 (Blecher, Effros, and Zarikian) [16] A linear map T : X X on an operator space X is in Ball(M l (X)) if and only if [ ] [ ] T xij xij, all n N and [x ij ], [y ij ] M n (X). y ij The norms in the last centered equation are just the norms in M 2n,n (X). We remark that later simplifications of our original proof of this result were given by Paulsen [7], Werner, and by D.P.B.. (In fact Theorem 3.2 was inspired by a similar theorem of Werner in an ordered context. Later Werner showed in [14] how one may view operator space multipliers within his ordered context, and also showed how Theorem 3.2 may be deduced from his earlier result.) The condition in Theorem 3.2 is purely in terms of the vector space structure and matrix norms on X; and yet it often encodes operator algebraic structure. For example, if A is an operator algebra (with an identity of norm 1), then M l (A) = A completely isometrically homomorphically. In fact the regular representation λ : A B(A) defined by λ(a)(b) = ab, is a completely isometric homomorphism onto M l (A). y ij 4

From this last fact we can provide a recipe to recover the product in an operator algebra. Let us suppose that A is an operator algebra with a forgotten product. For simplicity let us first suppose that we do remember the identity element e. Form M l (A) using Theorem 3.2, and define θ: M l (A) A by θ(t ) = T (e). Then it follows from the fact in the last paragraph that the forgotten product on A is ab = θ(θ 1 (a)θ 1 (b)), for a, b A. If we have forgotten the identity element e too, then one may only hope to recover the product modulo the unitary subgroup of A, but this may be done in a similar manner to the previous case (see the remarks on p. 316 in [16] for details). Of course once one has such a recipe for recovering the product, one would expect Banach-Stone type theorems for complete isometries between operator algebras to be more or less immediate. D.P.B. has discussed this elsewhere, and of course such results all have their origin in Kadison s well known Banach-Stone theorem for isometries between C -algebras [18]. In summary, we are clearly beginning to substantiate our statement that left multipliers are a key to operator algebraic structure. Another fact along these lines is that the operator module actions (in the sense of [19]) of an operator algebra A on an operator space X, are in a bijective correspondence with completely contractive unital homomorphisms θ : A M l (X). Indeed if A is a C -algebra then the correspondence is also with unital -homomorphisms from A into A l (X). This result, originally due to D.P.B. (see [13], and [15] for an alternative approach), also follows in just a few lines from Theorem 3.2. 4 The C -algebra A l (X) As was the case for left multipliers, there are several equivalent definitions of A l (X). For example, it may be defined to be the diagonal C -algebra of M l (X); that is, the span of the Hermitian elements in that operator algebra. Alternatively, although we will not use this formulation, A l (X) may be defined to be the set of maps T : X X for which there is a map S : X X with T x, y = x, Sy, x, y X, where the inner product in the last formula is the operator-valued one inherited from the inner product on a noncommutative Shilov boundary of X. If H is a Hilbert space for example, and if H c is H thought of as a fixed column in B(H), then the inner product above on X = H c may be taken to be the usual inner product. A similar remark holds for C -modules, which yields the facts stated next. Example 4.1 (1) For a Hilbert space H, M l (H c ) = A l (H c ) = B(H) (see the notation above). (2) For a C -algebra A, A l (A) is the usual C -algebra multiplier algebra M(A) of A, whereas M l (A) is the usual left multiplier algebra LM(A) (see 3.12 in [20]). (3) Generalizing examples (1) and (2), if Y is a C -module, then M l (Y ) is the space of bounded module maps on Y, whereas A l (Y ) is the C -algebra of adjointable maps in the usual C - module sense (see [3]). These examples show that in certain cases, the spaces A l (X) and M l (X) do consist precisely of the maps on X of most interest. However we must issue a warning: for a general space both M l (X) and A l (X) may be one-dimensional! In fact the prevalence of spaces with only trivial onesided multipliers is disappointing, until one remembers that this is simply reflecting the lack of any trace of operator algebra or operator module structure in such spaces. Thus one should not expect the classical or noncommutative L 1 spaces to have any nontrivial one-sided multipliers. This is also related to a lack of noncommutative M-structure, as we shall see later. On the other hand one would expect L 1 to have noncommutative L-structure. Interestingly, classical and noncommutative L p spaces for p 2 may have nontrivial left multipliers and nontrivial noncommutative M-structure. This may be seen by first noting that by Example 4.1 (1) above, L 2 has plenty of left multipliers. So does L by the fourth last paragraph of Section 3. Using Theorem 3.2 it is not hard to see that the complex interpolation method, applied to two left multipliers, produces a left multiplier on the interpolated space, which in this case is L p (with a slightly nonstandard operator space structure ). 5

A result which is extremely important in applications is the following fact from [16]: Theorem 4.2 If X is a dual operator space, then A l (X) is a W -algebra. This result opens another door to the use of von Neumann algebra methods in operator space theory. For example, the type decomposition, or Murray-von Neumann equivalences, in A l (X), will have consequences for the structure of X. At the very least the theorem implies that for a dual space X, the algebra A l (X) will have lots of orthogonal projections (if it is nontrivial), and that these projections may be utilized in the ways that von Neumann algebraists and operator theorists are familiar with. These projections are called left M-projections on X, and they form a basic ingredient in the noncommutative M-ideal theory presented in the last two sections. In the remainder of this section we continue to discuss duality. We recall that operator algebra questions often come in two varieties, a norm/c -algebraic type or a weak*/von Neumann algebraic type. One often proves a theorem of the first type by first proving an analogous result of the second type, and then deducing the desired result by going to the second dual. For function algebras and algebras of operators there is often a technical obstacle to this approach; namely that the Shilov boundary (or injective envelope) rarely works well in tandem with duality. Although operator space multipliers may be defined in terms of the noncommutative Shilov boundary (or injective envelope), they behave extremely well with respect to duality. This is mostly because of the existence of the criterion in Theorem 3.2. For example, using this criterion it is easy to see that T Ball(M l (X)) if and only if T Ball(M l (X )). (2) Using duality properties of multipliers, one may prove versions of many of the basic theorems concerning operator algebras and their modules which are appropriate to dual spaces and the weak* topology. For example, the left multiplier approach and Theorem 3.2 were key to the first proof (see [17]) of the following nonselfadjoint version of Sakai s characterization of von Neumann algebras: Theorem 4.3 (Le Merdy-Blecher) An operator algebra A is completely isometrically isomorphic (via a weak* homeomorphism) to a weak* closed unital subalgebra of B(H), if and only if A is a dual operator space. 5 Right ideals in operator spaces The one-sided multipliers above are intimately related to D.P.B., Effros, and V.Z. s generalization of the classical M-ideal notion to right M-ideals in operator spaces (see [16]). First we recall the classical definitions, due to Alfsen and Effros [21, 1]. These are based on the notion of the direct sum of two Banach spaces; namely the algebraic sum of the spaces, with the norm (x, y) = max{ x, y }. An M-projection on a Banach space X is an idempotent linear map P : X X such that the map from X X X given by x (P (x), (I P )(x)) (3) is an isometry. This is simply saying that X = P (X) (I P )(X). The range of an M-projection is a called an M-summand. A closed subspace J of a Banach space X is called an M-ideal if J is the range of an M-projection on X. These objects have an extensive and useful theory. Thus in the text [1] the reader will find an extensive calculus for M-ideals, by which we mean a collection of hundreds of theorems concerning the properties of M-ideals, together with a long list of examples. For example the M-ideals in a C -algebra are just the closed two-sided ideals (see [21, 22]). Analysts realizing that they have encountered an M-ideal in the course of a calculation, often need only to consult such a text to obtain a desired conclusion. It is therefore natural to ask if there is a noncommutative, one-sided, notion of M-ideals, which at the very least has the property that the right M-ideals in MIN(X) are the classical M-ideals in X, and the right M-ideals in C -algebras are the closed right ideals. One could then hope to connect such objects to the rich existing one-sided ideal theory of C -algebras, which is intimately related 6

to noncommutative topology (see for example [23, 24]). Our goal in the rest of this article is to describe such a notion, to relate our definition to one-sided multipliers, and then begin to assemble a calculus for one-sided M-ideals. Such a program was proposed in 2000 by E. G. Effros; and this was initiated in [16], and in V. Z. s Ph. D. thesis [25]. It turns out that there are several equivalent definitions of left M-projections on an operator space X. For example they may be defined to be the idempotent maps P : X X such that the map in Eq. (3) is a complete isometry, but now viewed as a map from X to M 2,1 (X) = C 2 (X). Alternatively, the set of left M-projections on X is exactly the set of orthogonal projections P in the C -algebra A l (X). The range of a left M-projection is a right M-summand. As before, we call a closed subspace J of an operator space X a right M-ideal if J is a right M-summand in X. The last few definitions are stated only in terms of the vector space structure, and matrix norms. The following examples show that they often encode important algebraic information. Example 5.1 (1) The M-ideals in a Banach space X are exactly the right M-ideals in MIN(X). (2) The right M-ideals in a C -algebra are exactly the closed right ideals. (3) The right M-ideals in a C -module are exactly the closed submodules. (4) The right M-ideals in an operator algebra are exactly the closed right ideals with a left contractive approximate identity. As we said earlier, one would not expect spaces like L 1 to have interesting one-sided M-structure. In the sequel [26] to the paper [16], we show for example that preduals of von Neumann algebras have no nontrivial one-sided M-ideals. Rather, such spaces will be interesting from the standpoint of one-sided L-ideal theory, where the latter is the dual theory to that of one-sided M-ideals. 6 The calculus of one-sided M-ideals In this section we briefly present some new results, and also describe a lengthy study which we have undertaken of one-sided M-ideals and multipliers. As mentioned in the last section, we wish to generalize the basic theorems from classical M-ideal theory, in order to develop a theory which will be useful for some problems in noncommutative functional analysis. As we shall see here, this often reduces to proving that left multipliers have certain properties, or to von Neumann algebra techniques and spectral theory (recall that A l (X) is a von Neumann algebra if X is a dual space). Indeed most of the basic theory of classical M-ideals does generalize, but usually requires quite different, noncommutative, arguments! And perhaps surprisingly, the results do not become untidy in the noncommutative framework, although sometimes we have to add reasonable hypotheses. As one would expect, one also finds truly new phenomena in the new framework (one example of this is the fact that we mentioned above Theorem 4.2, of the existence of nontrivial one-sided M-ideals in L p spaces). To illustrate some of the points above, consider the following result: Theorem 6.1 The closed span in an operator algebra of a family of closed right ideals, each possessing a left contractive approximate identity for itself, is a right ideal also possessing a left contractive approximate identity. A result like this is useful in practice, since the presence of an approximate identity is often key to calculations involving approximations. It is also interesting that this result is not true for Banach algebras, as one may easily see by testing three- or four-dimensional examples. In fact we claim that the result is essentially a one-sided M-ideal phenomenon. Indeed, Theorem 6.1 is immediate from Example 5.1 (4), and the following general result: Theorem 6.2 The closed span of a family {J λ : λ Λ} of right M-ideals in an operator space X, is a right M-ideal in X. 7

We sketch the proof of this theorem because it is quite instructive, and shows the noncommutative nature of our arguments. By the definition of right M-ideals, and basic Banach space duality theory (passage to the second dual X ), one sees that the theorem will follow if it is the case that the weak* closed span of families of right M-summands of a dual space Y, is again a right M-summand of Y. By Theorem 4.2, A l (Y ) is a W -algebra. By basic properties of families of projections in von Neumann algebras, it is easy to reduce the theorem to the assertion that the weak* closed span of two right M-summands is again a right M-summand. In fact a much more striking result is true, the map P P (Y ) is a lattice isomorphism between the lattice of projections in the W -algebra A l (Y ), and the lattice of right M-summands of Y. The meet in the latter lattice is intersection, while the join is given by the weak* closed span. In particular we claim that if P, Q are two left M-projections, then the weak* closure of P (Y ) + Q(Y ) is (P Q)(Y ). To see this we use the fact that P Q is the weak* limit in A l (Y ) of (P + Q) 1/n, which in fact is valid for any two projections P, Q in a von Neumann algebra. If x is in (P Q)(Y ) one can show that x = (P Q)(x) is the weak* limit of (P + Q) 1/n (x). By a spectral theory argument one can see that (P + Q) 1/n (x) is in the norm closure of P (Y ) + Q(Y ). Thus x, and hence (P Q)(Y ), is contained in the weak* closure of P (Y ) + Q(Y ). The reverse containment is quite obvious. There are two main techniques for proving results about one-sided M-ideals. We just saw an illustration of the first technique: heavy use of the fact that A l ( ) is a W -algebra for dual spaces, and the consequent permission to use von Neumann algebra facts and spectral theory. The second main technique is to first establish a result for A l (X) or M l (X), often using Theorem 3.2 for this, and then to use the fact that the projections in these algebras are just the left M-projections. We give just one illustration of this latter technique, an application to dual spaces. It follows from Eq. (2) that for any operator space X, we may regard M l (X) as a subalgebra of M l (X ). It is easy to see that this inclusion embeds A l (X) as a C -subalgebra of the W -algebra A l (X ). More is true if X is a dual operator space; in this case we claim that there is a canonical conditional expectation E from the W -algebra A l (X ) onto its subalgebra A l (X), and moreover E is weak* continuous with respect to the weak* topologies for which these algebras are W -algebras (see Theorem 4.2). We sketch the key idea in the proof of this claim. We define E by E(T ) = ι Y T ι X, where Y is a predual of X and ι Y is the canonical complete isometry from Y into Y. If T Ball(M l (X )), and x, y X then [ ] [ ] [ ] [ ] [ ] E(T )(x) = ι Y (T (ι X (x))) y ι Y (ι T (ιx (x)) X(y)) ιx (x) ι X (y) = x ι X (y), y where the second inequality follows from Theorem 3.2. A similar argument with matrices shows that E(T ) satisfies the criterion in Theorem 3.2. Thus E is a contraction from M l (X ) into M l (X). It is then easy to show that E maps A l (X ) into A l (X). Checking the remaining assertions of the claim is then routine. The claim established in the last paragraph has for example the following corollaries: Corollary 6.3 The weak* closure of a one-sided M-ideal in a dual operator space X is a one-sided M-summand of X. To see this, suppose that J is the right M-ideal, and consider the canonical left M-projection P of X onto J. Let Q = E(P ), where E is the conditional expectation above. By a weak* continuity argument one shows that Q is a left M-projection onto the weak* closure of J. Corollary 6.4 (Kaplansky density for M ideals) The unit ball of a one-sided M-ideal in a dual operator space is weak* dense in the unit ball of its weak* closure. To see this, suppose that J, P, Q are as discussed in the paragraph above. If x X is in the unit ball of the weak* closure of J, by Goldstine s theorem we may choose a net (x t ) in Ball(J) 8

converging in the weak* topology of X to P (x). Thus x t = Q(x t ) = E(P )(x t ) converges in the weak* topology of X to E(P )(x) = Q(x) = x. These corollaries are also true for general dual Banach spaces, with the word one-sided removed throughout. This may be seen by applying these results to MIN(X) for a Banach space X. We were surprised to find no trace of such results in the M-ideal literature. References [1] Harmand, P., Werner, D. & Werner, W. (1993) M-ideals in Banach spaces and Banach algebras (Lecture Notes in Mathematics 1547, Springer-Verlag, Berlin-Heidelberg-New York). [2] Rieffel, M. A. (1982) Proceedings of Symposia in Pure Mathematics, Am. Math. Soc, Providence, RI 38, 285-298. [3] Lance, E. C. (1995) Hilbert C -modules A toolkit for operator algebraists, (London Math. Soc. Lecture Notes, Cambridge University Press). [4] Kadison, R. V. & Ringrose J. R. (1997) Fundamentals of the Theory of Operator Algebras (Amer. Math. Soc., Providence, RI), Vol. 1 & 2. [5] Effros, E. G. & Ruan, Z. J. (2000) Operator Spaces, (Oxford University Press, Oxford). [6] Pisier G. (2003) Introduction to operator space theory, (London Math. Soc. Lecture Note Series 294, Cambridge University Press). [7] Paulsen, V. I. (2002) Completely bounded maps and operator algebras, (Cambridge University Press). [8] Blecher, D. P., Ruan, Z. J. & Sinclair, A. M. (1990) J. Funct. Anal. 89, 188-201. [9] Arveson, W. B. (1969) Acta Math. 123, 141-224 [10] Arveson, W. B. (1972) Acta Math. 128, 271-308. [11] Hamana, M. (1979) Publ. R.I.M.S. Kyoto Univ. 15, 773-785. [12] Hamana, M. (1999) Math. J. Toyama University 22, 77-93. [13] Blecher, D. P. (2001) J. Funct. Anal. 182, 280 343. [14] W. Werner, J. Funct. Anal., In Press. [15] Blecher, D. P. & Paulsen V. I. (2001) Pacific J. Math. 200, 1 17. [16] Blecher, D. P., Effros, E. G., & Zarikian, V. (2002) Pacific J. Math. 206, 287 319. [17] Blecher, D. P. (2001) J. Funct. Anal. 183, 498 525. [18] Kadison, R. V. (1951) Ann. of Math. 54, 325-338. [19] Christensen, E., Effros, E. G. & Sinclair, A. M. (1987) Inv. Math. 90, 279-296. [20] Pedersen, G. K. (1979) C -algebras and their automorphism groups, (Academic Press). [21] Alfsen, E. M. & Effros E. G. (1972) Ann. of Math. 96, 98 173. [22] Smith, R. R. & Ward J. D. (1978) J. Funct. Anal. 27, 337 349. [23] Effros, E. G. (1963) Duke Math. J. 30, 391 412. [24] Akemann, C. A. (1969) J. Funct. Anal. 4, 277-294. [25] Zarikian, V. (2001) Complete one-sided M-ideals in operator spaces (Ph.D. thesis, UCLA). [26] Blecher, D. P., Smith, R. R. & V. Zarikian (2002) J. Op. Theory, In Press. 9