arxiv:math/ v1 [math.qa] 18 Sep 2000

Similar documents
Traces for star products on symplectic manifolds

FLABBY STRICT DEFORMATION QUANTIZATIONS AND K-GROUPS

arxiv:hep-th/ v1 10 Mar 2003

The Erwin Schrodinger International Boltzmanngasse 9. Institute for Mathematical Physics A-1090 Wien, Austria

Classifying Morita equivalent star products

Natural star products on symplectic manifolds and quantum moment maps

arxiv: v1 [math.qa] 23 Jul 2016

UNIVERSIDADE FEDERAL DO PARANÁ Luiz Henrique Pereira Pêgas THE HOCHSCHILD-KOSTANT-ROSENBERG THEOREM FOR SMOOTH MANIFOLDS. Curitiba, 2010.

Involutions and Representations for Reduced Quantum Algebras

Lecture I: Constrained Hamiltonian systems

ON CONTRAVARIANT PRODUCT CONJUGATE CONNECTIONS. 1. Preliminaries

arxiv:math/ v1 [math.dg] 29 Sep 1998

Del Pezzo surfaces and non-commutative geometry

Chern characters via connections up to homotopy. Marius Crainic. Department of Mathematics, Utrecht University, The Netherlands

Convergence of Star Products and the Nuclear Weyl Algebr

LECTURE 26: THE CHERN-WEIL THEORY

A star product for the volume form Nambu-Poisson structure on a Kähler manifold.

arxiv: v1 [math.qa] 1 Jul 2015

SYMPLECTIC MANIFOLDS, GEOMETRIC QUANTIZATION, AND UNITARY REPRESENTATIONS OF LIE GROUPS. 1. Introduction

CYCLIC HOMOLOGY AND THE BEILINSON-MANIN-SCHECHTMAN CENTRAL EXTENSION. Ezra Getzler Harvard University, Cambridge MA 02138

Quantizations and classical non-commutative non-associative algebras

Deformation Quantization

Deformation groupoids and index theory

THE NONCOMMUTATIVE TORUS

STRICT QUANTIZATIONS OF ALMOST POISSON MANIFOLDS

Noncommutative Geometry and Applications to Physics. Jonathan Rosenberg University of Maryland

HYPERKÄHLER MANIFOLDS

Morita Equivalence in Deformation Quantization. Henrique Bursztyn

SEMISIMPLE LIE GROUPS

LECTURE 1: LINEAR SYMPLECTIC GEOMETRY

Lecture 11: Clifford algebras

CANONICAL METRICS AND STABILITY OF PROJECTIVE VARIETIES

Hochschild and cyclic homology of a family of Auslander algebras

Atiyah classes and homotopy algebras

arxiv:q-alg/ v2 18 Nov 1997

GEOMETRIC QUANTIZATION

MILNOR SEMINAR: DIFFERENTIAL FORMS AND CHERN CLASSES

The symplectic structure on moduli space (in memory of Andreas Floer)

Vector fields in the presence of a contact structure

Two-sided multiplications and phantom line bundles

ISOMORPHISMS OF POISSON AND JACOBI BRACKETS

Physical justification for using the tensor product to describe two quantum systems as one joint system

DEFORMATIONS OF ALGEBRAS IN NONCOMMUTATIVE ALGEBRAIC GEOMETRY EXERCISE SHEET 1

ON ISOTROPY OF QUADRATIC PAIR

class # MATH 7711, AUTUMN 2017 M-W-F 3:00 p.m., BE 128 A DAY-BY-DAY LIST OF TOPICS

Group Actions and Cohomology in the Calculus of Variations

arxiv: v1 [math.gr] 8 Nov 2008

Lecture on Equivariant Cohomology

arxiv:math/ v2 [math.qa] 19 Mar 2002

EXERCISES IN POISSON GEOMETRY

Lie bialgebras real Cohomology

Complexes of Hilbert C -modules

Notes on p-divisible Groups

Clifford Algebras and Spin Groups

THE MODULAR CLASS OF A LIE ALGEBROID COMORPHISM

Collected trivialities on algebra derivations

Metrics and Holonomy

Noncommutative geometry, quantum symmetries and quantum gravity II

Higgs Bundles and Character Varieties

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

The Gauss-Manin Connection for the Cyclic Homology of Smooth Deformations, and Noncommutative Tori

LECTURE 16: LIE GROUPS AND THEIR LIE ALGEBRAS. 1. Lie groups

REVERSALS ON SFT S. 1. Introduction and preliminaries

How to sharpen a tridiagonal pair

Global Seiberg-Witten quantization for U(n)-bundles on tori

Formal power series rings, inverse limits, and I-adic completions of rings

Non-separable AF-algebras

De Rham Cohomology. Smooth singular cochains. (Hatcher, 2.1)

arxiv: v1 [math.rt] 11 Sep 2009

Lie groupoids, cyclic homology and index theory

The Spinor Representation

A Bridge between Algebra and Topology: Swan s Theorem

CONSIDERATION OF COMPACT MINIMAL SURFACES IN 4-DIMENSIONAL FLAT TORI IN TERMS OF DEGENERATE GAUSS MAP

arxiv:math-ph/ v1 25 Feb 2002

Infinitesimal Einstein Deformations. Kähler Manifolds

Acceleration bundles on Banach and Fréchet manifolds

Kleine AG: Travaux de Shimura

LECTURE IV: PERFECT PRISMS AND PERFECTOID RINGS

BULLETIN (New Series) OF THE AMERICAN MATHEMATICAL SOCIETY Volume 33, Number 1, January 1996

OF AZUMAYA ALGEBRAS OVER HENSEL PAIRS

SYMPLECTIC LEAVES AND DEFORMATION QUANTIZATION

Lie Algebra of Unit Tangent Bundle in Minkowski 3-Space

LECTURE 2: SYMPLECTIC VECTOR BUNDLES

MAT 5330 Algebraic Geometry: Quiver Varieties

BERNARD RUSSO. University of California, Irvine

Deformations of coisotropic submanifolds in symplectic geometry

ABSTRACT DIFFERENTIAL GEOMETRY VIA SHEAF THEORY

Fiberwise two-sided multiplications on homogeneous C*-algebras

Injective semigroup-algebras

THE QUANTUM CONNECTION

Symplectic varieties and Poisson deformations

THE EULER CHARACTERISTIC OF A LIE GROUP

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

arxiv:hep-th/ v1 29 Nov 2000

SYMMETRIC SUBGROUP ACTIONS ON ISOTROPIC GRASSMANNIANS

Dirac Structures on Banach Lie Algebroids

arxiv:alg-geom/ v1 29 Jul 1993

SERRE FINITENESS AND SERRE VANISHING FOR NON-COMMUTATIVE P 1 -BUNDLES ADAM NYMAN

On projective classification of plane curves

Pacific Journal of Mathematics

Transcription:

Deformation Quantization of Hermitian Vector Bundles arxiv:math/0009170v1 [math.qa] 18 Sep 2000 Henrique Bursztyn Department of Mathematics UC Berkeley 94720 Berkeley, CA, USA Stefan Waldmann Département de Mathématique Université Libre de Bruxelles Campus Plaine, C. P. 218 Boulevard du Triomphe B-1050 Bruxelles Belgique September 2000 Abstract Motivated by deformation quantization, we consider in this paper -algebras A over rings C = R(i), where R is an ordered ring and i 2 = 1, and study the deformation theory of projective modules over these algebras carrying the additional structure of a (positive) A-valued inner product. For A = C (M), M a manifold, these modules can be identified with Hermitian vector bundles E over M. We show that for a fixed Hermitian star-product on M, these modules can always be deformed in a unique way, up to (isometric) equivalence. We observe that there is a natural bijection between the sets of equivalence classes of local Hermitian deformations of C (M) and Γ (End(E)) and that the corresponding deformed algebras are formally Morita equivalent, an algebraic generalization of strong Morita equivalence of C -algebras. We also discuss the semi-classical geometry arising from these deformations. henrique@math.berkeley.edu Research supported by a fellowship from CNPq, Grant 200481/96-7. Stefan.Waldmann@ulb.ac.be Research supported by the Communauté française de Belgique, through an Action de Recherche Concertée de la Direction de la Recherche Scientifique. 1

1 Introduction The concept of formal deformation quantization was first introduced in [2] and its goal is to construct quantum observable algebras by means of formal deformations (in the sense of Gerstenhaber [20]) of classical Poisson algebras, usually given by the algebra of complex-valued smooth functions on a Poisson manifold. The deformed associative algebra structures arising in this way are called star-products. Their existence and classification have been established through the joint effort of many authors (see [3, 14, 17, 27, 31, 32, 34, 43] and [21, 39, 42] for surveys on the subject). The aim of the present paper is to study a certain class of modules over these star-product algebras. In order to better understand how deformation quantization relates to the usual formalism of quantum mechanics through operators on Hilbert spaces, one is naturally led to consider representations of star-product algebras. By using the natural order structure in the ring R[[λ]] and considering Hermitian star-products (i.e., star-products for which the pointwise complex conjugation of functions is a -involution), a theory of -representations of star-product algebras on formal pre-hilbert spaces can be developed, similar to the usual representation theory of C -algebras. In fact, these ideas can be carried out in the more general setting of -algebras over C = R(i), where R is an ordered ring and C is its ring extension by i, with i 2 = 1. This category encompasses both star-product algebras and operator algebras. This approach was started in [7] with a formal version of the GNS construction and has been further investigated in [4 6, 41]. In this context, one can formulate the notion of formal Morita equivalence [9, 11], which is a generalization of Rieffel s notion of strong Morita equivalence of C -algebras [36]. An important role in this theory is played by finitely generated projective modules over unital algebras A carrying the additional structure of an A-valued inner product. For A = C (M), M an arbitrary manifold, these modules can be naturally identified with (smooth sections of) smooth Hermitian vector bundles over M. The main subject of the present paper is the deformation quantization of projective inner product A-modules. We remark that other authors have introduced quantization of vector bundles in different contexts, such as geometric quantization, quantum groups and non-commutative geometry [8, 12, 22, 23, 37]. This paper is organized as follows. In Section 2.1 we collect some basic facts about -algebras over ordered rings. Section 2.2 reviews some aspects of the deformation theory of -algebras and their projections. In Section 2.3, we discuss deformations of projective modules with inner-products. The main result in this section is Proposition 2.8. We emphasize here the additional Hermitian structure, since the result without inner product seems to be well-known, although not stated in the literature in this form. In Section 3.1, we specialize the discussion of Section 2.3 to the case of Hermitian vector bundles E M, where M is a Poisson manifold with a star-product. We show existence and uniqueness of deformation quantization of these objects, see Definition 3.1 and Theorem 3.2. Using this result, we also show that there is a natural bijection between the sets of equivalence classes of local Hermitian deformations of the -algebras C (M) and Γ (End(E)) in Proposition 3.3. This generalizes the result in [28], where the the case of trivial bundles is treated through a different approach. We show in Section 4 that corresponding deformations are formally Morita equivalent (Proposition 4.3). The important ingredient for this result is the notion of a strongly full projection (see Definition 4.1). A discussion about the underlying semi-classical geometry corresponding to the deformations just mentioned (in the sense of [35]) is presented in Section 3.2. We remark that techniques used here have been previously used by other authors, see e.g. [16, 18, 38]. The idea of deforming projections is present in Fedosov s index theorems [18] (see also [31, 32]) and in the star-product formulation of multicomponent WKB expansions by Weinstein and Emmrich [16] (see also [15]). It would be interesting to explore the connections between the present paper and these topics. We also believe this paper is an important step in extending the 2

approach to WKB approximations taken in [5, 6] to the multicomponent case. Finally, let us mention that various authors in particle physics have considered noncommutative gauge fields, see e.g. [24, 29] and the references therein. Here the fields can be viewed as sections of deformed vector bundles, but mainly trivial bundles have been treated so far. The Hermitian metrics (and their deformations) should play an important role in this context since they typically provide the potential term in the Lagrangians of physical models. This will be the subject of future investigations. 2 Algebraic preliminaries 2.1 -algebras over ordered rings We will recall here some basic definitions concerning -algebras over ordered rings. Further details can be found in [7, 9 11]. All algebras in this paper will be assumed to be associative. Let R be an ordered ring, i.e. an associative, commutative, and unital ring with a subset P R so that R is the disjoint union R = P {0} P and P P P, P + P P. An element a P is called positive and we denote it by a > 0. We remark that if R is ordered, then the ring R[[λ]] of formal power series has a natural ordering given by r=0 a rλ r > 0 if a r0 > 0, where r 0 is the first index with nonvanishing coefficient. The important examples of ordered rings in deformation quantization are R and R[[λ]]. We also define C = R(i) to be the quadratic ring extension of R by i, with i 2 = 1. Complex conjugation z z is defined in the usual way. In the following we shall assume for simplicity that Q R. Let A be an associative algebra over C, equipped with a -involution, i.e. an involutive C- antilinear antiautomorphism : A A. We call A a -algebra over C. We define Hermitian, unitary and normal elements in A in the usual way. A linear functional ω : A C is called positive if ω(a A) 0 for all A A. An element A A is then called positive if ω(a) 0 for all positive linear functionals ω. The set of positive elements in A is denoted by A +. Note that all elements of the form a 1 A 1 A 1 + + a n A na n, with A j A and a j 0 in C, are positive. These elements are called algebraically positive and they form a set denoted by A ++. As an example of such - algebras, consider A = C (M), the algebra of complex-valued smooth functions on a manifold M with -involution given by complex conjugation. In this case, with these definitions, positive functionals correspond to positive Borel measures with compact support and positive elements in A are positive functions as expected, see e.g. [9, App. B]. We remark that all these notions also make sense for star-product algebras (C (M)[[λ]], ) by means of the order structure of R[[λ]], see [6, 10]. 2.2 Formal deformations of -algebras and projections We shall discuss here some aspects of the formal deformation theory of associative algebras and -algebras over C. A good part of it is well-known and the reader is refered to [10, 20] for more details. Let k be a commutative and unital ring and let A be a k-algebra. A formal deformation of A (in the sense of Gerstenhaber, see e.g. [20]) is an associative k[[λ]]-bilinear multiplication on A[[λ]] of the form A A = C r (A,A )λ r, A,A A, (2.1) r=0 3

where each C r is a Hochshild 2-cochain and C 0 : A A A is the original product on A. If A is unital, we require in addition that the unit element 1 A is still the unit element with respect to. We denote the deformed algebra by A = (A[[λ]], ). We recall that two deformations of A, A 1 = (A[[λ]], 1 ) and A 2 = (A[[λ]], 2 ), are called equivalent if there exist k-linear maps T r : A A, r 1, so that T = id + r=1 T rλ r : A 1 A 2 satisfies A 1 A = T 1 (T(A) 2 T(A )), A,A A[[λ]]. (2.2) Let us now consider A to be a -algebra over C = R(i), R ordered. (For simplicity, whenever we refer to a -algebra A, it will be implicitly assumed that the underlying ring k is C.) A formal deformation A = (A[[λ]], ) of A is called Hermitian if the natural extension of the -involution in A to A[[λ]] is still an involution with respect to, i.e. (A A ) = A A for all A,A A. In this paper, deformations of -algebras will be always assumed to be Hermitian. We recall that if A 1, A 2 are two Hermitian deformations of A which are equivalent then there actually exists an equivalence T satisfying, in addition, T(A ) = T(A), for all A A (see [33, Prop. 5.6]). So equivalence transformations between Hermitian deformations will be assumed to preserve the involution. We observe that if A is a k-algebra and A is a formal deformation of A, then M n (A) can be identified with M n (A)[[λ]] as a k[[λ]]-module and naturally defines a deformation of M n (A). Also note that if A is a -algebra, we can define a -involution on M n (A) in the usual way and if A is a Hermitian deformation of A, M n (A) naturally defines a Hermitian deformation of M n (A). Lemma 2.1 Let A be a Hermitian deformation of a unital -algebra A over C. Let L 0 M n (A) be invertible and let S = r=0 S rλ r M n (A) be Hermitian with S 0 = L 0 L 0. Then there exist L r M n (A),r 1, such that L = r=0 L rλ r satisfies S = L L. Proof: We define L recursively. Suppose L 0, L 1,..., L k 1 M n (A) are such that L k 1 = L 0 + L 1 λ +... + L k 1 λ k 1 satisfies S L k 1 L k 1 = b k λ k + o(λ k+1 ). Note that since S is Hermitian, so is b k. We need to find L k so that L k = k j=0 L jλ j satisfies S = L k L k up to order λ k+1. But this happens if and only if L k L 0 + L 0L k = b k. Then L k = 1 2 (b kl 1 0 ) is a solution. Corollary 2.2 Let A be a unital -algebra over C and A a Hermitian deformation of A. Then any unitary U 0 M n (A) can be deformed into a unitary U = r=0 U rλ r M n (A). Let A be a k-algebra and let P 0 M n (A) be an idempotent, i.e. P0 2 = P 0. It is well-kown that if A is any formal deformation of A, we can always deform P 0, that is, find an idempotent P M n (A) so that P = P 0 +o(λ) (see e.g. [16, 18, 20]). In particular, the explicit formula (e.g. [18, Eq. (6.1.4)]) P = 1 ( 2 + P 0 1 ) 1 (2.3) 2 1 + 4(P0 P 0 P 0 ) shows that, in the case A is a -algebra and A is a Hermitian deformation, P can be chosen to be a projection (i.e. a Hermitian idempotent) if P 0 is a projection. Lemma 2.3 Let A be a k-algebra and suppose P 0,Q 0 M n (A) are idempotents. Let A be a deformation of A and P = r=0 P rλ r,q = r=0 Q rλ r M n (A) be deformations of P 0,Q 0, respectively. Then the map I : P 0 M n (A)Q 0 [[λ]] P M n (A) Q given by is a k[[λ]]-module isomorphism. I(P 0 LQ 0 ) = P (P 0 LQ 0 ) Q, L M n (A)[[λ]], (2.4) 4

Proof: The k[[λ]]-linearity and the injectivity of I are obvious since P and Q are deformations of P 0 and Q 0. To prove surjectivity, let L = r=0 L rλ r P M n (A) Q be given. Then P L Q = L whence P 0 L 0 Q 0 = L 0. Thus defining S 0 := L 0 P 0 M n (A)Q 0, we have I(S 0 ) = L up to order λ 0. Since I(S 0 ) L P M n (A) Q starts with order λ, we can repeat the argument to find a S 1 P 0 M n (A)Q 0 such that I(S 0 + λs 1 ) coincides with L up to order λ. Then a simple induction proves that I is onto. Keeping the notation as in Lemma 2.3, we denote the deformed product of M n (A) by L S = r=0 C r(l,s)λ r, L,S M n (A). Note that if we consider I : P 0 M n (A)Q 0 [[λ]] M n (A) = M n (A)[[λ]], we can write I = r=0 I rλ r and a simple computation shows that I r (B) = C m (C k (P i,b),q j ), for B P 0 M n (A)Q 0. (2.5) i+j+k+m=r We note that I is just a deformation of the natural inclusion P 0 M n (A)Q 0 M n (A). Let us consider a k-algebra A, P 0 M n (A) an idempotent, A = (A[[λ]], ) a formal deformation of A and P a deformation of P 0. It is clear that P 0 M n (A)P 0 and P M n (A) P are algebras, which are unital if A is unital. Note that if A is a -algebra, P 0 M n (A) is a projection, and A is a Hermitian deformation of A, with P a projection deforming P 0, then P 0 M n (A)P 0 and P M n (A) P are in fact -algebras. From Lemma 2.3 we find Corollary 2.4 The map I : P 0 M n (A)P 0 [[λ]] P M n (A) P defined in (2.4) induces a formal deformation of P 0 M n (A)P 0. Moreover, if A is a -algebra, P 0 is a projection and A is Hermitian, then I induces a Hermitian deformation of P 0 M n (A)P 0. 2.3 Deformations of projective inner-product A-modules Let A be a unital k-algebra and let E be a module over A. We will essentially restrict ourselves to right modules, but the reader will have no problem to adapt all the definitons and results to come to left modules. Let R 0 : E A E denote the right A-action on E, R 0 (x,a) = x A for x E, A A. Let A = (A[[λ]], ) be a formal deformation of A and suppose there exist k-bilinear maps R r : E A E, for r 1, such that the map R = R r λ r : E[[λ]] A E[[λ]] (2.6) r=0 makes E[[λ]] into a module over A. We will denote R(x,A) = x A, for x E, A A. Definition 2.5 We call E = (E[[λ]], ) a deformation of the (right) A-module E corresponding to A = (A[[λ]], ). Two deformations E = (E[[λ]], ), E = (E[[λ]], ) are equivalent if there exists an A-module isomorphism T : E E of the form T = id + r=1 T rλ r, with k-linear maps T r : E E. Here we will be only interested in finitely generated projective modules (f.g.p.m.). Proposition 2.6 Let A be a unital k-algebra and A = (A[[λ]], ) be a deformation of A. Let E be a (right) f.g.p.m. over A. Then there exists a deformation E of E corresponding to A so that E is a f.g.p.m. over A. Moreover, this deformation is unique up to equivalence and hence, in particular, every deformation of E is finitely generated and projective. 5

Proof: For the existence, note that since E is f.g.p.m., it follows that we can identify E = P 0 A n, for some n 1 and P 0 M n (A) idempotent. Let P M n (A) be an idempotent deforming P 0 and consider the (right) f.g.p.m. over A given by P A n. By Lemma 2.3 (choosing Q 0 to be 1 in the upper right corner and zero elsewhere), we can use the isomorphism I : E[[λ]] P A n to pull this A-module structure back to E[[λ]], i.e. x A := I 1 (P x A) = x A + o(λ) for x E and A A. So E = (E[[λ]], ) is a f.g.p. deformation of E. Now assume E = (E[[λ]], ) is another deformation of E. Let C : E E/λE = E and C : E E /λe = E be the natural projections, which are surjective A-module homomorphisms. Then it follows by projectivity of E that there exists an A-module homomorphism T : E E satisfying C T = C. Since E = E = E[[λ]] as k[[λ]]-modules, we can write T = r=0 T rλ r and it is readly seen that T 0 = id. So T is an equivalence. Suppose now A is a -algebra. A (right) f.g.p.m. E over A is called an inner-product module if it is equipped with a positive definite, A-valued, and A-right linear inner product, i.e. with a map h 0 : E E A satisfying h 0 (x,αy + βz) = αh 0 (x,y) + βh 0 (x,z), h 0 (x,y) = h 0 (y,x), h 0 (x,y A) = h 0 (x,y)a, h 0 (x,x) A +, and h 0 (x,x) = 0 = x = 0, where x,y,z E, α,β C, and A A. A left inner-product module is defined analogously but h 0 is required to be C-linear and A-left linear in the first argument. The A-right linear endomorphisms of E are denoted by End A (E) and the inner product determines the subalgebra B A (E) End A (E) of those endomorphisms which have an adjoint with respect to h 0. Then B A (E) becomes a -algebra over C. Let A be a Hermitian deformation of A. Let E = (E[[λ]], ) be a corresponding deformation of E and suppose there exist h r : E E A such that h = h r λ r (2.7) defines a positive definite, A-valued, A-right linear inner product on E[[λ]]. r=0 Definition 2.7 We call E = (E[[λ]],, h) a Hermitian deformation of the inner-product module (E,h 0 ) corresponding to A. Two Hermitian deformations E = (E[[λ]],,h), E = (E[[λ]],,h ) are called equivalent if there is an equivalence T = id + r=0 T rλ r : E E (as in Definition 2.5) satisfying h (T(x),T(y)) = h(x,y), x,y E. Let A be a unital -algebra and let P 0 M n (A) be a projection. Consider the f.g.p.m. over A given by E = P 0 A n. We observe that E has a canonical A-valued inner product h 0, namely the restriction to P 0 A n of the canonical A-valued inner product on the free-module A n given by x,y = n i=1 x i y i. Note that in this case M n (A) = End A (A n ) = B A (A n ) with the above -involution, the same holding for P 0 M n (A)P 0 = End A (P 0 A n ) = B A (P 0 A n ). Proposition 2.8 Let A be a unital -algebra and P 0 M n (A) be a projection. Let A be a Hermitian deformation of A and consider the A-module E = P 0 A n, equipped with its canonical A-valued inner product h 0. Then there exists a Hermitian deformation of E corresponding to A, which is unique up to equivalence. Proof: As in Proposition 2.6, we choose P M n (A), a projection deforming P 0, and consider the A- module P A n, which we know to define a deformation E = (E[[λ]], ) of E. Let h be the A-valued inner product on E obtained from, restricted to P A n. A simple computation shows that h is a deformation of h 0 and hence E = (P A n, h) is a Hermitian deformation of (E, h 0 ). Let E = (E[[λ]],, h ) be another Hermitian deformation of (E, h 0 ). By Proposition 2.6, we may assume that E = E as an A-module, with some A-valued inner product h deforming h 0. Recall that any A-valued 6

inner product, on the free A-module A n can be written as, =, H for some Hermitian element H M n (A), where x, y = n i=1 x i y i, x, y A n. Since P A n A n is projective, one can check that the same holds for this submodule. Hence, there is a Hermitian element H End A (E) so that h (, ) = h(, H ). But since h and h are deformations of h 0, we can write H = r=0 H rλ r with H 0 = id. It then follows from Lemma 2.1 that we can find U = id + r=0 U rλ r so that H = U U. It is then clear that U : E E is the desired equivalence. Let us finally discuss the deformation of isometries. An element V 0 End A (E), with E = P 0 A n, is called an isometry of E if V 0 is invertible and h 0 (V 0 x,v 0 y) = h 0 (x,y) for all x,y E. Clearly such a V 0 gives rise to a unitary Ṽ0 M n (A) with Ṽ0P 0 = P 0 Ṽ 0 and every such unitary yields an isometry of E by restriction. Proposition 2.9 Let A be a Hermitian deformation of A and E = (E[[λ]],,h) a Hermitian deformation of E = P 0 A n. Then for every isometry V 0 End A (E) there exists a deformation V = V 0 + r=1 λr V r of V 0 into an isometry V of E. Proof: Since all deformations of E are equivalent we choose a deformation P of the projection P 0. Moreover, we choose a unitary Ũ M n(a) with Ũ = r=0 λr Ũ r and Ũ0 = Ṽ0 which is possible due to Lemma 2.1. Now P = Ũ P Ũ is again a projection with classical limit P 0. Thus it defines a deformation (E[[λ]],, h ) which is equivalent to the first one by an equivalence transformation T = id + r=1 λr T r. Moreover, Ũ descends to a unitary A-right linear map U : (E[[λ]],, h) (E[[λ]],, h ) with lowest order U 0 = V 0. Then V = T 1 U is the desired deformation of V 0. 3 Deformation quantization of Hermitian vector bundles 3.1 Deformation quantization Let A = C (M) be the algebra of complex-valued smooth functions on a manifold M. This algebra has a natural -involution given by complex conjugation. Let E M be a complex vector bundle over M, with fiber dimension k 1. Consider E = Γ (E), the space of smooth sections of E, equipped with its natural right A-module structure. We recall that E is a f.g.p.m. over A (see [26, Ch. I, Thm. 6.5], noticing that the proof there works for any manifold due to [30, Lem. 2.7]). Finally, let B = End A (E) = Γ (End(E)) be the complex algebra of smooth sections of the endomorphism bundle End(E) M. Note that if E is Hermitian, i.e. equipped with a Hermitian fiber metric h 0, then there is a corresponding positive definite A-valued, A-right linear inner product on E, also denoted by h 0. This defines a -involution on B. Let A = (C (M)[[λ]], ) be a deformation of A. Recall that a star-product, given by f g = r=0 C r(f,g)λ r, is called local/differential/of Vey type if each C r is local/differential/differential of order r in each argument. Definition 3.1 Let A = (C (M)[[λ]], ) be a deformation of A. A deformation quantization of E is a deformation of E = Γ (E) in the sense of Definition 2.5. If E is equipped with a Hermitian fiber metric h 0 and A is a Hemitian deformation of A, then a Hermitian deformation quantization of (E,h 0 ) is a deformation of (E,h 0 ) as in Definition 2.7. A deformation is called local/differential/of Vey type if the corresponding R r and h r (as in (2.6), (2.7)) are local/differential/differential of order r in each argument. Recall that any two Hermitian metrics on a complex vector bundle are equivalent (see [26, Ch. I,Thm. 8.8]), hence we can identify E with P 0 A n, for some n 1 and some projection P 0 M n (A), equipped with its canonical A-valued inner product h 0. 7

Theorem 3.2 Let E be a complex (Hermitian) vector bundle over a Poisson manifold M and let A = (C (M)[[λ]], ) be a (Hermitian) deformation quantization of A. Then there exists a (Hermitian) deformation quantization of E corresponding to A, which is unique up to equivalence. Moreover, if is local/differential/of Vey type, the deformation of E can be chosen to be of the same type. Proof: Existence and uniqueness of (Hermitian) deformations follow from Propositions 2.6, 2.8 and the observation before the theorem. Suppose now is local/differential/of Vey type. Choose a deformation P of P 0 and let us consider the A-action on E[[λ]] induced by I as in (2.4). Note that if we write I = r=0 I rλ r, it follows from (2.5) that each I r : E M n (A) is local/differential/differential of order r. Moreover, I 1 = r=0 J rλ r has the same property. From x A = I 1 (P x A) and h(x, y) = P x, P y, it follows directly that R r and h r have the same desired properties. It is well-known that the complex algebras A = C (M) and B = Γ (End(E)) are Morita equivalent and hence, as such, have the same algebraic deformation theory (see [20, Sect. 16]). We will now observe that these algebras in fact have the same local and Hermitian deformation theories. See [28] for the case of local deformations of a trivial bundle. Let M loc (B) = {local deformations of B}, i.e. elements in M loc (B) are local star-products on B[[λ]]. Suppose 1, 2 M loc (B) are equivalent, through an equivalence transformation T = id + r=1 T rλ r. We call T local if each T r is local. We remark that the argument in [13, Lem. 1.1.4] shows that any equivalence transformation between local star-products is automatically local. We define Def loc (B) as the quotient of M loc (B) by (local) equivalences. Similarly, we define Def loc (B) to be the set of local Hermitian deformations of B, up to equivalence. We denote the equivalence class of in Def ( ) (B) by [ ]. loc Recall that if we identify Γ (E) = P 0 A n, where P 0 M n (A) is an idempotent (projection, in the Hermitian case), we can write B = P 0 M n (A)P 0. Let P M n (A) be an idempotent (projection) deforming P 0. Note that, by Corollary 2.4, the algebra P M n (A) P defines a (Hermitian) deformation of B via I. Moreover, it follows from (2.5) that, if is local (or differential/of Vey type), then so is the deformation of B defined by I. (It would be interesting to compare this construction of star-products on B with Fedosov s construction [17, Sect. 7] in the symplectic case.) Note also that if P is another deformation of P 0, then P M n (A) P and P M n (A) P induce equivalent deformations of B. To see that, recall that P M n (A) P = End A (P A n ) and P M n (A) P = End A (P A n ) and P A n and P A n are equivalent as A-modules by Proposition 2.6. It is simple to check that this gives rise to a well-defined map Φ : Def ( ) loc (A) Def ( ) loc (B). Proposition 3.3 The map Φ : Def ( ) loc (C (M)) Def ( ) loc (Γ (End(E))) is a bijection. Proof: As we have remarked, the algebras A = C (M) and B = Γ (End(E)) are Morita equivalent, and E = Γ (E) = P 0 A n is a (B-A)-bimodule defining this equivalence. By symmetry of Morita equivalence, it follows that there exists an idempotent Q 0 M m (B), for some m 1, so that E = B m Q 0 (B m as row vectors) as a left B-module and A = End B (B m Q 0 ) = Q 0 M n (B)Q 0 as a unital algebra. In the Hermitian case, we note that, by [25, Thm. 26], we can actually choose Q 0 satisfying Q 0 = Q 0. Note that since A is commutative (and so is Q 0 M n (B)Q 0 ), it follows from [9, Cor. 7.7], [1, Thm. 4.2] that in fact A and Q 0 M n (B)Q 0 are isomorphic as -algebras (these algebras are Morita -equivalent in the sense of [1]). Therefore, we can define a map ˆΦ : Def ( ) loc (B) Def( ) loc (A) just as we did for Φ. Let [ ] Def ( ) loc (A) and A = (A[[λ]], ). Let P be a deformation of P 0 and B = (B[[λ]], ) be the deformation induced by P M n (A) P, in such a way that [ ] = Φ([ ]). Let E = P A n, which is naturally a (B-A)-bimodule. Note that, by Morita theory, we have A = End B (E). Now pick Q M n (B), 8

a deformation of Q 0 and let E = B m Q. Then [ ] = ˆΦ([ ]) = ˆΦ Φ([ ]) is induced by Q M m (B) Q, and A = (A[[λ]], ) can be identified with End B (E ). Finally, it is not hard to check (see the existence part of Proposition 2.6) that E and E are both deformations of E = P 0 A n = B m Q 0 corresponding to B. It then follows from Proposition 2.6 (for left modules) that these deformations are equivalent and hence so are and. Therefore ˆΦ Φ = id and a similar argument shows that Φ ˆΦ = id. This concludes the proof. 3.2 The semi-classical limit We shall now compute the first order term of the deformed module E = (E[[λ]], ) over A = (A[[λ]], ), where A is a unital -algebra, a Hermitian deformation and E = P 0 A n for P 0 M n (A) a projection. The deformed module structure is defined via a deformation P M n (A) of P 0 and the isomorphism I : P 0 A n [[λ]] P A n as in Lemma 2.3. A simple computation yields R 1 (x,a) = P 0 C 1 (x,a) (3.1) for the first order term of, where C 1 (x,a) is defined by C 1 (x,a) i = C 1 (x i,a) for x A n and A A. In particular R 1 does not depend on the chosen deformation of P 0. Let us assume that A is commutative. It is well-known that in this case the skew-symmetric part of C 1 is a Poisson bracket for A. In order to get a real Poisson bracket, we use the convention {A 1,A 2 } := 1 i (C 1(A 1,A 2 ) C 1 (A 2,A 1 )), (3.2) for A 1,A 2 A. Thus {A 1,A 2 } = {A 1,A 2 } follows from the fact that is a Hermitian deformation. Let us assume furthermore that C 1 is skew-symmetric (this yelds no loss of generality for starproducts since any differential cocycle C 1 in this case is cohomologous to its skew-symmetric part [40]). Then C 1 (A 1,A 2 ) = i 2 {A 1,A 2 }. Let us consider the bracket {, } E : E A E given by obtained from the semi-classical limit of. {x,a} E = P 0 {x,a} = 2 i R 1(x,A), x E,A A (3.3) Proposition 3.4 The bracket {, } E defines the structure of a right Poisson module in the sense of [35, Def. 3.1] on E. It is a Poisson module in the sense of [35, Def. 3.2] if and only if the curvature R E : A A End A (E), defined by vanishes. R E (A 1,A 2 )x = {{x,a 1 } E,A 2 } E {{x,a 2 } E,A 1 } E {x, {A 1,A 2 }} E, (3.4) Proof: From (3.3) it easily follows that {, } E satisfies the natural Leibniz rules (see the last two equations in [35, Eq. (16)]) and this implies the first statement. A simple computation shows that R E (A 1, A 2 ) End A (E). Note that R E = 0 is equivalent to the first equation in [35, Eq. (16)] and the last statement follows directly from [35, Def. 3.2] Suppose A = C (M), where M is a Poisson manifold. Let P 0 M n (A) be a projection and let E M be given by the image of P 0, so that Γ (E) = P 0 A n. Let d denote the natural flat connection defined on the trivial vector bundle M C n M, given by component-wise exterior differentiation. Then := P 0 d defines a connection on E, sometimes called the Levi-Civita connection of E. 9

Corollary 3.5 Let E = Γ (E) and {, } E be as in (3.3). Then (E, {, } E ) is a Poisson module (in which case E is called a Poisson vector bundle) if and only if the Levi-Civita connection is flat on each symplectic leaf of M. Proof: Note that if x Γ (E) and f C (M), we have {x, f} E = Xf x, where X f is the Hamiltonian vector field corresponding to f. Observe that the curvature tensor corresponding to, R, satisfies R E (f, g)x = R (X f, X g )x, for all f, g C (M) and x Γ (E). This implies the result. We note that {, } E defines a linear contravariant connection on E M (by D df x = {x,f} E ), which is just the one induced by [19, Sect. 2]. For a Hermitian star-product on M, let be the corresponding deformation of Γ (End(E)) = P 0 M n (A)P 0 induced by a deformation P of P 0 and the isomorphism I : P 0 M n (A)P 0 [[λ]] P M n (A) P as in Corollary 2.4. Note that the center of Γ (End(E)), denoted by Z, is isomorphic to C (M) through f fp 0. As discussed in [35, Prop. 1.2], the skew-symmetric part of the semiclassical limit of endows the pair (Γ (End(E)),Z) with the structure of a Poisson fibred algebra (see [35, Def. 1.1]). For L,S M n (A), let {L,S} = 1 i (C 1(L,S) C 1 (S,L)), where C 1 (L,S) M n (A) is defined by C 1 (L,S) i,j = n r=1 C 1(L i,r,s r,j ). Proposition 3.6 The Poisson fibred algebra bracket {, } : Γ (End(E)) Z Γ (End(E)) induced by (Γ (End(E))[[λ]], ) is given by {L 0,u} = P 0 {L 0,u}P 0, for L 0 Γ (End(E)) and u Z. Moreover, the Poisson bracket defined by {, } on Z coincides with the original Poisson bracket {, } on C (M). Proof: For L 0, S 0 P 0 M n (A)P 0 = Γ (End(E)), write L 0 S 0 = I 1 (I(L 0 ) I(S 0 )) = r=0 B r(l 0, S 0 )λ r. It is not hard to check that B 1 (L 0, S 0 ) = P 0 C 1 (L 0, S 0 )P 0. (3.5) It is then clear that {L 0, S 0 } := 1 i (B 1(L 0, S 0 ) B 1 (S 0, L 0 )) = P 0 {L 0, S 0 }P 0. The bracket {, } defines an action of the Poisson algebra (Z, {, } ) on Γ (End(E)) by derivations and this implies that P 0 {P 0, }P 0 = P 0 {, P 0 }P 0 = 0. Hence the Leibniz rule for {, } yields {fp 0, gp 0 } = P 0 {f, g}p 0 = {f, g}p 0, f, g C (M), and this immediately shows that the bracket {, } on Z coincides with {, } after the identification Z = C (M). 4 Strongly full projections and formal Morita equivalence It is known that unital algebras which are Morita equivalent have equivalent algebraic deformation theory and, moreover, corresponding deformations are again Morita equivalent (see [20, Sect. 16]). We will show in this section that local, Hermitian deformations which are related by Φ are actually formally Morita equivalent, which is a notion stronger than the classical Morita equivalence and related to strong Morita equivalence of C -algebras (see [36]). We now briefly recall the definitions, see [9, 11] for details. Let A,B be -algebras over C = R(i), where R is an ordered ring. Consider a (B-A)-bimodule E with an A-valued, A-right linear positive semi-definite, full inner product, as well as a B- valued, B-left linear, positive semi-definite, full inner product Θ, such that x,by = B x,y, Θ x,ya = Θ xa,y, and Θ x,y z = x y,z for all A A, B B, and x,y,z E. Here positivity is understood in the sense of positive algebra elements of a -algebra (see Section 2.1). Fullness means E,E = A and Θ E,E = B, respectively. Let (π,h) be a -representation of A on a pre- Hilbert space over C and consider the A-balanced tensor product E A H with the inner product 10

x ψ, y φ = ψ, π( x, y )φ. If this inner product is positive semi-definite for all (π, H) then E is said to satisfy property (P). The analogous property for -representations of B is called (Q). If E satisfies all these requirements then E is called an equivalence bimodule and A and B are called formally Morita equivalent. If in addition the actions of A and B on E are non-degenerate then E is called non-degenerate. Note that we are dealing here with unital -algebras only. We remark that this purely algebraic notion is equivalent to strong Morita equivalence when applied to C -algebras, see [11]. Let A be a -algebra over C and P 0 M n (A) be a projection. Let E = P 0 A n, considered as a right A-module and left End A (E) = P 0 M n (A)P 0 module. Note that both actions are nondegenerate. We noted in Section 2.3 that E has a canonical A-valued inner product h 0. Consider the map Θ, : E E End A (E), defined by Θ x,y z = xh 0 (y,z), for x,y,z E. The following definition will give a sufficient condition to guarantee that E, equipped with h 0 and Θ,, is an equivalence bimodule. Definition 4.1 A projection P 0 M n (A) is called strongly full if there exists an invertible element τ A such that tr P 0 = (ττ ) 1. It turns out that this is indeed a stronger version of the usual notion of full projections (recall that P 0 M n (A) is full if M n (A)P 0 M n (A) = M n (A)). Theorem 4.2 Let P 0 M n (A) be a strongly full projection. Then E is a non-degenerate equivalence bimodule and thus End A (E) and A are formally Morita equivalent. Proof: Let x, y E A n. Then a straightforward computation shows the relation Θ x,p0e iτθ P0e iτ,y = Θ x,y. (4.1) i It follows immediately that Θ x,x is a positive algebra element and that Θ, is full. The fullness of, follows from P 0 e i τ, P 0 e i τ = τ tr P 0 τ = 1. (4.2) i We observe that property (P) can be easily shown as in [9, Sect. 6]. In order to prove (Q), we recall that this just means (P) for the complex-conjugate bimodule E. Equivalently, we can consider -representations of End A (E) on pre-hilbert spaces from the right. Thus let (π, H) be such a -representation of End A (E) from the right and let φ 1,..., φ r H as well as x 1,..., x r E. Then by (4.1) φi π ( Θ xi,x j ), φj = φi π (Θ xi,p 0e k τ),φ j π ( ) Θ xj,p 0e k τ 0 i,j φ i x i, φ j x j = i,j i,j,k and this shows the positivity needed for (Q). This concludes the proof. We remark that, by Lemma 2.1, it immediately follows that deformations of strongly full projections are again strongly full. Proposition 4.3 Let A be a Hermitian deformation of a -algebra A and P 0 M n (A) be a strongly full projection. Then every deformation P M n (A) of P 0 is again strongly full and therefore End A (P A n ) and A are formally Morita equivalent via the non-degenerate equivalence bimodule E = P A n. Note that if A = C (M) and P 0 M n (A) is a nowhere zero projection, then tr P 0 is nonzero and constant on connected components of M. Hence P 0 is strongly full. 11

Theorem 4.4 Let E M be a (non-zero) Hermitian vector bundle over a Poisson manifold M with star-product. Then every deformation E of E = Γ (E) is a non-degenerate equivalence bimodule and therefore End A (E) and A are formally Morita equivalent. Corollary 4.5 Consider the bijective map Φ : Def loc (C (M)) Def loc (Γ (End(E))) as in Proposition 3.3. If A = (C (M)[[λ]], ) and B = (Γ (End(E))[[λ]], ) are Hermitian deformations such that [ ] = Φ([ ]), then A and B are formally Morita equivalent. The formal Morita equivalence of C (M) and Γ (End(E)) follows from Theorem 4.4 by considering the undeformed product for the trivial Poisson bracket. This was shown in [11, Sect. 6] by a more direct argument. We also remark that formal Morita equivalence implies that the - representation theories of the involved -algebras on pre-hilbert spaces over C[[λ]] are the same, see [9, Thm. 5.10]. Acknowledgements We would like to thank Michael Artin, Martin Bordemann, Michel Cahen, Laura DeMarco, Rui L. Fernandes, Viktor Ginzburg, Simone Gutt, Tim Swift and Alan Weinstein for valuable discussions and remarks. References [1] Ara, P.: Morita equivalence for rings with involution. Algebr. Represent. Theory 2.3 (1999), 227 247. [2] Bayen, F., Flato, M., Frønsdal, C., Lichnerowicz, A., Sternheimer, D.: Deformation Theory and Quantization. Ann. Phys. 111 (1978), 61 151. [3] Bertelson, M., Cahen, M., Gutt, S.: Equivalence of Star Products. Class. Quantum Grav. 14 (1997), A93 A107. [4] Bordemann, M., Neumaier, N., Waldmann, S.: Homogeneous Fedosov Star Products on Cotangent Bundles I: Weyl and Standard Ordering with Differential Operator Representation. Commun. Math. Phys. 198 (1998), 363 396. [5] Bordemann, M., Neumaier, N., Waldmann, S.: Homogeneous Fedosov star products on cotangent bundles II: GNS representations, the WKB expansion, traces, and applications. J. Geom. Phys. 29 (1999), 199 234. [6] Bordemann, M., Waldmann, S.: Formal GNS Construction and WKB Expansion in Deformation Quantization. In: Sternheimer, D., Rawnsley, J., Gutt, S. (eds.): Deformation Theory and Symplectic Geometry, Mathematical Physics Studies no. 20, 315 319. Kluwer Academic Publisher, Dordrecht, Boston, London, 1997. [7] Bordemann, M., Waldmann, S.: Formal GNS Construction and States in Deformation Quantization. Commun. Math. Phys. 195 (1998), 549 583. [8] Brzeziński, T., Majid, S.: Quantum group gauge theory on quantum spaces. Comm. Math. Phys. 157.3 (1993), 591 638. [9] Bursztyn, H., Waldmann, S.: Algebraic Rieffel Induction, Formal Morita Equivalence and Applications to Deformation Quantization. Preprint math.qa/9912182 (December 1999). To appear in the J. Geom. Phys. [10] Bursztyn, H., Waldmann, S.: On Positive Deformations of -Algebras. Preprint math.qa/9910112 (October 1999). Contribution to the Proceedings of the Conference Moshé Flato 1999. Talk given by S. Waldmann. 12

[11] Bursztyn, H., Waldmann, S.: -Ideals and Formal Morita Equivalence of -Algebras. Preprint math.qa/0005227 (May 2000). [12] Connes, A.: Noncommutative geometry. Academic Press, Inc., San Diego, CA, 1994. [13] Deligne, P.: Déformations de l algèbre des fonctions d une variété symplectique: comparaison entre Fedosov et De Wilde, Lecomte. Selecta Math. (N.S.) 1.4 (1995), 667 697. [14] DeWilde, M., Lecomte, P. B. A.: Existence of Star-Products and of Formal Deformations of the Poisson Lie Algebra of Arbitrary Symplectic Manifolds. Lett. Math. Phys. 7 (1983), 487 496. [15] Emmrich, C., Römer, H.: Multicomponent Wentzel-Kramers-Brillouin Approximation on arbitrary Symplectic Manifolds: A Star Product Approach. J. Math. Phys. 39.7 (1998), 3530 3546. [16] Emmrich, C., Weinstein, A.: Geometry of the transport equation in multicomponent WKB approximations. Comm. Math. Phys. 176.3 (1996), 701 711. [17] Fedosov, B. V.: A Simple Geometrical Construction of Deformation Quantization. J. Diff. Geom. 40 (1994), 213 238. [18] Fedosov, B. V.: Deformation Quantization and Index Theory. Akademie Verlag, Berlin, 1996. [19] Fernandes, R.: Connections in Poisson Geometry I: Holonomy and Invariants. math.dg/0001129. [20] Gerstenhaber, M., Schack, S. D.: Algebraic Cohomology and Deformation Theory. In: Hazewinkel, M., Gerstenhaber, M. (eds.): Deformation Theory of Algebras and Structures and Applications, 13 264. Kluwer Academic Press, Dordrecht, 1988. [21] Gutt, S.: Variations on deformation quantization. Preprint ULB math.dg/0003107 (March 2000). [22] Hawkins, E.: Geometric Quantization of Vector Bundles. math.qa/9808116. [23] Hawkins, E.: Quantization of equivariant vector bundles. Comm. Math. Phys. 202.3 (1999), 517 546. [24] Jurco, B., Schupp, P., Wess, J.: Noncommutative gauge theory for Poisson manifolds. Preprint hep-th/0005005 (2000). [25] Kaplansky, I.: Rings of operators. W. A. Benjamin, Inc., New York-Amsterdam, 1968. [26] Karoubi, M.: K-theory. Springer-Verlag, Berlin, 1978. An introduction, Grundlehren der Mathematischen Wissenschaften, Band 226. [27] Kontsevich, M.: Deformation Quantization of Poisson Manifolds, I. Preprint q-alg/9709040 (September 1997). [28] Lecomte, P., Roger, C.: Formal deformations of the associative algebra of smooth matrices. Lett. Math. Phys. 15.1 (1988), 55 63. [29] Madore, J., Schraml, S., Schupp, P., Wess, J.: Gauge Theory on Noncommutative Spaces. Preprint hep-th/0001203 (2000). [30] Munkres, J. R.: Elementary differential topology. Princeton University Press, Princeton, N.J., 1963. Lectures given at Massachusetts Institute of Technology, Fall, 1961. Annals of Mathematics Studies, No. 54. [31] Nest, R., Tsygan, B.: Algebraic Index Theorem. Commun. Math. Phys. 172 (1995), 223 262. [32] Nest, R., Tsygan, B.: Algebraic Index Theorem for Families. Adv. Math. 113 (1995), 151 205. [33] Neumaier, N.: Local ν-euler Derivations and Deligne s Characteristic Class of Fedosov Star Products. Preprint Freiburg FR-THEP-99/3 math.qa/9905176 (May 1999). [34] Omori, H., Maeda, Y., Yoshioka, A.: Weyl Manifolds and Deformation Quantization. Adv. Math. 85 (1991), 224 255. [35] Reshetikhin, N., Voronov, A., Weinstein, A.: Semiquantum geometry. J. Math. Sci. 82.1 (1996), 3255 3267. Algebraic geometry, 5. 13

[36] Rieffel, M. A.: Morita equivalence for operator algebras. 285 298. Amer. Math. Soc., Providence, R.I., 1982. [37] Rieffel, M. A.: Projective modules over higher-dimensional noncommutative tori. Canad. J. Math. 40.2 (1988), 257 338. [38] Rosenberg, J.: Rigidity of K-theory under deformation quantization. q-alg/9607021. [39] Sternheimer, D.: Deformation Quantization: Twenty Years After. math.qa/9809056 (September 1998). [40] Vey, J.: Déformation du Crochet de Poisson sur une Variété symplectique. Comm. Math. Helv. 50 (1975), 421 454. [41] Waldmann, S.: Locality in GNS Representations of Deformation Quantization. Commun. Math. Phys. 210 (2000). [42] Weinstein, A.: Deformation Quantization. Séminaire Bourbaki 46ème année 789 (1994). [43] Weinstein, A., Xu, P.: Hochschild cohomology and characterisic classes for star-products. In: Khovanskij, A., Varchenko, A., Vassiliev, V. (eds.): Geometry of differential equations. Dedicated to V. I. Arnold on the occasion of his 60th birthday, 177 194. American Mathematical Society, Providence, 1998. 14