2000 Mathematics Subject Classification. 58B34; Secondary 46L65, 53D17.
<ph f="pplb7t">The Structure of Noncommutative Deformations</ph>

Eli Hawkins

Department of Mathematics The University of Western Ontario London, Ontario N6A 5B7, Canada mrmuon@mac.com

1 Introduction

The notion of a noncommutative deformation is not uncommon in contemporary mathematics. Quantum mechanics can be viewed as a noncommutative deformation of classical mechanics. Quantum groups are noncommutative deformations of Lie groups. The Podleś spheres and other quantum group homogeneous spaces are deformations. The noncommutative torus is a deformation of a torus. E   -theory is a bivariant K   -theory constructed from equivalence classes of deformations. Strict and formal deformation quantization are two mathematical settings for studying deformations.
If the starting point of a noncommutative deformation is a geometrical space, then it is natural to try to view the deformation geometrically. At best, we can try to promote a 1-parameter family of algebras to a 1-parameter family of noncommutative geometries. In [13, I showed that this is not always possible. There are obstructions to deforming integration, 1   -forms, or a metric.
Any noncommutative deformation of a smooth manifold is characterized by a Poisson structure (an antisymmetric bivector). These obstructions are expressed in the language of Poisson geometry. In particular, the necessary condition for deformation of 1   -forms is the existence of a flat, torsion-free “contravariant connection”.
The first main result (Sec. 2 ) in this paper is an obstruction to deforming higher degree differential forms. This obstruction is a tensor M l m i j k   which I call the “metacurvature” of the flat, torsion-free contravariant connection.
The most general fully fledged example of a noncommutative geometric deformation was given by Connes and Landi [8. Their construction applies to a compact Riemannian manifold with a torus group acting by isometries. An invariant Poisson structure on the torus determines the deformation. This raises the question of how much more general is the most general deformation.
Using all the available obstructions, I show in Section 6.1 that a noncommutative deformation of a compact Riemannian manifold corresponds to a Poisson structure that can be expressed locally in terms of commuting Killing vectors. Globally, the Poisson structure is induced by an invariant Poisson structure on a group of isometries of a covering space. Using this structure, I sketch a generalization of the Connes-Landi construction. This construction indicates that my obstructions are not only necessary but sufficient conditions for the existence of a noncommutative geometric deformation.
One of my obstructions is a condition for deforming integration into a trace on the noncommutative algebras. This is independent of the other obstructions, so it is possible to consider the other obstructions alone. Classifying the solutions to this weaker set of conditions is more difficult in general, but I investigate the simplest case of 2 dimensions and find that the only solutions are a flat torus with a constant Poisson bivector and a round sphere with the Poisson structure corresponding to the Podleś “standard” sphere. The latter case corresponds to a noncommutative geometric deformation constructed by Da ̧ browski and Sitarz; this example satisfies a weakened version of Connes' axioms for noncommutative geometry. In the remainder of this section, I present further background. In Section 2 , I derive the metacurvature tensor as the obstruction to the existence of a differential graded Poisson algebra; I prove a formula for the metacurvature in the simplest case of a symplectic manifold. In Section 3 , I use this formula to classify noncommutative deformations of compact 2-dimensional Riemannian manifolds, temporarily disregarding integration. In Section 4 , I reformulate the obstruction regarding integration. In Section 5 , I present several lemmata showing how the metacurvature and related structures behave in a symplectic realization. These are tools for the proofs in Section 6 where I show that the Poisson bivector for a deformation of a compact Riemannian manifold can be decomposed locally into products of Killing vectors.
Globally, I construct a symmetry group from such a Poisson structure and show how any such Poisson structure can be constructed. After some more background, in Section 7.3 I sketch the construction of a geometric deformation for any compatible Poisson structure on a compact Riemannian manifold. Finally, in Section 8 , I give a few simple examples of Poisson structures on Riemannian manifolds.

1.1 Notation

C ) ( )   will denote the space of smooth (infinitely differentiable), C   -valued functions on a smooth manifold   . Γ ( , V )   will denote the space of smooth sections of a vector bundle V   over   . Ω p ( ) : = Γ ( , p T * )   will denote the space of smooth differential p   -forms, and Ω p ( , V )   the space of smooth p   -forms with coefficients in V   .
When discussing deformations directly, it is necessary to use complex functions and sections. However, the connections and brackets that I work with all preserve real sections.
I will mostly use index-free notation for tensors. However, it is occasionally necessary to resort to index notation. In index notation, a vertical bar denotes a covariant derivative. Summation is implicit over any repeated index.
If   is a foliation of   , then T   denotes the tangent bundle to the foliation; this is a bundle over   . The differential forms along   are Ω * ( ) : = Γ ( , * T * )   .
A vector field acts on functions as a directional derivative operator as in X ( f )   .
Multivectors (vectors, bivectors, et cetera) are sections of the exterior powers * T   of the tangent bundle. I will use the symbol   to denote not only the contraction of a vector into a differential form, but also the contraction of a multivector into a differential form. This is such that, for instance, ( X Y ) ε = Y ( X ε ) = ε ( X , Y , . . . )   .
An exponent on a form or multivector always denotes an exterior power.

1.2 Lie Algebroids

Several of the principal structures that I use here are unified by the concepts of a Lie algebroid and a connection with respect to a Lie algebroid.
Definition 1.1. A Lie algebroid is a vector bundle V   over a smooth manifold   with a vector bundle homomorphism ι : V ) T   (the anchor map) and a Lie algebra structure on Γ ( , V )   such that:
  • (1) For v , w Γ ( , V )   and f C ) ( )   [ v , f w ] = f [ v , w ] + ι v ( f ) w ;  
  • (2) The anchor map intertwines the V   -bracket with the Lie bracket of vector fields [ ι v , ι w ] = ι [ v , w ] .  
The tangent bundle is itself a Lie algebroid with the identity as anchor map.
A Lie algebra is precisely a Lie algebroid over a point. The holomorphic tangent bundle of a complex manifold is a Lie algebroid. The tangent bundle to a foliation is a Lie algebroid; the anchor is the inclusion map to T   .
The most important Lie algebroid here is the cotangent bundle to a Poisson manifold. In this case, I denote the anchor as # : T * ) T   ; it is defined by ( # α ) i = π j i α j .   The bracket is the Koszul bracket which I denote as [ , ] π   ; it is uniquely defined by the defining properties of a Lie algebroid and the condition that [ d f , d g ] π = d { f , g } .   The usual notion of a connection for a vector bundle naturally generalizes to Lie algebroids.
Definition 1.2. A connection on a vector bundle W   with respect to a Lie algebroid V   is a map   from sections of V   to first order differential operators on sections of W   such that for any f C ) ( )   , v Γ ( , V )   , and w Γ ( , W )   : f v w = f v w   and v f w = f v w + ι v ( f ) w .  
The definition of curvature for such a connection is formally identical to the usual definition: K ( v , w ) = v w - w v - [ v , w ] .   It is simple to check that the definitions imply this is a tensor; specifically, it is a section of 2 V * End ( W )   . In particular, if W = V   , then we can define torsion as T ( v , w ) = v w - w v - [ v , w ] .   This is also a tensor, a section of 2 V * V   . The connections of this kind appearing in this paper are all torsion-free.
Definition 1.3. Given a foliation   , a partial connection on   is a connection on T   with respect to T   . Given a Poisson manifold, a contravariant connection [10, 19is a connection with respect to T *   ; I denote a contravariant connection as D   .

1.3 Deformations and Previous Results

Let A 0   be an algebra.
Definition 1.4. A deformation of A 0   is an algebraic extension of the form, 0 ) ħ A ) A ) P A 0 ) 0   where ħ   is a central multiplier of A   , and for a A   , ħ 2 a = 0 a ħ A .  
This definition is the weakest possible one for the purposes of this paper; it is essentially equivalent to the definition I gave in [13. One could easily replace the last assumption with the simpler but marginally more restrictive condition that ħ   is not a zero-divisor: ħ a = 0 a = 0 .   If A 0 = C ) ( )   , then we should think of A   as the algebra of smooth functions on a larger noncommutative space. This is a sort of noncommutative cobordism.
If A   is a deformation of C ) ( )   , then this definition allows us to extract a Poisson bracket from the commutator in A   . This must satisfy the Jacobi identity and is given geometrically by a Poisson bivector, π Γ ( , 2 T )   .
I now summarize my principal results from [13.
Let ε Ω n ( )   be a volume form. If integration by ε   on   can be smoothly deformed to a trace, then π   and ε   must satisfy the compatibility condition
d ( π ε ) = 0 . (1.1)
If Ω 1 ( )   and the gradient map d : C ) ( ) ) Ω 1 ( )   can be smoothly deformed, then there exists a flat, torsion-free contravariant connection D   on   .
If   has a Riemannian metric and this is extended into a deformed real spectral triple, then the above contravariant connection must be compatible with the metric in the sense that the contravariant derivative of the metric is 0   . This condition can be motivated in other ways, but it requires a sufficiently well defined notion of noncommutative geometry. In fact I will give an improved derivation if this condition in Section 6 For a given metric and Poisson structure there exists a unique torsion-free contravariant connection compatible with the metric. I call this the metric contravariant connection.
In this way, the last two conditions can be restated as: The metric contravariant connection is flat.
In the next section I will motivate and define one more condition. Given a flat, torsion-free contravariant connection, there exists a rank 5 tensor which I call the metacurvature. If there exists a deformation of differential forms in all degrees, then the metacurvature must be 0   .

2 Metacurvature

Suppose that Ω *   is a deformation of Ω * ( )   , the differential graded algebra of differential forms. Because Ω * ( )   is (graded) commutative, the graded commutator vanishes, 0 = [ σ , ρ ] : = σ ρ - ( - ) deg σ deg ρ ρ σ   for σ , ρ Ω * ( )   . For σ ^ , ρ ^ Ω *   , this implies that [ σ ^ , ρ ^ ] ħ Ω *   . Because of this we can define a generalized Poisson bracket on Ω * ( )   by { P ( σ ^ ) , P ( ρ ^ ) } = P ( 1 i ħ [ ρ ^ , σ ^ ] ) .  
Theorem 2.1. This defines a bracket on Ω * ( )   , making it a differential graded Poisson algebra.
Definition 2.1. A differential graded Poisson algebra is a graded vector space Ω *   with 3 operations: d   ,   , and { , }   , such that:
  • (1) d : Ω k ) Ω k + 1   is linear and a differential: d 2 = 0   .
  • (2) ( Ω * , )   is an associative, graded-commutative algebra.
  • (3) ( Ω * , { , } )   is a graded Lie algebra, i. e., { , }   is bilinear, degree 0   , antisymmetric
    { σ , ρ } = - ( - ) deg σ deg ρ { ρ , σ } , (2.1)
    and satisfies the graded Jacobi identity,
    { { σ , ρ } , λ } = { σ , { ρ , λ } } - ( - ) deg σ deg ρ { ρ , { σ , λ } } . (2.2)
  • (4) ( Ω * , d , )   is a differential graded algebra, i. e., it also satisfies the Leibniz identity,
    d ( σ ρ ) = d σ ρ + ( - ) deg σ σ d ρ . (2.3)
  • (5) ( Ω * , d , { , } )   is a differential graded Lie algebra, i. e., it also satisfies the Leibniz identity
    d { σ , ρ } = { d σ , ρ } + ( - ) deg σ { σ , d ρ } . (2.4)
  • (6) ( Ω * , , { , } )   is a (graded) Poisson algebra, i. e., it also satisfies the product identity
    { σ , ρ λ } = { σ , ρ } λ + ( - ) deg σ deg ρ ρ { σ , λ } . (2.5)
  • Proof. Of course, Ω * ( )   has the structure of a differential graded commutative algebra. We need to check the claims pertaining to the Poisson bracket.
    It inherits bilinearity and antisymmetry from the commutator, provided that it is well defined. Let σ ^ , ρ ^ Ω *   , σ = P ( σ ^ )   , and ρ = P ( ρ ^ )   ; we need to check that { σ , ρ } : = P ( 1 i ħ [ σ ^ , ρ ^ ] )   is uniquely determined by σ   and ρ   . The only way to change σ ^   without changing σ   is to add something of the form ħ λ ^   . However, [ σ ^ + ħ λ ^ , ρ ^ ] = [ σ ^ , ρ ^ ] + ħ [ λ ^ , ρ ^ ]   which is only changed by an element of ħ 2 Ω *   . Therefore the bracket is well-defined in σ   , and by symmetry, in ρ   .
    The graded Jacobi identity comes from the equivalent identity for the graded commutator.
In degree 0   , these structures make C ) ( )   a Poisson algebra. Hence   is a Poisson manifold.
This generalized Poisson bracket should not be confused with the Koszul bracket.
The Koszul bracket can be naturally extended to differential forms, but it is quite different from the Poisson bracket. The Koszul bracket is of degree - 1   and is uniquely determined by π   ; the Poisson bracket is of degree 0   and is not unique. For example, if α , β Ω 1 ( )   , then [ a , β ] π = - [ β , α ] π Ω 1 ( )   but { α , β } = { β , α } Ω 2 ( )   .
The product identities 2.5 for two functions and a 1   -form imply that the Poisson bracket of a function and 1   -form is given by a contravariant connection D   on T *   as { f , α } = D d f α .   The Jacobi identity 2.2 for two functions and a 1   -form imply that D   is flat:
K ( d f , d g ) α = D d f D d g α - D d g D d f α - D [ d f , d g ] π α = D d f D d g α - D d g D d f α - D d { f , g } α = { f , { g , α } } - { g , { f , α } } - { { f , g } , α } = 0 .
The Leibniz identity 2.4 for two functions implies that D   is torsion-free:
T ( d f , d g ) = D d f d g - D d g d f - [ d f , d g ] π = D d f d g - D d g d f - d { f , g } = 0 .
The contravariant connection on T *   naturally extends to the exterior powers, k T *   . This is compatible with the exterior product in the obvious way, and so the product identity 2.5 implies that the Poisson bracket of a function and differential form is given by D   as { f , σ } = D d f σ .   Now consider the identities involving the bracket of two 1   -forms. The product identity 2.5 for f C ) ( )   and α , β Ω 1 ( )   is { f α , β } = f { α , β } + α { f , β } .   The Leibniz identity 2.4 is in this case d { f , α } = { d f , α } + { f , d α } .   These identities uniquely determine the Poisson bracket of two 1   -forms. However, there is still one identity to be satisfied: the Jacobi identity,
0 = ? { f , { β , γ } } - { { f , β } , γ } - { { f , γ } , β } = D d f { β , γ } - { D d f β , γ } - { D d f γ , β } . (2.6)
Without assuming that this is satisfied, we can consider the properties of the right hand side of eq. 2.6 .
Theorem 2.2. A flat, torsion-free contravariant connection determines a tensor M l m i j k   symmetric in the contravariant indices and antisymmetric in the covariant indices, such that
M ( d f , β , γ ) = { f , { β , γ } } - { { f , β } , γ } - { { f , γ } , β } (2.7)
if M   is viewed as a trilinear map from 1   -forms to 2   -forms.
Definition 2.2. M   is the metacurvature.
  • Proof. Begin by taking eq. 2.7 as the definition of some trilinear map, and note that the right hand side of eq. 2.7 is explicitly symmetric in β   and γ   . We need to check that it is C ) ( )   -linear in either of these arguments. For any g C ) ( )   ,
    M ( d f , g β , γ ) - g M ( d f , β , γ ) = { f , g } { β , γ } - { { f , g } β , γ } + { f , β { g , γ } } - { f , β } { g , γ } - β { g , { f , γ } } = β ( - { { f , g } , γ } + { f , { g , γ } } - { g , { f , γ } } ) = 0 .
    The first two steps use the product identities; the last step uses the lower degree Jacobi identity.
    Now, consider the Jacobi identity 0 = { { f , g } , d h } - { f , { g , d h } } + { g , { f , d h } } .   Applying d   to this equation and using the Leibniz identities for d   shows that M ( d f , d g , d h ) = { f , { d g , d h } } - { { f , d g } , d h } - { { f , d h } , d g }   is symmetric under the exchange of f   and g   . Since it is C ) ( )   -linear in d g   it must also be C ) ( )   -linear in d f   . This shows that the right hand side of eq. 2.7 is indeed given by a C ) ( )   -trilinear map from 1   -forms to 2   -forms. Such a map is equivalent to a tensor M l m i j k   such that M ( α , β , γ ) l m = M l m i j k α i β j γ k   .
Theorem 2.3. The following are equivalent:
  • (1) A generalized Poisson bracket making Ω * ( )   a differential graded Poisson algebra.
  • (2) A Poisson structure on   and a flat, torsion-free contravariant connection with M = 0   .
  • Proof. We have already seen that the first structure determines the second.
    The product identity, { f σ , ρ } = f { σ , ρ } + σ D d f ρ   , shows that a generalized Poisson bracket { σ , ρ }   is first order differential in both arguments, so it can be constructed in a coordinate chart. Let x i   be the coordinates. Decomposing a differential form in these coordinates simply means writing it as a sum of products of functions and the basis 1   -forms d x i   . So, we can compute any bracket using the product identity 2.5 and the fundamental brackets:
    { f , d x i } = D d f d x i , { d x i , d x j } = d ( D d x i d x j ) ,  
    and the Poisson bracket of functions. This bracket satisfies all the product identities 2.5 because these are consistent with associativity. It satisfies the Leibniz identities 2.4 because they are consistent with the product identities.
    This leaves the Jacobi identities 2.2 . Using the coordinate decomposition and the product identities, any Jacobi identity reduces down to the Jacobi identities involving functions and 1   -forms. These are satisfied because of the assumptions that D   is flat and M = 0   . (The Jacobi identity for three 1   -forms is trivially implied by M = 0   .)
If the Poisson bivector π   is invertible, then its inverse is a symplectic 2   -form.
In that case, the flat, torsion-free, contravariant connection D   is related to a flat, torsion-free, covariant connection   by # D α β = # α # β   .
Theorem 2.4. If π   is symplectic then,
M l m i j k = D i D j D k ω l m (2.8a)
= - π a i π b j π c k ω d l ω e m π | a b c d e (2.8b)
  • Proof. Consider any point x   . Because   is flat, any vector in T x   extends to a   -constant vector field in a neighborhood of x   . Equivalently, any covector in T x *   extends to a D   -constant 1   -form in a neighborhood of x   . Because of this and symmetry, in order to compute M   at x   , it is sufficient to compute M ( α , α , α )   for any 1   -form α Ω 1 ( U )   with D α = 0   defined over some neighborhood U x   .
    Because D α = 0   , we have { f , α } = 0   for any f C ) ( )   . Using this, we compute
    { f d g , α } = f { d g , α } + d g { f , α } = f { d g , α } = f ( { d g , α } - d { g , α } ) = - f { g , d α } = - f D d g d α = - D f d g d α
    which shows that { β , α } = - D β d α   for any β Ω 1 ( U )   .
    Only the first term of 2.7 survives in
    M ( d f , α , α ) = { f , { α , α } } = D d f { α , α } = - D d f D α d α
    Because M   is just a tensor, this gives M ( α , α , α ) = - D α 2 d α .   The vector field # α   is covariantly constant ( # α = 0   ) so the Lie derivative is equal to the covariant derivative, # α = # α   . Applying this to ω   gives, # α ω = # α ω = d ( # α ω ) = d α .   The map #   naturally extends to differential forms as in # ω = - π   . Because π   is inverse to ω   , its derivative can be rewritten as # α ω = # - 1 # α π   .
    This gives the expressions,
    M ( α , α , α ) = - D α 2 # - 1 # α π (2.9)
    = - D α 3 # - 1 π = D α 3 ω . (2.10)
    This gives eq. 2.8a . To rewrite eq. 2.9 in terms of   , consider # M ( α , α , α ) = - # D α 2 # - 1 # α π = - # α 3 π   This gives eq. 2.8b .
This means that for a symplectic manifold, M = 0   if and only if π   is quadratic in the locally affine structure defined by the flat connection.
Equation 2.8a suggests an analogue of the Bianchi identity for the metacurvature.
In the symplectic case, the metacurvature is the third contravariant derivative of the symplectic 2   -form. In general, it behaves as if it is the third derivative of a 2   -form.
Proposition 2.5. D i M m n j k l   is totally symmetric in the contravariant indices.
  • Proof. It is sufficient to prove that D i M m n j k l = D j M m n i k l   .
    We can rewrite eq. 2.7 slightly as M ( d f , β , γ ) = D d f { β , γ } - { D d f β , γ } - { β , D d f γ } .   Formally, this expresses M   as the contravariant derivative of the Poisson bracket on two 1   -forms. The proof of the symmetry of the second derivative is formally the same as if { , }   were a C ) ( )   -bilinear map.

3 2 Dimensions

Suppose that a Riemannian manifold is deformed into a noncommutative geometry.
In particular, assume that differential forms are deformed in a way that is compatible with the metric. By the results of [13and Thm. 2.3 , this means that the metric contravariant connection is flat and has vanishing metacurvature, M = 0   .
Da ̧ browski and Sitarz [9have constructed an interesting example of noncommutative geometry on the Podleś “standard” sphere, a noncommutative deformation of S 2   .
This is particularly interesting because it satisfies some, but not all, of Connes' axioms for noncommutative geometry. Because of this, the homological and spectral dimensions are not 2   and there is no trace corresponding to integration on S 2   .
This example indicates that it is interesting to consider the compatibility conditions related to differential forms, but to ignore the compatibility condition related to integration. This weaker set of conditions is much more difficult to analyze in general, but in 2   dimensions it is fairly simple.
Theorem 3.1. Let   be a compact 2   -dimensional Riemannian manifold with a nonzero Poisson structure. Suppose that there exists a deformation of differential forms compatible with the metric. Then   is either a torus with a constant metric and Poisson bivector or a sphere with constant curvature and the Poisson structure corresponding to the Podleś standard sphere.
  • Proof. By assumption, the Poisson bivector π   is not identically 0   , therefore there is (at least) an open submanifold Σ : = { x | π ( x ) 0 }   . Let Σ ~   be its universal covering space.
    Where π ( x )   does not vanish, it is invertible, so Σ ~   is symplectic. Over Σ ~   , #   intertwines D   with a flat, contravariant connection. This is the Levi-Civita connection of a flat metric g i j   . The Poisson structure intertwines these metrics: g i j = π i k π j l g k l .   Because Σ ~   is two dimensional, this can be expressed more simply. Let ε Ω 2 ( Σ )   be the volume form of g   . A bivector in 2 dimensions has only one independent component, so we can write π   in the form π = h ε - 1   where h   is a scalar function. The relationship between the metrics is therefore a conformal rescaling, g i j = h - 2 g i j   .
    The flatness condition implies that there exists an isometric submersion of ( Σ ~ , g i j   to the Euclidean plane, R 2   . By Thm. 2.4 , M = 0   implies that π   is quadratic, and since the volume form ε   is constant on R 2   , h   is (the pull-back of) a quadratic function on R 2   It may be that the submersion of Σ ~   to R 2   is an isomorphism. If it is not, then there exist points of R 2   on the boundary of Σ ~   . Consider a curve of finite (Euclidean) length that ends on a boundary point x 0 R 2   . If h ( x 0 ) 0   , then the length of the lifted curve in ( Σ ~ , g )   is finite. This implies that the lifted curve has an endpoint x 1 \ Σ   . Because π ( x 1 ) = 0   , this implies that h - 1 ) 0   at the end of the curve. However, the polynomial h   is bounded along this finite curve. This contradiction shows that we must have h = 0   at all boundary points of Σ ~   .
    Because   is compact, Σ   must have finite area. This area can be computed by integrating ε = h - 2 ε   over a fundamental domain in Σ ~   . If ( Σ ~ , g )   is not isometric to R 2   , then h   would vanish at the boundary and this area integral would diverge.
    Therefore Σ ~ = R 2   .
    This also shows that h   is nonvanishing in R 2   . We can choose Cartesian coordinates x   and y   such that, h = a + b x 2 + c y 2 .   An elementary calculation shows that the curvature of g i j   is h 2 h - ( h ) 2 = 2 a ( b + c ) - 2 b ( b - c ) x 2 + 2 c ( b - c ) y 2 .   Of course, the curvature of a compact surface is bounded, so this must be bounded.
    In other words, the x 2   and y 2   terms must vanish, hence c = b   .
    There are now two possibilities. If b = 0   , then g i j   is flat. Because h = a   is constant, it does not diverge even at infinity, hence h - 1   does not vanish anywhere on   , and Σ =   .   is a flat Riemannian surface, therefore it is a torus. π = a - 1 ε - 1   is constant.
    If b 0   , then it must have the same sign as a   (so that h   does not vanish). By the above formula, the curvature is 4 a b > 0   , so   must be a sphere. The metric is d s 2 = ( a + b [ x 2 + y 2 ] ) - 2 ( d x 2 + d y 2 ) .   We can put this in a more standard form if we use the complex coordinate ζ : = ( a b ) 1 / 2 ( x + i y )   : d s 2 = ( a b ) - 1 ( 1 + ζ ζ ¯ ) - 2 d ζ d ζ ¯ .   The Poisson bracket is given by, { ζ , ζ ¯ } = 2 i { x , y } = 2 i h = 2 a i ( 1 + ζ ζ ¯ ) .   This is the Poisson structure corresponding to the Podleś standard sphere.

4 Divergence

In [13I derived d ( π ε ) = 0   as a compatibility condition between a Poisson structure and volume form. The result of this section allows this condition to be restated in terms of a contravariant connection. I present this here partly because it facilitates the simplest proof of Lem. 5.4 , but I am stating a much more general theorem than I will really need, because it may be of independent interest.
Given a contravariant connection D   and a differential form σ Ω p ( )   , we can define a “contravariant divergence” by ( D σ ) i 1 i 2 . . . i p - 1 : = D i α j i 1 . . . i p - 1   .
Theorem 4.1. Let D   be a torsion-free contravariant connection. There is a vector field φ Γ ( , T )   such that for any σ Ω * ( )  
D σ = φ σ - δ σ (4.1)
where δ σ : = π d σ - d ( π σ )   is the Koszul-Brylinski codifferential [1, 14. In particular, for α Ω 1 ( )  
D α = φ α - π d α (4.2)
and for ε Ω n ( )   a volume form,
φ ε = - d ( π ε ) (4.3)
if and only if D ε = 0   .
  • Proof. We first note that for any f , g C ) ( )   D ( f d g ) = f D d g + { g , f } .   This gives the product property D d ( f g ) = f D d g + g D d f   meaning that D d   is a first order differential operator and is equivalent to some φ Γ ( , T )   as D d f = φ ( f )   .
    Now, D ( f d g ) = φ ( f d g ) - π ( d f , d g )   which implies eq. 4.2 .
    Let σ Ω p ( )   and X Γ ( , p - 1 T )   . The contravariant exterior derivative can be expressed as - [ X , π ] = D X .   This is dual to the Koszul-Brylinski codifferential: [ X , π ] σ = ( - ) p X δ σ + δ ( X σ ) .   We can also write this in terms of the contravariant divergence
    ( D X ) σ = D ( X σ ) + ( - ) p X ( D σ )
    = - ( - ) p X ( φ σ ) - δ ( X σ ) + ( - ) p X ( D σ ) .
    From these we can solve for X ( D σ )   and prove the general result.
    The special case of a volume form ε Ω n ( )   follows from the simplification δ ε = - d ( π ε )   .
The last result here shows that if D ε = 0   then φ   is a modular vector (see [22).
The operator δ   is the boundary operator defining Poisson homology. The modular class in Poisson homology is independent of ε   . It is easy to construct D   to give φ   any desired value, but unless φ   belongs to the modular class, there cannot exist a volume form compatible with D   .

5 Realizations

Metacurvature is rather difficult to compute in general. It is only defined when D   is flat and torsion-free, but these are highly nontrivial conditions. In principle, it is possible to write an explicit formula for the metacurvature of a metric contravariant connection, but this would be hopelessly complicated. Only in the symplectic case is the metacurvature easy to compute and understand. For this reason, my strategy for analyzing the condition M = 0   on a Poisson manifold is to relate the Poisson manifold to a symplectic manifold. Fortunately, there is a well established way of doing this: symplectic realization.
Definition 5.1. A Poisson map is a diffeomorphism of Poisson manifolds φ : 1 ) 2   such that the pull-back of functions intertwines the Poisson brackets, φ * { f , g } 2 = { φ * f , φ * g } 1 .   A (local) symplectic realization of a Poisson manifold   is a Poisson map φ : ¯ )   with ¯   symplectic. A symplectic realization is full if it is a submersion.
There are variations on this definition in the literature [21, 20. It is common to assume that φ   is a surjective submersion and a complete Poisson map. However, I will not need such properties, and I do not want to assume a priori the global integrability necessary to satisfy them. Instead, I will use full local realizations as a tool for studying the local differential geometry of a Poisson manifold.
Given a symplectic realization, the preimages of points in   are the leaves of a foliation : = ker φ   . The symplectic orthogonal subbundle T : = { v T ¯ | ω ( v , w ) = 0 w T }   is integrable and thus defines another foliation,   .
The symplectic foliation S   of a Poisson manifold is defined by its tangent distribution.
The tangent fiber over x   is T x S : = Im # x T x .   In general, the symplectic foliation is only a singular foliation, but over any open region where rk π   is constant, S   is a regular foliation and T S   is actually a bundle.
Let r e g   be the union of open sets over which rk π   is constant.
Definition 5.2. An isotropic realization is a full symplectic realization such that   .
If φ : ¯ )   is an isotropic realization, then ¯   is of minimal dimension and Im φ r e g   . In this case, the symplectic leaves in   are the images of the   -leaves, and the   -leaves are precisely the preimages of the symplectic leaves.

5.1 Realizations and Connections

Recall that for a foliated manifold, a foliated chart is a coordinate chart with two types of coordinates: transverse and tangent. This is such that any leaf in the coordinate neighborhood is identified with a subset where the transverse coordinates are constant. The tangent coordinates are thus coordinates along the leaf and the transverse coordinates are coordinates on the set of leaves.
On a symplectic manifold, a contravariant connection is equivalent to a covariant connection; they are intertwined by the map #   . In general, if   is a Poisson manifold with a symplectic realization φ : ¯ )   , the composed map # φ * : Ω 1 ( ) ) Γ ( ¯ , T ¯ )   plays the role that #   does in the symplectic case. Let n = dim   and 2 N = dim ¯   . Thus dim = 2 N - n   and dim = n   .
Lemma 5.1. Let φ : ¯ )   be a full symplectic realization. If D   is a contravariant connection on   , then there exists a unique partial connection   on   such that, for α , β Ω 1 ( )   ,
# φ * ( D α β ) = # φ * α ( # φ * β ) . (5.1)
If D   is flat or torsion-free, then so is   . If D   is flat and torsion-free then around any point of ¯   , there exists an   -foliated coordinate chart such that   is simply given by partial derivatives.
  • Proof. First, observe that for α Ω 1 ( )   , the pullback φ * α Ω 1 ( ¯ )   is normal to the distribution T   . Therefore # φ * α Γ ( ¯ , T )   . Sections of this form span Γ ( ¯ , T )   as a C ) ( ¯ )   -module, therefore eq. 5.1 effectively defines X Y   for all X , Y Γ ( ¯ , T )   because 5.1 is consistent with the product rule.
    The definition of a Poisson map, and the identity d { f , g } = [ d f , d g ] π   imply that φ *   intertwines Koszul brackets. This then shows that for α , β Ω 1 ( )   [ # φ * α , # φ * β ] = # [ φ * α , φ * β ] π ¯ = # φ * [ α , β ] π .   This identity implies that the torsions and curvatures of D   and   are intertwined by # φ *   . Therefore if D   is flat or torsion-free, then so is   .
    Around any point of ¯   , there exists a neighborhood U ¯   such that the leaves of | U   are simply connected and the leaf space U /   is Hausdorff and contractible. The flat, torsion-free partial connection   is precisely equivalent to a locally affine structure on each leaf of   . Because each leaf of | U   is simply connected, it can be identified as an open subset of an affine space (of dimension n   ). These form a bundle of affine spaces over U /   . Because U /   is contractible, it can be identified with an open subset of R 2 N - n   and there exists a trivialization of the bundle of affine spaces. Together, this gives an identification of U   with an open subset of R 2 N - n × R n   . This is the desired foliated chart.
Definition 5.3. A cotangent curve [11is a curve ( u , ξ ) : R I ) T *   such that # ξ ( t ) = u ˙ ( t )   . A cotangent geodesic [10is a cotangent curve such that
D ξ ξ = 0 . (5.2)
Lemma 5.2. Let φ : ¯ )   be a full symplectic realization, and D   a contravariant connection on   . Any curve v : I ) ¯   in an   -leaf descends to a unique cotangent curve ( u , ξ ) : I ) T *   such that v ˙ ( t ) = # φ * ( ξ ( t ) )   . For any cotangent curve ( u , ξ )   , there exists such an   -curve v   , and ( u , ξ )   is a cotangent geodesic if and only if v   is a geodesic in an   -leaf.
  • Proof. The first claim is trivial; we simply define u = φ v   , and check that v ˙ ( t ) = # φ * ( ξ ( t ) )   is consistent with the definition of a cotangent curve.
    For the second claim, we must first choose a point in the preimage φ - 1 ( u ( t ) )   for some t I   . Integrating then defines v   for the rest of the interval I   .
    The third claim follows from the definition of the lifted partial connection; v ˙ v ˙ = # φ * ( D ξ ξ )   identifies the geodesic equations above and below.
For the remainder of this section, D   is a flat and torsion-free contravariant connection and φ : ¯ )   is a (local) full symplectic realization.
Definition 5.4. A flat   -foliated chart is one where the induced partial connection is given by partial derivatives. Let us say that a tensor on ¯   is   -constant (or linear, or quadratic, etc.) if in any flat   -foliated chart it is constant (linear, quadratic) in the parallel coordinates along   .
In fact, (although I won't need to prove it) a tensor is   -polynomial if this is satisfied for some   -foliated atlas. The concept really only depends upon the partial connection   . The reason is that   extends to a flat partial connection on T ¯   with respect to   , and this is unique modulo linear changes. This fact is related to the existence of a natural flat partial connection on the conormal bundle to a foliation.
Lemma 5.3. M = 0   if and only if the Poisson bivector π ¯   of ¯   is   -quadratic.
  • Proof. Around any point of ¯   , consider a flat foliated chart in a neighborhood U   . With the symplectic structure of ¯   , the trivial connection (partial derivatives) defines a flat, torsion-free contravariant connection on U   . Consider the metacurvature M ¯   of U   , and its relationship to the metacurvature M   of   .
    The definition of a Poisson map states that φ *   intertwines Poisson brackets of functions. The relationship between D   and   means that φ * : Ω * ( ) ) Ω * ( U )   intertwines the Poisson brackets of functions and 1   -forms. The product and Leibniz identities then imply that φ *   intertwines all these generalized Poisson brackets of differential forms.
    The definition of the metacurvature then shows that metacurvatures are intertwined as follows: For α , β , γ Ω 1 ( )   , φ * [ M ( α , β , γ ) ] = M ¯ ( φ * α , φ * β , φ * γ ) .   Now, assume that M = 0   . This implies that 0 = M ¯ ( α , β , γ )   for any α , β , γ   normal to   . Because U   is symplectic, we can actually compute M ¯   from the third derivative of the Poisson bivector π ¯   on U   . Equation 2.8b shows that the assumption M = 0   is equivalent to 0 = X a X b X c π ¯ , a b c d e   for all X Γ ( U , T )   . In other words, π ¯   is   -quadratic in this chart. However, since this works for any flat chart around any point, we can say that π ¯   is   -quadratic.
Lemma 5.4. If there exists ε Ω n ( )   such that d ( π ε ) = 0   and D ε = 0   , then the symplectic volume form 1 N ! ω N   is   -constant.
  • Proof. Again, around an arbitrary point of ¯   , consider a flat   -foliated chart.
    For any α Ω 1 ( )   , the pull-back of its contravariant divergence is a partial divergence in this chart: φ * ( D α ) = ( # φ * α ) .   In the notation of Thm. 4.1 , d ( π ε ) = 0   and D ε = 0   mean that φ = 0   . By eq. 4.2 , for any f C ) ( )   , D d f = 0   . Lifting this to ¯   , we have
    0 = φ * ( D d f ) = ( # φ * d f ) . (5.3)
    Now, ( # φ * d f ) ω N N ! = d ( φ * f ) ω N - 1 ( N - 1 ) !   is exact, so
    0 = d [ ( # φ * d f ) ω N N ! ]
    = ( # φ * d f ) ω N N ! + # φ * d f ω N N ! .
    The first term vanishes by eq. 5.3 . At any point of ¯   , # φ * d f   can give any vector in T   . Therefore the symplectic volume form ω N N !   is   -constant.
Corollary 5.5. If M = 0   and there exists ε Ω n ( )   with D ε = 0   and d ( π ε ) = 0   , then the symplectic form on ¯   is   -polynomial.
  • Proof. ω = π ¯ N - 1 ω N N !  
Lemma 5.6. If D   is compatible with a metric d s 2   on   , then there exists a unique flat metric d s 2   (with the same signature) on the leaves of   such that,
# φ * α , # φ * β = φ * [ α , β ] (5.4)
where ,   is the d s 2   inner product on T   and ,   is the d s 2   inner product on T *   .
  • Proof. This expression clearly defines an inner product on every fiber of T   . As above, D   determines a flat partial connection   along   . Because of the way the metrics and connections are intertwined, the   derivative of the lifted metric d s 2   vanishes.
    Along any leaf of   ,   becomes simply a (flat) connection, and d s 2   a metric.
    The connection is the Levi-Civita connection, thus the metric is flat.

6 Riemannian Manifolds

A noncommutative deformation of differential forms is characterized by a contravariant connection for which the torsion, curvature, and metacurvature vanish. In [13I showed that if a Riemannian manifold is deformed into a real spectral triple, then the contravariant connection is compatible with the metric. I will now briefly present a more robust derivation of this condition, using a much weaker notion of spectral triple. I expect that this compatibility condition can be derived from any concrete notion of noncommutative geometry.
Recall that a spectral triple consists of a Hilbert space   , an involutive algebra of bounded operators A   , and a self-adjoint unbounded operator D   , such that the commutator of D   with any element of A   is bounded.
The geometry of a Riemannian manifold   can be encoded algebraically in a spectral triple. Let A = C ) 0 ( )   . Let D   be any Dirac-type operator, and let   be the Hilbert space of square-integrable sections of the bundle on which D   acts. The metric can be recovered because of the identity ( [ D , f ] - ) 2 = - d f , d f   where f C ) 0 ( )   and ,   is the metric pairing. A spectral triple is thus a generalization of a Riemannian manifold.
Differential forms can be constructed from a spectral triple (see [5and Sec. 7.1 ). In particular, Ω D 1 ( A )   is the A   -bimodule generated by bounded operators of the form d a : = [ D , a ] -   , for a A   . In the above example, Ω D 1 ( A ) = Ω 0 1 ( )   .
Now suppose that there exists a smooth noncommutative deformation of a spectral triple describing a Riemannian manifold. Suppose that the differential forms constructed from this deformation are a smooth deformation of the differential graded algebra Ω 0 * ( )   . We have seen in Section 2 that this deformation of Ω 0 * ( )   is described to leading order by a Poisson structure and a contravariant connection D   with 0   torsion, curvature, and metacurvature.
To see how this contravariant connection is related to the metric, consider the simple identity
[ a , ( [ D , b ] - ) 2 ] - = [ a , [ D , b ] - ] - [ D , b ] - + [ D , b ] - [ a , [ D , b ] - ] - . (6.1)
A commutator such as [ a , [ D , b ] - ] -   corresponds to a generalized Poisson bracket such as { f , d h }   . At first order in ħ   , the identity 6.1 gives { f , d h , d h } = 2 d h , { f , d h } = 2 d h , D d f d h ,   but this is simply the condition that the contravariant derivative of the metric is 0   .
Since D   is torsion-free, this means that D   is (by definition) the metric contravariant connection.
Together with the compatibility with the volume form, this gives the following notion of compatibility.
Definition 6.1. For the purposes of this paper, I will say that a metric and Poisson structure are compatible if
  • (1) The metric contravariant connection D   is flat.
  • (2) The metacurvature (of D   ) vanishes: M = 0   .
  • (3) the Poisson structure is compatible with the Riemannian volume form:
    d ( π ε ) = 0 .  

6.1 Local Structure

In this subsection,   is a compact Riemannian manifold with a metric g i j   and a compatible Poisson structure π   . The analysis here is based on repeatedly using two simple principles: A continuous function on a compact manifold is bounded, and a bounded polynomial is constant.
The local norm of the Poisson bivector π   is defined by | π | 2 : = 1 2 π i j π k l g i k g j l .   This is a continuous function and   is compact, so there is also a global norm, π : = max x | π ( x ) | .   The norm of any multivector or differential form is defined analogously.
Lemma 6.1. For any local full realization of   , let ω 0 Ω 2 ( )   be the pull-back of the symplectic form to the   -leaves; this is   -constant.
  • Proof. First, consider any maximally extended cotangent geodesic ( u , ξ ) : R I ) T *   .
    The geodesic equation 5.2 implies that | ξ ( t ) |   is constant, so the “speed” of the curve u   is bounded as | u ˙ ( t ) | = | # ξ ( t ) | | π ( u ( t ) ) | | ξ | π | ξ | .   This means that if ( u , ξ )   cannot be extended beyond (say) t = t 1   , then it ends at a point x 1 : = lim t ) t 1 u ( t ) .   However, if we construct a symplectic realization over x 1   , then ( u , ξ )   lifts to a straight line in an   -leaf. There is no obstruction to extending a straight line, therefore I = R   .
    For any t 0 R   , there exists a symplectic realization φ : ¯ )   over u ( t 0 )   . By Lem. 5.6 , this defines flat Riemannian metrics on the   -leaves in ¯   . By Lem. 5.2 , ( u , ξ )   lifts to a geodesic v   in the   -leaf over u   . The pull-back of the local norm | π |   to ¯   is φ * | π | = | ω 0 | ,   the norm of ω 0   with respect to the flat metric on   -leaves. By Cor. 5.5 , ω   , and thus ω 0   and | ω 0 | 2   are   -polynomial. This means that | π ( u ( t ) ) | 2 = | ω 0 ( v ( t ) ) | 2   is a polynomial for t t 0   . Hence, | π ( u ( t ) ) | 2   is a polynomial for all t R   .
    This polynomial is bounded by π 2   , therefore it is constant. Through every point of   there are cotangent geodesics passing in any direction tangent to the symplectic foliation S   , therefore | π |   is S   -constant.
    Now, for any symplectic realization, this shows that | ω 0 | = φ * | π |   is   -constant.
    Because ω 0   is   -polynomial, this is enough to prove that ω 0   is   -constant.
Lemma 6.2. Around any regular point with 2 m = rk π   there exists an S   -foliated chart with parallel coordinates { x a } a = 1 2 m   and transverse coordinates { z α } α = 1 n - 2 m   such that:
  • (1) The components of π   are constant in this coordinate system.
  • (2) The metric is of the form
    d s 2 = g a b ( d x a - A α a d z α ) ( d x b - A β b d z β ) + g α β d z α d z β (6.2)
    with g a b   and g α β   independent of x   . A α a   is polynomial in x   , and for fixed z   and α   , A α a   are the components of a Hamiltonian vector field in the symplectic leaf.
  • Proof. Let φ : U )   be an isotropic realization over an arbitrary point x 0 r e g   , with U   small enough that the leaf space U /   is Hausdorff, connected, and simply connected. Over U   , the bundle T   decomposes canonically into the direct sum of T   (which is the kernel of ω 0   ) and its orthogonal complement. Let 2 m = rk π ( x 0 )   .
    We can construct an adapted coordinate system as follows. First, choose some smooth transversal to   (intersecting φ - 1 ( x 0 )   ) with coordinates { z α } α = 1 n - 2 m   . Choose a smooth basis of T   over the transversal such that the last n - 2 m   basis vectors lie in T   , the first 2 m   lie in the orthogonal complement, and ω 0   is constant in this basis. Extend the coordinate system to a neighborhood of the transversal using the exponential map of   . Let { x a } a = 1 2 m   be the coordinates coming from the first 2 m   basis vectors and { y α } α = 1 n - 2 m   the coordinates coming from the last n - 2 m   .

    Figure 1 . In this sketch, the planes represent   -leaves; the lines represent   -leaves.

    By construction, this is a flat   -foliated chart with transverse coordinates z   .
    It is also an   -foliated chart with transverse coordinates y   and z   . To put this another way, the z   's parameterize the set of   -leaves, the x   's parameterize the set of   -leaves in each   -leaf, and the y   's are coordinates on each   -leaf, as depicted in the figure.
    The y   - y   -components of ω   vanish because   is isotropic. The z   - z   -components vanish because   is coisotropic. The x   - y   -components vanish because (by definition)   is symplectically orthogonal to   . With this in mind, the symplectic form reduces to ω = 1 2 ω a b d x a d x b + ω a β d x a d z β + ω α β d y α d z β .   the first term is ω 0   , and therefore ω a b   is constant.
    The Poisson bivector π ¯   on U   is of course given by inverting this symplectic form.
    The Poisson bivector on Im φ   is given by the x   and y   components of π ¯   , but only the x   - x   -components may be nonzero. Thus the nonzero part of π   is π a b   which is the inverse matrix of ω a b   , but ω a b   are components of ω 0   and are thus constant. Because of the orthogonal decomposition, the metric on the   -leaves is of the form d s 2 = g a b d x a d x b + g α β d y α d y β .   These components are functions of the z   's alone.
    The metric on Im φ   is constructed by using the symplectic form to lower the indices of the contravariant form of d s 2   . This gives d s 2 = g c d ( ω c a d x a + ω c α d z α ) ( ω d b + ω d β d z β ) + g γ δ ω γ α ω δ β d z α d z β .   This is of the form 6.2 with A α a = π a b ω b α .   Because ω   is closed and ω a b   constant, ω a α d x a   is exact. Therefore A α a   is a Hamiltonian vector field in the sense claimed.
These results can be restated in purely geometric terms:
  • (1) The symplectic foliation S   is a Riemannian foliation [15of r e g   . That is, the metric descends to a well-defined transverse metric g α β   . This is a metric on the leaf space, to the extent that the leaf space is meaningful.
  • (2) The induced leaf metric g a b   is flat, thus the leaves are locally affine and in this sense the symplectic form is constant on each leaf.
  • (3) Locally, S   looks like a bundle of symplectic affine spaces. Flowing orthogonally to S   defines a connection   (for which A α a   is the potential). The structure Lie algebra is that of polynomial Hamiltonian vector fields on R 2 m   .
Corollary 6.3. The regular symplectic leaves are (intrinsically) geodesically complete.
  • Proof. We can compute | π m |   explicitly from the above coordinate expressions. It is S   -constant.
    Suppose that u : [ 0 , 1 ) )   is an incomplete geodesic in a regular 2 m   -dimensional leaf. Because   is compact, this extends continuously to u : [ 0 , 1 ] )   . The point u ( 1 )   does not lie on the leaf, thus rk π [ u ( 1 ) ] < 2 m   and | π m [ u ( 1 ) ] | = 0   . However, | π m [ u ( t ) ] |   is constant and nonzero for t [ 0 , 1 )   , contradicting the continuity of | π m |   .
The extrinsic curvature of S   is K : = 1 2 d s 2   . This is a section of T * S T * S N * S   , where N * S   is the conormal bundle to S   . This is easily computed for the metric form 6.2 : K a b α = 1 2 ( g a b , α + A a α , b + A b α , a )   where latin indices are lowered with the leaf metric g a b   . Commas denote partial derivatives.
Lemma 6.4. A α a   only depends linearly on the x   's (i. e., it is S   -linear).
  • Proof. First, consider the trace of the extrinsic curvature: ( tr K ) α : = K a α a = 1 2 g a b g a b , α + A α , a a .   The last term vanishes because A α a   is Hamiltonian. So K a α a   is constant along S   .
    (This is actually a coordinate-independent statement; the foliation determines a canonical flat partial connection of N * S   along S   .) The components R a b c d   of the Riemann tensor parallel to S   can be computed from the extrinsic curvature R a b c d = K a c α K b d α - K a d α K b c α .   The trace of this is
    g a c g b d R a b c d = tr K 2 - ( tr K ) 2 = K a b α K a b α - K a α a K b b α . (6.3)
    The last term is S   -constant. This does not necessarily extend to a continuous function beyond r e g   . However, 6.3 is bounded in terms of the norm of the Riemannian curvature of   . This means that K   must be bounded. Since A   is polynomial in x   , so is K   and it must be independent of x   .
    This shows that A a α , b + A b α , a   is independent of x   , which implies that A α a   is linear in x   .
This shows that the structure group of   reduces to I S p ( 2 m )   , the group of linear symplectomorphisms of R 2 m   . This is very much like the geometry of a Kaluza-Klein model.
The curvature of   is F α β a = A β , α a - A α , β a + A α b A β , b a - A β b A α , b a .   This should be thought of as a vector field (over R 2 m   ) valued 2   -form (over the leaf space).
Lemma 6.5. In some neighborhood of any regular point where rk π = 2 m   , there exist commuting Killing vectors { X A } A = 1 2 m   which span T S   . The Poisson bivector can be expressed as
π = 1 2 Π A B X A X B (6.4)
where the matrix Π A B   is constant and nondegenerate.
  • Proof. The scalar curvature R   of   can be computed in terms of the transverse metric, the leaf metric, and   . The only term of R   which is not necessarily S   -constant is - 1 4 F 2 : = - 1 4 g a b g α γ g β δ F α β a F γ δ b .   Because R   is bounded, F α β a   must be bounded. However, F α β a   must be polynomial in x   , therefore it is independent of x   .
    In other words, the curvature is translation valued. This means that locally, by a gauge (coordinate) transformation, A α a   can be made independent of x   . With such a coordinate choice, all components of the metric 6.2 are independent of x   . This means that the basis vectors in x   -directions are commuting Killing vectors. Let { X A } A = 1 2 m   be this basis; that is, X A a = δ A a   and X A α = 0   . The components of π   in this coordinate system become the components of π   in this basis. Renaming these as Π A B   , we have eq. 6.4 .
The final step is to show that Lem. 6.5 implies that the decomposition 6.4 exists in a neighborhood of any point, not just a regular point. To do this, I will view Killing vectors as coming from a larger bundle. If X Γ ( , T )   is a Killing vector, then X   is antisymmetric (with respect to the metric), so X   and X   together form a section of the bundle i s o ( T ) = T 2 T   ; moreover, the second derivative can be expressed in terms of X   and the Riemann tensor.
Based on this, define a connection on i s o ( T )   by ~ Y ( X , w ) : = ( Y X - w ( Y ) , Y w - R ( Y , X ) )   where Y Γ ( , T )   and ( X , w ) Γ ( , i s o ( T ) )   . If X   is a Killing vector then 0 = ~ ( X , X )   . Conversely, any ~   -constant section is given by a Killing vector in this way.
Let 2 m   be the maximum rank of π   .
Theorem 6.6. Over any simply connected open subset of   , π   is given by 2 m   commuting Killing vectors as π = 1 2 Π A B X A X B   with constant coefficients. The set { ( X A , X A ) } A = 1 2 m   is a basis of ~   -constantsections of a flat subbundle V i s o ( T )   .
  • Proof. This proof consists of repeatedly applying a simple principle: If a continuoussection vanishes over some neighborhood of any regular point, then it vanishesover r e g   and by continuity it vanishes over   . This can be applied to any propertythat can be expressed as the vanishing of a continuous section. Let x 0 r e g   be an arbitrary regular point. By Lem. 6.5 , in some neighborhood U   of x 0   , we have the decomposition 6.4 . From this, we can construct X ~ A : = ( X A , X A )   and
    π ~ = 1 2 Π A B X ~ A X ~ B (6.5)
    over U   . Because each X A   is a Killing vector, 0 = ~ X ~ A   , and so 0 = ~ π ~ .   This π ~ Γ ( U , 2 i s o ( T ) )   is made up of three components; the first if π   itself, and the others are sections of T s o ( T )   and 2 s o ( T )   . A priori, π ~   is only defined over U   , and appears to depend upon a choice ofdecomposition. However, observe that A k i j : = 1 2 ( π | k i j - π k j | i - π k i | j ) = Π A B X A i X B | k j   and B k l i j : = A l | k i j + π i m R k l m j = Π A B X A | k i X B | l j   are the other components of π ~   . Obviously, A   and B   are well defined tensors over   . Although B   is not explicitly a section of s o ( T )   , it is over U   ; by continuity, it is over   .
    So, we can define π ~ Γ ( , 2 i s o ( T ) )   as the global section with components π   , A   , and B   . This satisfies 0 = ~ π   over U   ; by continuity, this is true over   .
    This implies that π ~   has constant rank (which must be 2 m   ). It thus spans a subbundle V i s o ( T )   . Over U   , the decomposition 6.5 shows that the sections { X ~ A } A = 1 2 m   span V   . Since 0 = ~ X ~ A   , the restriction of ~   to V   is flat over U   ; by continuity, it is flat over   .
    Any ~   -constant section of V   over U   is a constant linear combination of { X ~ A } A = 1 2 m   ; its T   -component is the same linear combination of { X A } A = 1 2 m   . Therefore, the T   -components of ~   -constant sections of V   over U   are mutually commuting Killing vectors; by continuity, this is true over any domain in   .
    Now forget the X A   and X ~ A   used above. Over any simply connected neighborhood of any point in   , there exists a basis { X ~ A } A = 1 2 m   of ~   -constant sections of V   . In this basis, π ~   has constant components, Π A B   . Defining X A   as the T   -component of X ~ A   , this gives the desired decomposition of π   .

6.2 Global Structure

Theorem 6.7. Let   be a connected, compact Riemannian manifold with a compatible Poisson structure. There exists a covering ~   and a Lie group G   of isometries of ~   such that:
  • (1)   is the quotient ~ / Γ   by a discrete, cocompact subgroup Γ G   .
  • (2) The Poisson structure on ~   is induced by an Ad G   -invariant bivector Π 2 g   , where g   is the Lie algebra of G   .
  • (3) The span of Π   (in g   ) densely generates a connected abelian normal subgroup T G   .
  • (4) Γ T = { e }   and the subgroup T Γ G   generated by Γ   and T   is dense.
  • Proof. Choose some (arbitrary) base point in   . Let Γ   be the holonomy group for the flat bundle V i s o ( T )   , regarded as a discrete group. Γ   is a quotient of the fundamental group, so we can define ~   as the covering of   with covering group Γ   . Let G 1   be the (Lie) group of isometries of ~   that preserve the Poisson structure (i.e., Poisson isometries).
    By construction, Γ G 1   , = ~ / Γ   , and V   is globally flat. So, the Killing vectors X A   exist globally over ~   , and π = 1 2 Π A B X A X B   over ~   .
    Since the Killing vectors X A   commute, the decomposition shows that they preserve π   . Hence, they are elements of the Lie algebra of G 1   . Let T G 1   be the Lie subgroup densely generated by { X A } A = 1 2 m   . This is abelian because the X A   's commute.
    Because π   is G 1   -invariant, Π   is Ad G 1   -invariant and T   is normal.
    Define G   as the closure of the subgroup T Γ G 1   . Because G G 1   is a closed subgroup, it is a Lie group. By construction Γ G   and T G   .
    The definition of Γ   implies that the adjoint action of Γ   on T   is a faithful representation.
    Because T   is abelian, the vectors X A   are T   -invariant. Therefore Γ T = { e }   .
    For some regular point x ~   , let H x G Isom ( ~ )   be the subgroup leaving x   fixed. because   (and hence ~   ) is Riemannian (with positive definite metric) H x   is compact. Let O x   be the closure of the symplectic leaf through the image of x   in   . This is naturally identified with the double quotient H x \ G / Γ   . Because   is compact, O x   must be compact and so G / Γ   is compact. In other words, Γ G   is cocompact.
Note that an Ad G   -invariant bivector in 2 g   is the same thing as a bi-invariant (left and right invariant) bivector field on G   . In fact [20,Thm. 10.4a bi-invariant Poisson structure always comes from an abelian normal subgroup in this way.
Theorem 6.8. Let   be an (arbitrary) Riemannian or pseudo-Riemannian manifold.
If ~   is some covering of   , and G   is a Lie group acting by isometries on ~   such that ~ / G   is a quotient of   , then any Ad G   -invariant Poisson bivector Π 2 g   induces a Poisson structure on   which is compatible with the metric in the sense of Definition 6.1 .
  • Proof. Obviously, Π   induces a G   -invariant Poisson structure on ~   , but   is the quotient of   by a subgroup of G   , so this induces a Poisson structure on   . The compatibility conditions in question are all local, so it is sufficient to check them on ~   .
    By Theorem 10.4 of [20, Π   spans an abelian ideal t g   . So Π 2 t   . This means that the Poisson structure on ~   can be written as 1 2 Π A B X A X B   where { X A }   is a basis of commuting Killing vectors spanning t   .
    The volume condition is straightforward to check (   denotes the Lie derivative):
    d ( π ε ) = 1 2 Π A B d [ ( X A X B ) ε ) = Π A B X A ( X B ε ) - 1 2 Π A B [ X A , X B ] ε = 0
    because the vectors X A   preserve the volume form and commute.
    Some contravariant connection D   is defined by D α β = Π A B ( X A β ) X B β ,   for any α , β Ω 1 ( ~ )   . This is compatible with the metric because the X A   's are Killing vectors.
    It is sufficient to compute the torsion using exact 1   -forms. Noting that, D d f d h = Π A B X A ( f ) d [ X B ( h ) ] ,   we have
    T ( d f , d h ) = D d f d h - D d h d f - d { f , h } = Π A B [ X A ( f ) d ( X B h ) + d ( X A f ) X B ( h ) - d ( X A f X B h ) ] = 0 .
    This shows that this D   is in fact the metric contravariant connection.
    Again, let t   be the abelian Lie algebra spanned by { X A }   . Given any regular point x 0 ~   , any covector at x 0   extends to a t   -invariant 1   -form in a neighborhood of x 0   .
    Let γ   be such a t   -invariant 1   -form. Its contravariant derivative vanishes, D γ = 0   .
    Consequently, K ( α , β ) γ = 0   and so K = 0   at x 0   . The curvature vanishes at every regular point, therefore K = 0   everywhere.
    The first expression for D   generalizes to give the Poisson bracket of a function and a differential form, { f , σ } = Π A B X A ( f ) X B σ .   The Leibniz identity 2.4 then implies that the bracket of two 1   -forms must be, { α , β } = Π A B X A α X B β .   Again, let γ   be a t   -invariant 1   -form on some neighborhood of x 0   . These formulæ show that any generalized Poisson bracket of γ   with a function or 1   -form must vanish. Thus by eq. 2.7 , M ( d f , β , γ ) = 0   . This shows that M = 0   at x 0   , hence at every regular point, hence everywhere.
These two results show that in the case of a compact Riemannian manifold, a Poisson structure is compatible with the metric if and only if it is induced in this way from a bi-invariant Poisson structure on a group.
Theorem 6.7 thus shows how to construct all examples of compatible Poisson structures. They are classified by triples ( G , Γ , Π )   of a Lie group, a cocompact discrete subgroup and a bi-invariant Poisson structure, such that Γ   and the span of Π   densely generate G   .
Whenever the metacurvature obstruction vanishes, the de Rham complex of differential forms becomes a differential graded Poisson algebra. This is precisely the natural sufficient condition for the generalized Poisson brackets to descend to de Rham cohomology. It is thus natural to ask what this gives in the cases we have been considering. The answer is disappointingly trivial.
Proposition 6.9. If   is a compact Riemannian manifold with a compatible Poisson structure, then the induced Poisson bracket on de Rham cohomology is 0   .
  • Proof. Firstly, using the Leibniz identity 2.3 , we can extrapolate to an explicit formula for the generalized Poisson bracket of two differential forms σ , ρ Ω * ( ~ )   , { σ , ρ } = Π A B X A σ X B ρ .   If σ   and ρ   are closed, then this can be rewritten as
    { σ , ρ } = Π A B d ( X A σ ) d ( X B ρ ) = d [ ( X A σ ) X B ρ ] .
    The last expression in brackets is just a contraction of ρ   with D σ Ω ( ~ , T ~ )   . In this way, the formula descends to   . If σ , ρ Ω * ( )   are closed forms on   , then D σ Ω * ( , T )   and so { σ , ρ }   is exact.
There is no obvious reason for the bracket on cohomology to vanish in greater generality; this question can be considered for any Poisson manifold that has a contravariant connection with vanishing torsion, curvature, and metacurvature.
Unfortunately, the only other example that I discuss in this paper is S 2   ; in that case the bracket on cohomology vanishes simply because the cohomology is trivial.

7 Spectral Triples

I have shown that the compatibility conditions of Definition 6.1 are necessary for the existence of a deformed noncommutative geometry which respects differential forms and integration. I have not shown if these conditions are sufficient. As I shall explain, this appears to be essentially true provided that the Poisson structure is itself suitably integrable.
I have tried to be as general as possible by not tying my arguments to a specific notion of noncommutative geometry more than necessary. In order to discuss the sufficiency of my compatibility conditions, it is appropriate to be a bit more concrete.

7.1 Differential Forms

Connes [5has given a very general recipe for constructing a differential graded algebra of “noncommutative differential forms” Ω D * ( A )   from a spectral triple. Given the Dirac operator and algebra of smooth functions on a compact spin manifold, this recovers the differential forms: Ω D * [ C ) ( ) ] = Ω * ( ) .   The construction applies provided that the commutator of D   with any element of A   is bounded. We begin by building a universal differential graded algebra from A   . Let Ω 0 ( A ) = A   and Ω 1 ( A ) A A   the kernel of the multiplication map A A ) A   . The differential begins with d : Ω 0 ( A ) ) Ω 1 ( A ) , a 1 a - a 1 .   Finally, Ω k ( A )   is the k   -fold tensor product of Ω 1 ( A )   over A   , and d   is defined in general by the Leibniz identity. Note that Ω k ( A ) A ( k + 1 )   .
The formula a 0 a 1 a k a 0 D a 1 D . . . D a k   defines a map from A ( k + 1 )   to operators. Restricting this to Ω k ( A )   gives a representation by bounded operators ρ : Ω * ( A ) ) ( )   . The kernel of ρ   is neither a differential nor graded ideal.
Instead define J 0 ker ρ Ω * ( A )   as the subspace spanned by homogeneous elements; then J : = J 0 + d J 0   is a differential graded ideal. Finally, Ω D ( A ) : = Ω * ( A ) / J   .
This doesn't require a spectral triple that could be reasonably regarded geometrically.
For example, we could take D = 0   , in which case Ω 0 0 ( A ) = A   and Ω 0 * > 0 ( A ) = 0   .

7.2 Axioms

In [6, Connes presented a system of axioms for a real spectral triple. This is the most completely and restrictively defined notion of noncommutative geometry.
For reference, I summarize the axioms here. See [12for the most detailed discussion.
An n   -dimensional real spectral triple consists of:   a Hilbert space; A   a *   -algebra of bounded operators; D   an unbounded self-adjoint operator; J   an antiunitary operator; and γ   a Z 2   -grading operator. These satisfy the following axioms.
Dimension: The resolvent ( D + i ) - 1   is a compact operator contained in the ideal n + ( )   .
Smoothness: For any a A   , the commutator [ D , a ] -   is bounded. Both a   and [ D , a ] -   are in the domain of any power of the derivation δ   defined by δ ( a ) = [ | D | , a ] -   .
Reality: For any a , b A   , a o p : = J a * J - 1   commutes with b   . J 2 = ± 1   , J D = ± D J   , and J γ = ± γ J   , with the signs depending upon the dimension n   modulo 8   .
First order: For any a , b A   , a o p   commutes with [ D , b ] -   .
Finiteness: The common domain )   of all powers of D   is a finitely generated, projective A   -module. There is an A   -valued pre-Hilbert module inner product such that for a A   , ψ , χ )   , Tr ω ( a ψ | χ ) | D | - n ) = ψ | a o p | χ .   Orientation: γ   is self-adjoint and commutes with A   . If n   is even then γ D = - D γ   ; if n   is odd then γ = 1   . There exists a Hochschild cycle c Z n ( A , A A o p )   such that ρ ( c ) = γ   for the representation ρ : Z * ( A , A A o p ) ) ( )   is defined by ρ ( a 0 b o p a 1 a n ) = a 0 b o p [ D , a 1 ] - . . . [ D , a n ] - .   Poincaré duality: The Kasparov product with the K   -homology class [ D ] K R n ( A A o p )   is an isomorphism: A [ D ] : K * ( A ) ~ K * ( A o p ) = K * ( A ) .  

7.3 Converse Construction

If a noncommutative deformation of the geometry of   exists, then we must have in particular a noncommutative deformation of the algebra of smooth functions on   . Whether this exists is a more fundamental issue than compatibility with geometry, and in a way it is a separate question.
The results of Section 6.2 show that a compatible Poisson structure comes from a homogeneous Poisson manifold G / Γ   . This is the model that everything is constructed from. Assuming that G / Γ   can be suitably deformed, I will sketch the construction of real spectral triples for a deformation of   .
I will focus on real spectral triples, so it is necessary to assume that   is a spin manifold. I assume that   is a compact, Riemannian spin manifold with a Γ   -covering ~   , Γ G   is a discrete, cocompact subgroup, Π 2 g   is an invariant Poisson bivector, and the action of G   on ~   extends to the spinor bundle. In order to satisfy the last condition, we might need to replace G   with some finite covering.
A geometric deformation of   exists if there is a suitable G   -equivariant deformation of G / Γ   . I will not only assume that there exists a deformation B   of C ) ( G / Γ )   , but also:
  • (1) B   is a dense subalgebra of sections of a continuous field, B   , of C *   -algebras over an interval.
  • (2) There is an action of G   on B   which extends the action of G   on C ) ( G / Γ )   , i. e., this is an equivariant deformation.
  • (3) For every value of the parameter ħ   , let ħ B ħ   be the image of B   in B ħ   . This is precisely the domain of the action of the universal enveloping algebra of g   on B ħ   .
  • (4) There exists a G   -invariant tracial state τ ħ : B ħ ) C   .
It is elementary to construct a deformation of G   with respect to Π   . In some cases, the techniques of Rieffel [17can then be used to construct a deformation of G / Γ   .
However, it is not clear whether G / Γ   can be suitably deformed in general.

7.3.1 Algebra

In preparation for constructing the deformed algebras, we can reexpress   as = ~ / Γ = ~ × G G / Γ .   This means that the algebra of smooth functions C ) ( )   is naturally identified with the G   -invariant C ) ( G / Γ )   -valued smooth functions on ~   : C ) ( ) = C ) ( ~ , C ) [ G / Γ ] ) G .   The (right) action of g G   is defined by the pullback by the right action of g - 1   on ~   and the the left action of g   on G / Γ   .
We can mimic this construction in the noncommutative case substituting ħ   for C ) ( G / Γ )   . There is still a right action of G   on ħ   . The right action of g G   on C ) ( ~ , ħ )   is given by simultaneously applying the action of g   on ħ   and pulling back by the action of g - 1   on ~   . With this action, we can define the algebra A ħ : = C ) ( ~ , ħ ) G .   This generalizes the “twisting by a torus” construction of Connes and Landi [8(see also [7).

7.3.2 Hilbert Space

Let S )   be the spinor bundle. The A ħ   -module ħ ) : = Γ ( ~ , S ħ ) G   is an analogue of the smooth sections of S   over   , but in order to construct a spectral triple we need a Hilbert space analogous to the L 2   -sections of S   . For this, we need an inner product.
The inner product of spinors comes fundamentally from a bundle homomorphism S ¯ S ) n T * .   For two spinor sections ψ , φ Γ ( , S )   , the local inner product is ψ ¯ φ Ω n ( )   .
Integrating this over   gives the Hilbert inner product.
Now, for two G   -invariant sections ψ , φ Γ ( ~ , S ħ ) G   the local inner product is ψ ¯ φ Ω n ( ~ , ħ ) G   . If we apply the G   -invariant trace, this becomes τ ħ ( ψ ¯ φ ) Ω n ( ~ ) G .   It would be a mistake to try to integrate this over ~   , since it does not fall off at all. Instead note that Ω n ( ~ ) G Ω n ( ~ ) Γ = Ω n ( ) .   With this identification, we can integrate over   and define the Hilbert inner product as ψ | φ : = τ ħ ( ψ ¯ φ ) .   It is worth observing how this gives the correct inner product at ħ = 0   . In that case, ψ ¯ φ   can already be identified with an n   -form on   . There is a singular foliation of   by the images of the G   orbits. The trace τ 0   averages ψ ¯ φ   over each of these leaves, which does not change the integral over   .

7.3.3 Dirac Operator

The Dirac operator of   is defined on sections of S   . This extends trivially (and G   -equivariantly) to sections of S ħ   over ~   . Restricting to G   -invariant sections defines the Dirac operator D ħ   on ħ   .
If the algebras C ) ( G / Γ )   and ħ   can be identified as G   -modules, then this induces a unitary map from L 2 ( , S )   to ħ   which intertwines the Dirac operators. This is thus an isospectral deformation. We can identify the Hilbert spaces and regard the Dirac operator as constant. Because this is isospectral, the dimension axiom is trivially satisfied.

7.3.4 Real Structure

The classical real structure is given by an antilinear bundle automorphism C : S ) S   . Combining this with the involution on ħ   gives the real structure C *   on S ħ   , and hence on ħ   .
This leads to the obvious A ħ   -bimodule structure on ħ   . The various signs remain the same as in the commutative case. The first order axiom is easy to verify.

7.3.5 Smoothness

The action of D ħ   on Γ ( ~ , S ħ ) G   can be rewritten partly in terms of the Lie algebra g   . The common domain of all powers of D ħ   is simply ħ ) = Γ ( ~ , S ħ ) G   .
The finiteness axiom can be checked from this. Smoothness follows similarly.

7.3.6 Differential Forms

Let C b ) ( ~ , ħ )   denote the algebra of smooth ħ   -valued functions that are bounded in all derivatives. This is a tensor product of C b ) ( ~ )   with ħ   . With D 1   and the Hilbert B ħ   -module L 2 ( ~ , S B ħ )   this forms a spectral triple (in a slightly generalized sense). The algebra of noncommutative differential forms is simply Ω D 1 * [ C b ) ( ~ , ħ ) ] = Ω b * ( ~ , ħ ) .   Let ρ 1 : Ω * [ C b ) ( ~ ) ] ) B ħ [ L 2 ( ~ , S B ħ ) ]   be the representation used in the construction.
The image consists of bounded-adjointable operators which act locally over ~   .
The same is true for the representation ρ 2 : Ω * ( A ħ ) ) ( ħ )   . Consequently, ker ρ 2 = ker ρ 1 Ω * ( A ħ )   . This means that Ω D ħ ( A ħ )   is the differential-graded subalgebra of Ω b * ( ~ , ħ )   generated by A ħ   in degree 0   . Thus Ω D ħ ( A ħ ) = Ω * ( ~ , ħ ) G .   With this it is clear that the differential graded algebra Ω D ħ ( A ħ )   is smoothly deformed from Ω * ( )   .

7.3.7 Orientation

By construction, the volume form ε ~ Ω n ( ~ )   is G   -invariant. This means that it can be identified with ε ~ 1 Ω n ( ~ , ħ ) = Ω D ħ n ( A ħ )   . Viewing this as an equivalence class in Ω n ( A ħ )   , some element of this class should be the Hochschild cycle with image γ   as required by the orientation axiom.

7.3.8 Poincaré Duality

This axiom is essentially impossible to check at this level of generality. It depends upon the stability of K   -theory under the deformation from C ) ( )   to A ħ   . This is not a general property, but there is a strong tendency for K   -theory to be preserved in deformations, see [18.
It is certainly plausible that Poincaré duality will be preserved in an isospectral deformation which preserves K   -theory. For example, if the dimension is even, we can consider the intersection product of even K   -theory classes which are determined by idempotents e , e M a t m ( A ħ )   . Varying ħ   , the intersection product should vary continuously, but it is an integer and thus constant.

8 Examples

8.1 Torus

The only compact example in two dimensions is a flat torus with a constant symplectic structure. In this case, G   is the 2-torus and Γ   is trivial. The construction of the deformed geometry is the canonical example of noncommutative geometry.
More generally, any antisymmetric n × n   -matrix Π   defines a compatible Poisson structure on a flat n   -dimensional torus, and there is a corresponding noncommutative torus deformation. The group G   is T n   in this case. Although Π   may be degenerate, the dimension of the the group T G   may still be larger than rk Π   . A simple example is R 3 / Z 3   with π = x ( y + 2 z )   . This is the effect I referred to in defining T   to be densely generated by the span of Π   .

8.2 Flat Manifolds

Let ~ = T 4 = R 4 / Z 4   be the 4   -dimensional torus with coordinates x 1 , x 2 , y 1 , y 2   (on R 4   ) and symplectic form ω = d x 1 d y 1 + d x 2 d y 2   . The mapping s : ( x 1 , y 1 , x 2 , y 2 ) ( y 1 , - x 1 , x 2 + 1 4 , y 2 )   generates a free action of Z 4   on ~   . Define : = ~ / Z 4   .
The abelian group of Poisson isometries generated by the span of Π   is the full group of translations, T 4   . The covering group is Γ = Z 4   generated by s   . The group generated by these is the semidirect product G = Z 4 T 4   ; the generator of Z 4   acts on T 4   by a quarter rotation in the x 1   - y 1   -plane. In this case,   is of the locally homogeneous form = H \ G / Γ   with H = Γ = Z 4   .
As always, we should construct a noncommutative deformation by first deforming the homogeneous model G / Γ = T 4   . This gives a noncommutative 4-torus. The algebra ħ   is generated by 4 unitaries U 1   , U 2   , V 1   , and V 2   with the relations
U 1 V 1 = e ( 2 π ) 2 i ħ V 1 U 1 , U 2 V 2 , = e ( 2 π ) 2 i ħ V 2 U 2 ,
and all other pairs of generators commuting. Corresponding to s   is the automorphism s *   defined by: s * ( U 1 ) = V 1   , s * ( V 1 ) = U 1 - 1   , s * ( U 2 ) = i U 2   , and s * ( V 2 ) = V 2   . A ħ   is the s *   -invariant subalgebra of ħ   .
This example is reminiscent of the classification of flat manifolds: Any flat compact Riemannian manifold is of the form = T n / Γ   with Γ   a finite group of free isometries of the flat torus (see [4). If there exists a constant Γ   -invariant Poisson structure on T n   , then this descends to a Poisson structure on   . Choosing ~ = T n   , we can take G = Γ T n   and this is another locally homogeneous example = Γ \ G / Γ   .
The standard quantization of T n   is equivariant under all linear transformations preserving the Poisson structure, therefore this gives a deformation of   .
In general, if   is a compact Riemannian manifold compatible with a symplectic structure, then it is flat and it must be of this form.

8.3 Heisenberg Manifolds

The Heisenberg group H   is the group of real 3 × 3   matrices of the form ( 1 x z 0 1 y 0 0 1 ) .   Let Λ H   be the subgroup of matrices with integer entries. The 3-dimensional nil-manifold (or “Heisenberg manifold”) is N i l 3 : = H / Λ   .
Let θ   be an irrational number. The Poisson bivector Π = ( x + θ y ) z   is bi-invariant on H   . In particular it is Λ   -invariant and defines a Poisson structure on N i l 3   . This is a regular Poisson structure. The leaves of the symplectic foliation are dense. N i l 3   is the total space of a circle bundle over the torus T 2   ; the symplectic foliation is the inverse image of a Kronecker foliation of the torus.
Obviously N i l 3 = H / Λ   describes N i l 3   in the desired form, but this is not the minimal description as in Thm. 6.7 . The universal covering H   is not the minimal covering of N i l 3   with the Poisson bivector given by global Killing vectors. Instead, we can take ~ = H / Z   where Z   is embedded in H   as the subgroup of matrices ( 1 0 z 0 1 0 0 0 1 )   with z Z   .
The covering group is Γ = Λ / Z = Z 2   . The image of Π   in the Heisenberg Lie algebra is 2-dimensional, spanned by ( 0 1 0 0 0 θ 0 0 0 ) and ( 0 0 1 0 0 0 0 0 0 ) .   This generates an abelian subgroup of Poisson isometries T = R × T 1   .
This covering manifold ~   is itself a group G = H / Z   because Z H   . Both Γ   and T   are contained in the group of right translations. In fact T Γ G   is dense. This is easiest to see by looking at the quotient G / T 1 = R 2   . Γ   maps injectively to the integer lattice Z 2 R 2   . T   maps to a 1-dimensional subgroup, a line of irrational slope. Together, these densely generate R 2   .
As a Poisson manifold, N i l 3   is homogeneous. The left action of G   on N i l 3 = G / Z 2   preserves the Poisson structure. However, it is not homogeneous as a Riemannian manifold. A compatible Riemannian metric is given by any right-invariant metric on G   . This cannot be left-invariant as well.
Rieffel constructed an equivariant deformation quantization of N i l 3   in [16. In fact, this was one of the very first examples of strict deformation quantization.
The other Heisenberg manifolds discussed there are simply finite quotients of N i l 3   . They can be seen as locally homogeneous examples subordinate to N i l 3   . Chakraborty and Sinha [3have constructed and analyzed spectral triples for these examples.

9 Conclusions

With the assumptions I have made about what constitutes noncommutative geometry, I have shown that noncommutative deformations are remarkably restricted.
The reader might wonder if the analysis here was all necessary. After all, the conclusion of Thm. 6.6 is quite simple. The reason for the complexity of the proof is that compactness leads to great simplification, but in a very indirect way. The noncompact solutions of the compatibility conditions are more complicated. For instance, consider 3   -dimensional Euclidean space with Cartesian coordinates x , y , z   , and a Poisson bracket defined by { x , y } = 1   , { x , z } = y   , and { y , z } = - x   . This satisfies the compatibility conditions, but the Poisson bivector cannot be decomposed into products of Killing vectors. This example can be noncommutatively deformed, which shows that there does not exist another local obstruction that would have simplified the analysis in Section 6.1 .
Another criticism is that my compatibility conditions are too restrictive. Although I think that my assumptions are well founded, some variation is possible. Different assumptions may lead to weaker conclusions, but I think that this has been a suitable starting point. The techniques I have used here may be useful for analyzing other scenarios.
The Da ̧ browski-Sitarz example of a spectral triple for the Podleś sphere is an example of noncommutative geometry where the orientation axiom is not satisfied.
This is partly because the homological dimension of the Podleś sphere is 0   rather than 2   . This sudden dimension drop is related to the fact that integration cannot be smoothly deformed to a trace. This example strongly suggests that it would be interesting to drop my compatibility condition 0 = d ( π ε )   between the Poisson structure and volume form. Unfortunately, this condition played a key role in the simplifications leading to the main results here. Nevertheless, the technique of using symplectic realizations to tame the metacurvature still applies. There may be other tricks that would make this problem tractable.
Some constructions of noncommutative differential forms for the fuzzy sphere and quantum groups have avoided these obstructions by using differential forms that do not correspond to classical differential forms. Instead of deforming 1   -forms, they deform 1   -forms plus some other “junk”. This approach seems unpleasantly ad hoc to me, but it is probably still possible to analyze this situation with my techniques.
The construction of spectral triples on noncommutatively deformed spaces is an active area of research. Theses examples are interesting, but in my view they cannot be considered as deformed geometries. For instance, the spectral triple for SU q ( 2 )   constructed by Chakraborty and Pal [2treats SU q ( 2 )   as a quantum group independent of SU ( 2 )   .
I hope to discuss the physical implications further in a future paper. These results rule out the idea of noncommutatively deformed 4-dimensional space-time.
At the same time, this spells out a possible structure for noncommutative extra dimensions. I have essentially classified noncommutative deformations of compact Riemannian manifolds in terms of the structure ( G , Γ , Π )   where G   is a Lie group, Γ G   is discrete and cocompact, Π 2 g   is an invariant Poisson structure, and G   is densely generated by Γ   and the Lie algebra ideal spanned by Π   . What is missing is a better understanding of this structure. For example, the Lie algebra g   is certainly not semisimple (it has a nontrivial ideal) but in all the examples that I know of, g   is actually nilpotent; it is not apparent whether g   is always nilpotent.
References

  1. Jean-Luc Brylinski: A Differential Complex for Poisson Manifolds. J. Diff. Geom. 28 (1988), 93–114.
  2. P. S. Chakraborty, A. Pal: Equivariant Spectral Triple on the Quantum SU ( 2 )   -group. K-Theory 28 (2003) 107–126. math.KT/0201004.
  3. P. S. Chakraborty, K. B. Sinha: Geometry on the Quantum Heisenberg Manifold. J. Funct. Anal. 203 (2003), 425–452. math.OA/0112270.
  4. Leonard S. Charlap: Bieberbach Groups and Flat Manifolds. Springer-Verlag, 1986.
  5. Alain Connes: Noncommutative Geometry. Academic Press, 1994.
  6. Alain Connes: Gravity Coupled with Matter and the Foundation of Non-commutative Geometry. Commun. Math. Phys. 182 (1996) 155–176. hep-th/9603053.
  7. Alain Connes, Michel Dubois-Violette: Noncommutative Finite-Dimensional Manifolds. I. Spherical manifolds and related examples. math.QA/0107070.
  8. Alain Connes, Giovanni Landi: Noncommutative Manifolds, the Instanton Algebra and Isospectral Deformations. Commun. Math. Phys. 221 (2001) 141–159. math.QA/0011194.
  9. Ludwik Da ̧ browski, Andrzej Sitarz: Dirac Operator on the Standard Podleś Quantum Sphere. Banach Center Publ. 61 (2003) 49–58. math.QA/0209048.
  10. Rui Loja Fernandes: Connections in Poisson Geometry. I. Holonomy and Invariants. J. Differential Geom. 54 (2000), 2, 303–365. math.DG/0001129.
  11. V. Ginzburg, A. Golubev: Holonomy on Poisson manifolds and the modular class. math.DG/9812153.
  12. J. Gracia-Bondía, J. C. Varilly, H. Figueroa: Elements of Noncommutative Geometry. Birkhäuser, 2001.
  13. Eli Hawkins: Noncommutative Rigidity. Commun. Math. Phys. 246 (2004) 211–235. math.QA/0211203.
  14. J. L. Koszul: Crouchet de Schouten-Nijenhuis et cohomologie. É. Cartan et les mathématiques d'aujourd'hui. Soc. Math. de France, Astérisque, hors série, 1985, 257–271.
  15. Pierre Molino: Riemannian Foliations. Birkhäuser, 1988.
  16. Marc A. Rieffel: Deformation Quantization of Heisenberg Manifolds. Commun. Math. Phys. 122, 531–562 (1989).
  17. Marc A. Rieffel: Proper Actions of Groups on C *   -Algebras. Mappings of Operator Algebras. 141–182. Progress in Mathematics 84. Birkhäuser, 1991.
  18. J. Rosenberg: Behavior of K   -theory under Quantization. Operator Algebras and Quantum Field Theory. 404–415. International Press, 1997.
  19. Izu Vaisman: On the Geometric Quantization of Poisson Manifolds. J. Math. Phys. 32 (1991) 3339–3345.
  20. Izu Vaisman: Lectures on the Geometry of Poisson Manifolds. Birkhäuser, 1994.
  21. Alan Weinstein: The Local Structure of Poisson Manifolds. J. Diff. Geometry 18 (1983) 523–557.
  22. Alan Weinstein: The modular automorphism group of a Poisson manifold. J. Geom. Phys. 23 (1997) 379–394.