WikiJournal Preprints/Poisson manifold

WikiJournal Preprints
Open access • Publication charge free • Public peer review

WikiJournal User Group is a publishing group of open-access, free-to-publish, Wikipedia-integrated academic journals. <seo title=" Wikiversity Journal User Group, WikiJournal Free to publish, Open access, Open-access, Non-profit, online journal, Public peer review "/>

<meta name='citation_doi' value=>

Article information

Abstract

In differential geometry, a field in mathematics, a Poisson manifold is a smooth manifold endowed with a Poisson structure. The notion of Poisson manifold generalises that of symplectic manifold, which in turn generalises the phase space from Hamiltonian mechanics.

A Poisson structure (or Poisson bracket) on a smooth manifold   is a function

 
on the vector space   of smooth functions on  , making it into a Lie algebra subject to a Leibniz rule (also known as a Poisson algebra).

Poisson structures on manifolds were introduced by André Lichnerowicz in 1977[1] and are named after the French mathematician Siméon Denis Poisson, due to their early appearance in his works on analytical mechanics.[2]


Introduction edit

From phase spaces of classical mechanics to symplectic and Poisson manifolds edit

In classical mechanics, the phase space of a physical system consists of all the possible values of the position and of the momentum variables allowed by the system. It is naturally endowed with a Poisson bracket/symplectic form (see below), which allows one to formulate the Hamilton equations and describe the dynamics of the system through the phase space in time.

For instance, a single particle freely moving in the  -dimensional Euclidean space (i.e. having   as configuration space) has phase space  . The coordinates   describe respectively the positions and the generalised momenta. The space of observables, i.e. the smooth functions on  , is naturally endowed with a binary operation called Poisson bracket, defined as  . Such bracket satisfies the standard properties of a Lie bracket, plus a further compatibility with the product of functions, namely the Leibniz identity  . Equivalently, the Poisson bracket on   can be reformulated using the symplectic form  . Indeed, if one considers the Hamiltonian vector field   associated to a function  , then the Poisson bracket can be rewritten as  

In more abstract differential geometric terms, the configuration space is an  -dimensional smooth manifold  , and the phase space is its cotangent bundle   (a manifold of dimension  ). The latter is naturally equipped with a canonical symplectic form, which in canonical coordinates coincides with the one described above. In general, by Darboux theorem, any arbitrary symplectic manifold   admits special coordinates where the form   and the bracket   are equivalent with, respectively, the symplectic form and the Poisson bracket of  . Symplectic geometry is therefore the natural mathematical setting to describe classical Hamiltonian mechanics.

Poisson manifolds are further generalisations of symplectic manifolds, which arise by axiomatising the properties satisfied by the Poisson bracket on  . More precisely, a Poisson manifold consists of a smooth manifold   (not necessarily of even dimension) together with an abstract bracket  , still called Poisson bracket, which does not necessarily arise from a symplectic form  , but satisfies the same algebraic properties.

Poisson geometry is closely related to symplectic geometry: for instance, every Poisson bracket determines a foliation whose leaves are naturally equipped with symplectic forms. However, the study of Poisson geometry requires techniques that are usually not employed in symplectic geometry, such as the theory of Lie groupoids and algebroids.

Moreover, there are natural examples of structures which should be "morally" symplectic, but fails to be so. For example, the smooth quotient of a symplectic manifold by a group acting by symplectomorphisms is a Poisson manifold, which in general is not symplectic. This situation models the case of a physical system which is invariant under symmetries: the "reduced" phase space, obtained by quotienting the original phase space by the symmetries, in general is no longer symplectic, but is Poisson.

History edit

Although the modern definition of Poisson manifold appeared only in the 70's–80's, its origin dates back to the nineteenth century. Alan Weinstein synthetised the early history of Poisson geometry as follows:

"Poisson invented his brackets as a tool for classical dynamics. Jacobi realized the importance of these brackets and elucidated their algebraic properties, and Lie began the study of their geometry."[3]

Indeed, Siméon Denis Poisson introduced in 1809 what we now call Poisson bracket in order to obtain new integrals of motion, i.e. quantities which are preserved throughout the motion.[4] More precisely, he proved that, if two functions   and   are integral of motions, then there is a third function, denoted by  , which is an integral of motion as well. In the Hamiltonian formulation of mechanics, where the dynamics of a physical system is described by a given function   (usually the energy of the system), an integral of motion is simply a function   which Poisson-commutes with  , i.e. such that  . What will become known as Poisson's theorem can then be formulated as

 
Poisson computations occupied many pages, and his results were rediscovered and simplified two decades later by Carl Gustav Jacob Jacobi.[2] Jacobi was the first to identify the general properties of the Poisson bracket as a binary operation. Moreover, he established the relation between the (Poisson) bracket of two functions and the (Lie) bracket of their associated Hamiltonian vector fields, i.e.
 
in order to reformulate (and give a much shorter proof of) Poisson's theorem on integrals of motion.[5] Jacobi's work on Poisson brackets influenced the pioneering studies of Sophus Lie on symmetries of differential equations, which led to the discovery of Lie groups and Lie algebras. For instance, what are now called linear Poisson structures (i.e. Poisson brackets on a vector space which send linear functions to linear functions) correspond precisely to Lie algebra structures. Moreover, the integrability of a linear Poisson structure (see below) is closely related to the integrability of its associated Lie algebra to a Lie group.

The twentieth century saw the development of modern differential geometry, but only in 1977 André Lichnerowicz introduced Poisson structures as geometric objects on smooth manifolds.[1] Poisson manifolds were further studied in the foundational 1983 paper of Alan Weinstein, where many basic structure theorems were first proved.[6]

These works exerted a huge influence in the subsequent decades on the development of Poisson geometry, which today is a field of its own, and at the same time is deeply entangled with e.g. non-commutative geometry, integrable systems, topological field theories and representation theory.

Formal definition edit

There are two main points of view to define Poisson structures: it is customary and convenient to switch between them.

As bracket edit

Let   be a smooth manifold and let   denote the real algebra of smooth real-valued functions on  , where the multiplication is defined pointwise. A Poisson bracket (or Poisson structure) on   is an  -bilinear map

 

defining a structure of Poisson algebra on  , i.e. satisfying the following three conditions:

  • Skew symmetry:  .
  • Jacobi identity:  .
  • Leibniz's Rule:  .

The first two conditions ensure that   defines a Lie-algebra structure on  , while the third guarantees that, for each  , the linear map   is a derivation of the algebra  , i.e., it defines a vector field   called the Hamiltonian vector field associated to  .

Choosing local coordinates  , any Poisson bracket is given by

 
for   the Poisson bracket of the coordinate functions.

As bivector edit

A Poisson bivector on a smooth manifold   is a bivector field   satisfying the non-linear partial differential equation  , where

 


denotes the Schouten–Nijenhuis bracket on multivector fields. Choosing local coordinates  , any Poisson bivector is given by

 
for   skew-symmetric smooth functions on  .

Equivalence of the definitions edit

Let   be a bilinear skew-symmetric bracket (also called an almost Lie bracket) satisfying Leibniz's rule; then the function   can be described a

 
for a unique smooth bivector field  . Conversely, given any smooth bivector field   on  , the same formula   defines an almost Lie bracket   that automatically obeys Leibniz's rule.

A bivector field, or the corresponding almost Lie bracket, is called an almost Poisson structure. An almost Poisson structure is Poisson if one of the following equivalent integrability conditions holds:[5]

  •   satisfies the Jacobi identity (hence it is a Poisson bracket);
  •   satisfies   (hence it a Poisson bivector);
  • the map   is a Lie algebra homomorphism, i.e. the Hamiltonian vector fields satisfy  ;
  • the graph   defines a Dirac structure, i.e. a Lagrangian subbundle   which is closed under the standard Courant bracket.[7]

Holomorphic Poisson structures edit

The definition of Poisson structure for real smooth manifolds can be also adapted to the complex case.

A holomorphic Poisson manifold is a complex manifold   whose sheaf of holomorphic functions   is a sheaf of Poisson algebras. Equivalently, recall that a holomorphic bivector field   on a complex manifold   is a section   such that  . Then a holomorphic Poisson structure on   is a holomorphic bivector field satisfying the equation  . Holomorphic Poisson manifolds can be characterised also in terms of Poisson-Nijenhuis structures.[8]

Many results for real Poisson structures, e.g. regarding their integrability, extend also to holomorphic ones.[9][10]

Holomorphic Poisson structures appear naturally in the context of generalised complex structures: locally, any generalised complex manifold is the product of a symplectic manifold and a holomorphic Poisson manifold.[11]

Symplectic leaves edit

A Poisson manifold is naturally partitioned into regularly immersed symplectic manifolds of possibly different dimensions, called its symplectic leaves. These arise as the maximal integral submanifolds of the completely integrable singular distribution spanned by the Hamiltonian vector fields.

Rank of a Poisson structure edit

Recall that any bivector field can be regarded as a skew homomorphism  . The image   consists therefore of the values   of all Hamiltonian vector fields evaluated at every  .

The rank of   at a point   is the rank of the induced linear mapping  . A point   is called regular for a Poisson structure   on   if and only if the rank of   is constant on an open neighborhood of  ; otherwise, it is called a singular point. Regular points form an open dense subset  ; when the map   is of constant rank, the Poisson structure   is called regular. Examples of regular Poisson structures include trivial and nondegenerate structures (see below).

The regular case edit

For a regular Poisson manifold, the image   is a regular distribution; it is easy to check that it is involutive, therefore, by Frobenius theorem,   admits a partition into leaves. Moreover, the Poisson bivector restricts nicely to each leaf, which become therefore symplectic manifolds.

The non-regular case edit

For a non-regular Poisson manifold the situation is more complicated, since the distribution   is singular, i.e. the vector subspaces   have different dimensions.

An integral submanifold for   is a path-connected submanifold   satisfying   for all  . Integral submanifolds of   are automatically regularly immersed manifolds, and maximal integral submanifolds of   are called the leaves of  .

Moreover, each leaf   carries a natural symplectic form   determined by the condition   for all   and  . Correspondingly, one speaks of the symplectic leaves of  . Moreover, both the space   of regular points and its complement are saturated by symplectic leaves, so symplectic leaves may be either regular or singular.

Weinstein splitting theorem edit

To show the existence of symplectic leaves also in the non-regular case, one can use Weinstein splitting theorem (or Darboux-Weinstein theorem).[6] It states that any Poisson manifold   splits locally around a point   as the product of a symplectic manifold   and a transverse Poisson submanifold   vanishing at  . More precisely, if  , there are local coordinates   such that the Poisson bivector   splits as the sum

 
where  . Notice that, when the rank of   is maximal (e.g. the Poisson structure is nondegenerate), one recovers the classical Darboux theorem for symplectic structures.

Examples edit

Trivial Poisson structures edit

Every manifold   carries the trivial Poisson structure

 
equivalently described by the bivector  . Every point of   is therefore a zero-dimensional symplectic leaf.

Nondegenerate Poisson structures edit

A bivector field   is called nondegenerate if   is a vector bundle isomorphism. Nondegenerate Poisson bivector fields are actually the same thing as symplectic manifolds  .

Indeed, there is a bijective correspondence between nondegenerate bivector fields   and nondegenerate 2-forms  , given by

 
where   is encoded by  . Furthermore,   is Poisson precisely if and only if   is closed; in such case, the bracket becomes the canonical Poisson bracket from Hamiltonian mechanics:
 
Non-degenerate Poisson structures on connected manifolds have only one symplectic leaf, namely   itself.

Linear Poisson structures edit

A Poisson structure   on a vector space   is called linear when the bracket of two linear functions is still linear.

The class of vector spaces with linear Poisson structures coincides actually with that of (dual of) Lie algebras. Indeed, the dual   of any finite-dimensional Lie algebra   carries a linear Poisson bracket, known in the literature under the names of Lie-Poisson, Kirillov-Poisson or KKS (Kostant-Kirillov-Souriau) structure:

 
where   and the derivatives   are interpreted as elements of the bidual  . Equivalently, the Poisson bivector can be locally expressed as
 
where   are coordinates on   and   are the associated structure constants of  ,

Conversely, any linear Poisson structure   on   must be of this form, i.e. there exists a natural Lie algebra structure induced on   whose Lie-Poisson bracket recovers  .

The symplectic leaves of the Lie-Poisson structure on   are the orbits of the coadjoint action of   on  .

Fibrewise linear Poisson structures edit

The previous example can be generalised as follows. A Poisson structure on the total space of a vector bundle   is called fibrewise linear when the bracket of two smooth functions  , whose restrictions to the fibres are linear, is still linear when restricted to the fibres. Equivalently, the Poisson bivector field   is asked to satisfy   for any  , where   is the scalar multiplication  .

The class of vector bundles with linear Poisson structures coincides actually with that of (dual of) Lie algebroids. Indeed, the dual   of any Lie algebroid   carries a fibrewise linear Poisson bracket,[12] uniquely defined by

 
where   is the evaluation by  . Equivalently, the Poisson bivector can be locally expressed as
 
where   are coordinates around a point  ,   are fibre coordinates on  , dual to a local frame   of  , and   and   are the structure function of  , i.e. the unique smooth functions satisfying
 
Conversely, any fibrewise linear Poisson structure   on   must be of this form, i.e. there exists a natural Lie algebroid structure induced on   whose Lie-Poisson backet recovers  .[13]

If   is integrable to a Lie groupoid  , the symplectic leaves of   are the connected components of the orbits of the cotangent groupoid  . In general, given any algebroid orbit  , the image of its cotangent bundle via the dual   of the anchor map is a symplectic leaf.

For   one recovers linear Poisson structures, while for   the fibrewise linear Poisson structure is the nondegenerate one given by the canonical symplectic structure of the cotangent bundle  . More generally, any fibrewise linear Poisson structure on   that is nondegenerate is isomorphic to the canonical symplectic form on  .

Other examples and constructions edit

  • Any constant bivector field on a vector space is automatically a Poisson structure; indeed, all three terms in the Jacobiator are zero, being the bracket with a constant function.
  • Any bivector field on a 2-dimensional manifold is automatically a Poisson structure; indeed,   is a 3-vector field, which is always zero in dimension 2.
  • Given any Poisson bivector field   on a 3-dimensional manifold  , the bivector field  , for any  , is automatically Poisson.
  • The Cartesian product   of two Poisson manifolds   and   is again a Poisson manifold.
  • Let   be a (regular) foliation of dimension   on   and   a closed foliated two-form for which the power   is nowhere-vanishing. This uniquely determines a regular Poisson structure on   by requiring the symplectic leaves of   to be the leaves   of   equipped with the induced symplectic form  .
  • Let   be a Lie group acting on a Poisson manifold   and such that the Poisson bracket of  -invariant functions on   is  -invariant. If the action is free and proper, the quotient manifold   inherits a Poisson structure   from   (namely, it is the only one such that the submersion   is a Poisson map).

Poisson cohomology edit

The Poisson cohomology groups   of a Poisson manifold are the cohomology groups of the cochain complex

 

where the operator   is the Schouten-Nijenhuis bracket with  . Notice that such a sequence can be defined for every bivector on  ; the condition   is equivalent to  , i.e.   being Poisson.

Using the morphism  , one obtains a morphism from the de Rham complex   to the Poisson complex  , inducing a group homomorphism  . In the nondegenerate case, this becomes an isomorphism, so that the Poisson cohomology of a symplectic manifold fully recovers its de Rham cohomology.

Poisson cohomology is difficult to compute in general, but the low degree groups contain important geometric information on the Poisson structure:

  •   is the space of the Casimir functions, i.e. smooth functions Poisson-commuting with all others (or, equivalently, smooth functions constant on the symplectic leaves);
  •   is the space of Poisson vector fields modulo Hamiltonian vector fields;
  •   is the space of the infinitesimal deformations of the Poisson structure modulo trivial deformations;
  •   is the space of the obstructions to extend infinitesimal deformations to actual deformations.

Modular class edit

The modular class of a Poisson manifold is a class in the first Poisson cohomology group: for orientable manifolds, it is the obstruction to the existence of a volume form invariant under the Hamiltonian flows.[14] It was introduced by Koszul[15] and Weinstein.[16]

Recall that the divergence of a vector field   with respect to a given volume form   is the function   defined by  . The modular vector field of an orientable Poisson manifold, with respect to a volume form  , is the vector field   defined by the divergence of the Hamiltonian vector fields:  .

The modular vector field is a Poisson 1-cocycle, i.e. it satisfies  . Moreover, given two volume forms   and  , the difference   is a Hamiltonian vector field. Accordingly, the Poisson cohomology class   does not depend on the original choice of the volume form  , and it is called the modular class of the Poisson manifold.

An orientable Poisson manifold is called unimodular if its modular class vanishes. Notice that this happens if and only if there exists a volume form   such that the modular vector field   vanishes, i.e.   for every  ; in other words,   is invariant under the flow of any Hamiltonian vector field. For instance:

  • symplectic structures are always unimodular, since the Liouville form is invariant under all Hamiltonian vector fields;
  • for linear Poisson structures the modular class is the infinitesimal modular character of  , since the modular vector field associated to the standard Lebesgue measure on   is the constant vector field on  . Then   is unimodular as Poisson manifold if and only if it is unimodular as Lie algebra;[17]
  • For regular Poisson structures the modular class is related to the Reeb class of the underlying symplectic foliation (an element of the first leafwise cohomology group, which obstructs the existence of a volume normal form invariant by vector fields tangent to the foliation).[18]

The construction of the modular class can be easily extended to non-orientable manifolds by replacing volume forms with densities.[16]

Poisson homology edit

Poisson cohomology was introduced in 1977 by Lichnerowicz himself;[1] a decade later, Brylinski introduced a homology theory for Poisson manifolds, using the operator  .[19]

Several results have been proved relating Poisson homology and cohomology.[20] For instance, for orientable unimodular Poisson manifolds, Poisson homology turns out to be isomorphic to Poisson cohomology: this was proved independently by Xu[21] and Evans-Lu-Weinstein.[17]

Poisson maps edit

A smooth map   between Poisson manifolds is called a Poisson map if it respects the Poisson structures, i.e. one of the following equivalent conditions holds (compare with the equivalent definitions of Poisson structures above):

  • the Poisson brackets   and   satisfy   for every   and smooth functions   ;
  • the bivector fields   and   are  -related, i.e.  ;
  • the Hamiltonian vector fields associated to every smooth function   are  -related, i.e.  ;
  • the differential   is a forward Dirac morphism.[7]

An anti-Poisson map satisfies analogous conditions with a minus sign on one side.

Poisson manifolds are the objects of a category  , with Poisson maps as morphisms. If a Poisson map   is also a diffeomorphism, then we call   a Poisson-diffeomorphism.

Examples edit

  • Given the product Poisson manifold  , the canonical projections  , for  , are Poisson maps.
  • The inclusion mapping of a symplectic leaf, or of an open subset, is a Poisson map.
  • Given two Lie algebras   and  , the dual of any Lie algebra homomorphism   induces a Poisson map   between their linear Poisson structures.
  • Given two Lie algebroids   and  , the dual of any Lie algebroid morphism   over the identity induces a Poisson map   between their fibrewise linear Poisson structure.

One should notice that the notion of a Poisson map is fundamentally different from that of a symplectic map. For instance, with their standard symplectic structures, there exist no Poisson maps  , whereas symplectic maps abound. More generally, given two symplectic manifolds   and   and a smooth map  , if   is a Poisson map, it must be a submersion, while if it is a symplectic map, it must be an immersion.

Symplectic realisations edit

A symplectic realisation on a Poisson manifold M consists of a symplectic manifold   together with a Poisson map   which is a surjective submersion. Roughly speaking, the role of a symplectic realisation is to "desingularise" a complicated (degenerate) Poisson manifold by passing to a bigger, but easier (non-degenerate), one.

Notice that some authors define symplectic realisations without requiring   to be a surjective submersion (so that, for instance, the inclusion of a symplectic leaf in a symplectic manifold is an example) and call full a symplectic realisation with such extra condition. Examples of (full) symplectic realisations include the following:

  • For the trivial Poisson structure  , one takes as   the cotangent bundle  , with its canonical symplectic structure, and as   the projection  .
  • For a non-degenerate Poisson structure   one takes as   the manifold   itself and as   the identity  .
  • For the Lie-Poisson structure on  , one takes as   the cotangent bundle   of a Lie group   integrating   and as   the dual map   of the differential at the identity of the (left or right) translation  .

A symplectic realisation   is called complete if, for any complete Hamiltonian vector field  , the vector field   is complete as well. While symplectic realisations always exist for every Poisson manifold (and several different proofs are available),[6][22][23] complete ones do not, and their existence plays a fundamental role in the integrability problem for Poisson manifolds (see below).[24]

Integration of Poisson manifolds edit

Any Poisson manifold   induces a structure of Lie algebroid on its cotangent bundle  , also called the cotangent algebroid. The anchor map is given by   while the Lie bracket on   is defined as

 
Several notions defined for Poisson manifolds can be interpreted via its Lie algebroid  :
  • the symplectic foliation is the usual (singular) foliation induced by the anchor of the Lie algebroid;
  • the symplectic leaves are the orbits of the Lie algebroid;
  • a Poisson structure on   is regular precisely when the associated Lie algebroid   is;
  • the Poisson cohomology groups coincide with the Lie algebroid cohomology groups of   with coefficients in the trivial representation;
  • the modular class of a Poisson manifold coincides with the modular class of the associated Lie algebroid  .[17]

It is of crucial importance to notice that the Lie algebroid   is not always integrable to a Lie groupoid.[25][26][24]

Symplectic groupoids edit

A symplectic groupoid is a Lie groupoid   together with a symplectic form   which is also multiplicative, i.e. it satisfies the following algebraic compatibility with the groupoid multiplication:  . Equivalently, the graph of   is asked to be a Lagrangian submanifold of  . Among the several consequences, the dimension of   is automatically twice the dimension of  . The notion of symplectic groupoid was introduced at the end of the 80's independently by several authors.[25][27][22][12]

A fundamental theorem states that the base space of any symplectic groupoid admits a unique Poisson structure   such that the source map   and the target map   are, respectively, a Poisson map and an anti-Poisson map. Moreover, the Lie algebroid   is isomorphic to the cotangent algebroid   associated to the Poisson manifold  .[28] Conversely, if the cotangent bundle   of a Poisson manifold is integrable (as a Lie algebroid), then its  -simply connected integration   is automatically a symplectic groupoid.[29]

Accordingly, the integrability problem for a Poisson manifold consists in finding a (symplectic) Lie groupoid which integrates its cotangent algebroid; when this happens, the Poisson structure is called integrable.

While any Poisson manifold admits a local integration (i.e. a symplectic groupoid where the multiplication is defined only locally),[28] there are general topological obstructions to its integrability, coming from the integrability theory for Lie algebroids.[30] Using such obstructions, one can show that a Poisson manifold is integrable if and only if it admits a complete symplectic realisation.[24] This fact can also be proved more directly, without using Crainic-Fernandes obstructions.[31]

The candidate   for the symplectic groupoid integrating a given Poisson manifold   is called Poisson homotopy groupoid and is simply the Weinstein groupoid of the cotangent algebroid  , consisting of the quotient of the Banach space of a special class of paths in   by a suitable equivalent relation. Equivalently,   can be described as an infinite-dimensional symplectic quotient.[26]

Examples of integrations edit

  • The trivial Poisson structure   is always integrable, a symplectic groupoid being the bundle of abelian (additive) groups   with the canonical symplectic form.
  • A non-degenerate Poisson structure on   is always integrable, a symplectic groupoid being the pair groupoid   together with the symplectic form   (for  ).
  • A Lie-Poisson structure on   is always integrable, a symplectic groupoid being the (coadjoint) action groupoid  , for   a Lie group integrating  , together with the canonical symplectic form of  .
  • A Lie-Poisson structure on   is integrable if and only if the Lie algebroid   is integrable to a Lie groupoid  , a symplectic groupoid being the cotangent groupoid   with the canonical symplectic form.

Submanifolds edit

A Poisson submanifold of   is an immersed submanifold   together with a Poisson structure   such that one of the following equivalent conditions holds:[6]

  • the immersion map   is a Poisson map;
  • the image of   is inside  ;
  • every Hamiltonian vector field  , for  , is tangent to  .

This definition is very natural and satisfies several good properties, e.g. the transverse intersection of two Poisson submanifolds is again a Poisson submanifold. However, it has also a few problems:

  • Poisson submanifolds are rare: for instance, the only Poisson submanifolds of a symplectic manifold are the open sets;
  • the definition does not behave functorially: if   is a Poisson map transverse to a Poisson submanifold   of  , the submanifold   of   is not necessarily Poisson.

In order to overcome these problems, one often uses the notion of a Poisson transversal (originally called cosymplectic submanifold).[6] This can be defined as a submanifold   which is transverse to every symplectic leaf   and such that the intersection   is a symplectic submanifold of  . It follows that any Poisson transversal   inherits a canonical Poisson structure   from  . In the case of a nondegenerate Poisson manifold   (whose only symplectic leaf is   itself), Poisson transversals are the same thing as symplectic submanifolds.

More general classes of submanifolds play an important role in Poisson geometry, including Lie–Dirac submanifolds, Poisson–Dirac submanifolds, coisotropic submanifolds and pre-Poisson submanifolds.[32]

References edit

  1. 1.0 1.1 1.2 Lichnerowicz, A. (1977). "Les variétés de Poisson et leurs algèbres de Lie associées". Journal of Differential Geometry 12 (2): 253–300. doi:10.4310/jdg/1214433987. 
  2. 2.0 2.1 Kosmann-Schwarzbach, Yvette (2022-11-29). "Seven Concepts Attributed to Siméon-Denis Poisson". SIGMA. Symmetry, Integrability and Geometry: Methods and Applications 18: 092. doi:10.3842/SIGMA.2022.092. https://www.emis.de/journals/SIGMA/2022/092/. 
  3. Weinstein, Alan (1998-08-01). "Poisson geometry". Differential Geometry and Its Applications. Symplectic Geometry 9 (1): 213–238. doi:10.1016/S0926-2245(98)00022-9. ISSN 0926-2245. 
  4. Poisson, Siméon Denis (1809). "Sur la variation des constantes arbitraires dans les questions de mécanique". Journal de l'École polytechnique [fr] 15e cahier (8): 266–344. https://babel.hathitrust.org/cgi/pt?id=mdp.39015074785596&view=1up&seq=280. 
  5. 5.0 5.1 Silva, Ana Cannas da; Weinstein, Alan (1999). Geometric models for noncommutative algebras. Providence, R.I.: American Mathematical Society. ISBN 0-8218-0952-0. OCLC 42433917. https://math.berkeley.edu/~alanw/Models.pdf. 
  6. 6.0 6.1 6.2 6.3 6.4 Weinstein, Alan (1983-01-01). "The local structure of Poisson manifolds". Journal of Differential Geometry 18 (3). doi:10.4310/jdg/1214437787. ISSN 0022-040X. 
  7. 7.0 7.1 Bursztyn, Henrique; Radko, Olga (2003). "Gauge equivalence of Dirac structures and symplectic groupoids". Annales de l’institut Fourier 53 (1): 309–337. doi:10.5802/aif.1945. ISSN 0373-0956. https://aif.centre-mersenne.org/item/AIF_2003__53_1_309_0/. 
  8. Laurent-Gengoux, C.; Stienon, M.; Xu, P. (2010-07-08). "Holomorphic Poisson Manifolds and Holomorphic Lie Algebroids". International Mathematics Research Notices 2008. doi:10.1093/imrn/rnn088. ISSN 1073-7928. https://academic.oup.com/imrn/article-lookup/doi/10.1093/imrn/rnn088. 
  9. Laurent-Gengoux, Camille; Stiénon, Mathieu; Xu, Ping (2009-12-01). "Integration of holomorphic Lie algebroids". Mathematische Annalen 345 (4): 895–923. doi:10.1007/s00208-009-0388-7. ISSN 1432-1807. https://doi.org/10.1007/s00208-009-0388-7. 
  10. Broka, Damien; Xu, Ping (2022). "Symplectic realizations of holomorphic Poisson manifolds". Mathematical Research Letters 29 (4): 903–944. doi:10.4310/MRL.2022.v29.n4.a1. ISSN 1945-001X. https://www.intlpress.com/site/pub/pages/journals/items/mrl/content/vols/0029/0004/a001/index.php. 
  11. Bailey, Michael (2013-08-01). "Local classification of generalize complex structures". Journal of Differential Geometry 95 (1). doi:10.4310/jdg/1375124607. ISSN 0022-040X. 
  12. 12.0 12.1 Coste, A.; Dazord, P.; Weinstein, A. (1987). "Groupoïdes symplectiques". Publications du Département de mathématiques (Lyon) (2A): 1–62. ISSN 2547-6300. http://www.numdam.org/item/PDML_1987___2A_1_0/. 
  13. Courant, Theodore James (1990). "Dirac manifolds". Transactions of the American Mathematical Society 319 (2): 631–661. doi:10.1090/S0002-9947-1990-0998124-1. ISSN 0002-9947. https://www.ams.org/tran/1990-319-02/S0002-9947-1990-0998124-1/. 
  14. Kosmann-Schwarzbach, Yvette (2008-01-16). "Poisson Manifolds, Lie Algebroids, Modular Classes: a Survey". SIGMA. Symmetry, Integrability and Geometry: Methods and Applications 4: 005. doi:10.3842/SIGMA.2008.005. http://www.emis.de/journals/SIGMA/2008/005/. 
  15. Koszul, Jean-Louis (1985). "Crochet de Schouten-Nijenhuis et cohomologie". Astérisque S131: 257–271. http://www.numdam.org/item/?id=AST_1985__S131__257_0. 
  16. 16.0 16.1 Weinstein, Alan (1997-11-01). "The modular automorphism group of a Poisson manifold". Journal of Geometry and Physics 23 (3): 379–394. doi:10.1016/S0393-0440(97)80011-3. ISSN 0393-0440. https://www.sciencedirect.com/science/article/pii/S0393044097800113. 
  17. 17.0 17.1 17.2 Evens, Sam; Lu, Jiang-Hua; Weinstein, Alan (1999). "Transverse measures, the modular class and a cohomology pairing for Lie algebroids". The Quarterly Journal of Mathematics 50 (200): 417–436. doi:10.1093/qjmath/50.200.417. https://academic.oup.com/qjmath/article-abstract/50/200/417/1515478?redirectedFrom=fulltext&login=false. 
  18. Abouqateb, Abdelhak; Boucetta, Mohamed (2003-07-01). "The modular class of a regular Poisson manifold and the Reeb class of its symplectic foliation". Comptes Rendus Mathematique 337 (1): 61–66. doi:10.1016/S1631-073X(03)00254-1. ISSN 1631-073X. 
  19. Brylinski, Jean-Luc (1988-01-01). "A differential complex for Poisson manifolds". Journal of Differential Geometry 28 (1). doi:10.4310/jdg/1214442161. ISSN 0022-040X. https://projecteuclid.org/journals/journal-of-differential-geometry/volume-28/issue-1/A-differential-complex-for-Poisson-manifolds/10.4310/jdg/1214442161.full. 
  20. Fernández, Marisa; Ibáñez, Raúl; de León, Manuel (1996). "Poisson cohomology and canonical homology of Poisson manifolds". Archivum Mathematicum 032 (1): 29–56. ISSN 0044-8753. https://eudml.org/doc/247851. 
  21. Xu, Ping (1999-02-01). "Gerstenhaber Algebras and BV-Algebras in Poisson Geometry". Communications in Mathematical Physics 200 (3): 545–560. doi:10.1007/s002200050540. ISSN 1432-0916. https://doi.org/10.1007/s002200050540. 
  22. 22.0 22.1 Karasev, M. V. (1987-06-30). "Analogues of the Objects of Lie Group Theory for Nonlinear Poisson Brackets". Mathematics of the USSR-Izvestiya 28 (3): 497–527. doi:10.1070/im1987v028n03abeh000895. ISSN 0025-5726. http://dx.doi.org/10.1070/IM1987v028n03ABEH000895. 
  23. Crainic, Marius; Marcut, Ioan (2011). "On the extistence of symplectic realizations". Journal of Symplectic Geometry 9 (4): 435–444. doi:10.4310/JSG.2011.v9.n4.a2. ISSN 1540-2347. https://www.intlpress.com/site/pub/pages/journals/items/jsg/content/vols/0009/0004/a002/abstract.php. 
  24. 24.0 24.1 24.2 Crainic, Marius; Fernandes, Rui (2004-01-01). "Integrability of Poisson Brackets". Journal of Differential Geometry 66 (1). doi:10.4310/jdg/1090415030. ISSN 0022-040X. 
  25. 25.0 25.1 Weinstein, Alan (1987-01-01). "Symplectic groupoids and Poisson manifolds". Bulletin of the American Mathematical Society 16 (1): 101–105. doi:10.1090/S0273-0979-1987-15473-5. ISSN 0273-0979. https://www.ams.org/journal-getitem?pii=S0273-0979-1987-15473-5. 
  26. 26.0 26.1 Cattaneo, Alberto S.; Felder, Giovanni (2001). "Poisson sigma models and symplectic groupoids". Quantization of Singular Symplectic Quotients (Basel: Birkhäuser): 61–93. doi:10.1007/978-3-0348-8364-1_4. ISBN 978-3-0348-8364-1. https://link.springer.com/chapter/10.1007/978-3-0348-8364-1_4. 
  27. Zakrzewski, S. (1990). "Quantum and classical pseudogroups. II. Differential and symplectic pseudogroups". Communications in Mathematical Physics 134 (2): 371–395. doi:10.1007/BF02097707. ISSN 0010-3616. https://projecteuclid.org/journals/communications-in-mathematical-physics/volume-134/issue-2/Quantum-and-classical-pseudogroups-II-Differential-and-symplectic-pseudogroups/cmp/1104201735.full. 
  28. 28.0 28.1 Albert, Claude; Dazord, Pierre (1991). Dazord, Pierre; Weinstein, Alan. eds. "Groupoïdes de Lie et Groupoïdes Symplectiques". Symplectic Geometry, Groupoids, and Integrable Systems. Mathematical Sciences Research Institute Publications (New York, NY: Springer US) 20: 1–11. doi:10.1007/978-1-4613-9719-9_1. ISBN 978-1-4613-9719-9. https://link.springer.com/chapter/10.1007%2F978-1-4613-9719-9_1. 
  29. Mackenzie, Kirill C. H.; Xu, Ping (2000-05-01). "Integration of Lie bialgebroids". Topology 39 (3): 445–467. doi:10.1016/S0040-9383(98)00069-X. ISSN 0040-9383. https://www.sciencedirect.com/science/article/pii/S004093839800069X. 
  30. Crainic, Marius; Fernandes, Rui (2003-03-01). "Integrability of Lie brackets". Annals of Mathematics 157 (2): 575–620. doi:10.4007/annals.2003.157.575. ISSN 0003-486X. 
  31. Álvarez, Daniel (2021-11). "Complete Lie algebroid actions and the integrability of Lie algebroids". Proceedings of the American Mathematical Society 149 (11): 4923–4930. doi:10.1090/proc/15586. ISSN 0002-9939. https://www.ams.org/proc/2021-149-11/S0002-9939-2021-15586-X/. 
  32. Zambon, Marco (2011). Ebeling, Wolfgang; Hulek, Klaus; Smoczyk, Knut. eds. "Submanifolds in Poisson geometry: a survey". Complex and Differential Geometry. Springer Proceedings in Mathematics (Berlin, Heidelberg: Springer) 8: 403–420. doi:10.1007/978-3-642-20300-8_20. ISBN 978-3-642-20300-8. https://link.springer.com/chapter/10.1007%2F978-3-642-20300-8_20. 

Further reading edit