Poisson manifold

Last updated

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.

Contents

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]

A Poisson structure on a manifold gives a way of deforming the product of functions on to a new product that is typically not commutative. This process is known as deformation quantization, since classical mechanics can be based on Poisson structures, while quantum mechanics involves non-commutative rings.

Introduction

From phase spaces of classical mechanics to symplectic and Poisson manifolds

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 of the manifold into symplectic submanifolds. 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 exhibit singularities, i.e. their "symplectic form" should be allowed to be degenerate. 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 quotienting the original phase space by the symmetries, in general is no longer symplectic, but is Poisson.

History

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 summarized 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

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

As bracket

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:

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

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

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.

Then the following integrability conditions are equivalent:

A Poisson structure without any of the four requirements above is also called an almost Poisson structure. [5]

Holomorphic Poisson structures

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. [7]

Many results for real Poisson structures, e.g. regarding their integrability, extend also to holomorphic ones. [8] [9]

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. [10]

Deformation quantization

The notion of a Poisson manifold arises naturally from the deformation theory of associative algebras. For a smooth manifold , the smooth functions form a commutative algebra over the real numbers , using pointwise addition and multiplication (meaning that for points in ). An th-order deformation of this algebra is given by a formula

for such that the star-product is associative (modulo ), but not necessarily commutative.

A first-order deformation of is equivalent to an almost Poisson structure as defined above, that is, a bilinear "bracket" map

that is skew-symmetric and satisfies Leibniz's Rule. [5] Explicitly, one can go from the deformation to the bracket by

A first-order deformation is also equivalent to a bivector field, that is, a smooth section of .

A bracket satisfies the Jacobi identity (that is, it is a Poisson structure) if and only if the corresponding first-order deformation of can be extended to a second-order deformation. [5] Remarkably, the Kontsevich quantization formula shows that every Poisson manifold has a deformation quantization. That is, if a first-order deformation of can be extended to second order, then it can be extended to infinite order.

Example: For any smooth manifold , the cotangent bundle is a symplectic manifold, and hence a Poisson manifold. The corresponding non-commutative deformation of is related to the algebra of linear differential operators on . When is the real line , the non-commutativity of the algebra of differential operators (known as the Weyl algebra) follows from the calculation that

Symplectic leaves

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 foliation spanned by the Hamiltonian vector fields.

Rank of a Poisson structure

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 subspace ; when , i.e. 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

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

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

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

Trivial Poisson structures

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

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 have only one symplectic leaf, namely itself, and their Poisson algebra become a Poisson ring.

Linear Poisson structures

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

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, [11] 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 . [12]

The symplectic leaves of are the cotangent bundles of the algebroid orbits ; equivalently, if is integrable to a Lie groupoid , they are the connected components of the orbits of the cotangent groupoid .

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 .

Other examples and constructions

Poisson cohomology

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:

Modular class

The modular class of a Poisson manifold is a class in the first Poisson cohomology group, which is the obstruction to the existence of a volume form invariant under the Hamiltonian flows. [13] It was introduced by Koszul [14] and Weinstein. [15]

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 a 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.

A 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:

Poisson homology

Poisson cohomology was introduced in 1977 by Lichnerowicz himself; [1] a decade later, Brylinski introduced a homology theory for Poisson manifolds, using the operator . [18]

Several results have been proved relating Poisson homology and cohomology. [19] For instance, for orientable unimodular Poisson manifolds, Poisson homology turns out to be isomorphic to Poisson cohomology: this was proved independently by Xu [20] and Evans-Lu-Weinstein. [16]

Poisson maps

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):

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

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.

Symplectic realisations

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 this last condition (so that, for instance, the inclusion of a symplectic leaf in a symplectic manifold is an example) and call full a symplectic realisation where is a surjective submersion. Examples of (full) symplectic realisations include the following:

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] [21] [22] complete ones do not, and their existence plays a fundamental role in the integrability problem for Poisson manifolds (see below). [23]

Integration of Poisson manifolds

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 :

It is of crucial importance to notice that the Lie algebroid is not always integrable to a Lie groupoid.

Symplectic groupoids

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. [24] [25] [21] [11]

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 . [26] Conversely, if the cotangent bundle of a Poisson manifold is integrable to some Lie groupoid , then is automatically a symplectic groupoid. [27]

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), [26] there are general topological obstructions to its integrability, coming from the integrability theory for Lie algebroids. [28] Using such obstructions, one can show that a Poisson manifold is integrable if and only if it admits a complete symplectic realisation. [23]

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. [29]

Examples of integrations

Submanifolds

A Poisson submanifold of is an immersed submanifold such that the immersion map is a Poisson map. Equivalently, one asks that 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:

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. [30]

See also

Related Research Articles

In differential geometry, a subject of mathematics, a symplectic manifold is a smooth manifold, , equipped with a closed nondegenerate differential 2-form , called the symplectic form. The study of symplectic manifolds is called symplectic geometry or symplectic topology. Symplectic manifolds arise naturally in abstract formulations of classical mechanics and analytical mechanics as the cotangent bundles of manifolds. For example, in the Hamiltonian formulation of classical mechanics, which provides one of the major motivations for the field, the set of all possible configurations of a system is modeled as a manifold, and this manifold's cotangent bundle describes the phase space of the system.

<span class="mw-page-title-main">Hamiltonian mechanics</span> Formulation of classical mechanics using momenta

In physics, Hamiltonian mechanics is a reformulation of Lagrangian mechanics that emerged in 1833. Introduced by Sir William Rowan Hamilton, Hamiltonian mechanics replaces (generalized) velocities used in Lagrangian mechanics with (generalized) momenta. Both theories provide interpretations of classical mechanics and describe the same physical phenomena.

<span class="mw-page-title-main">Poisson bracket</span> Operation in Hamiltonian mechanics

In mathematics and classical mechanics, the Poisson bracket is an important binary operation in Hamiltonian mechanics, playing a central role in Hamilton's equations of motion, which govern the time evolution of a Hamiltonian dynamical system. The Poisson bracket also distinguishes a certain class of coordinate transformations, called canonical transformations, which map canonical coordinate systems into canonical coordinate systems. A "canonical coordinate system" consists of canonical position and momentum variables that satisfy canonical Poisson bracket relations. The set of possible canonical transformations is always very rich. For instance, it is often possible to choose the Hamiltonian itself as one of the new canonical momentum coordinates.

In mathematics, a Lie algebroid is a vector bundle together with a Lie bracket on its space of sections and a vector bundle morphism , satisfying a Leibniz rule. A Lie algebroid can thus be thought of as a "many-object generalisation" of a Lie algebra.

In mathematics, a Lie groupoid is a groupoid where the set of objects and the set of morphisms are both manifolds, all the category operations are smooth, and the source and target operations

In mathematics, and especially differential geometry and gauge theory, a connection is a device that defines a notion of parallel transport on the bundle; that is, a way to "connect" or identify fibers over nearby points. A principal G-connection on a principal G-bundle over a smooth manifold is a particular type of connection which is compatible with the action of the group .

In physics, a first class constraint is a dynamical quantity in a constrained Hamiltonian system whose Poisson bracket with all the other constraints vanishes on the constraint surface in phase space. To calculate the first class constraint, one assumes that there are no second class constraints, or that they have been calculated previously, and their Dirac brackets generated.

In physics and mathematics, supermanifolds are generalizations of the manifold concept based on ideas coming from supersymmetry. Several definitions are in use, some of which are described below.

In mathematics and physics, a Hamiltonian vector field on a symplectic manifold is a vector field defined for any energy function or Hamiltonian. Named after the physicist and mathematician Sir William Rowan Hamilton, a Hamiltonian vector field is a geometric manifestation of Hamilton's equations in classical mechanics. The integral curves of a Hamiltonian vector field represent solutions to the equations of motion in the Hamiltonian form. The diffeomorphisms of a symplectic manifold arising from the flow of a Hamiltonian vector field are known as canonical transformations in physics and (Hamiltonian) symplectomorphisms in mathematics.

In mathematics, the Moyal product is an example of a phase-space star product. It is an associative, non-commutative product, , on the functions on , equipped with its Poisson bracket. It is a special case of the -product of the "algebra of symbols" of a universal enveloping algebra.

In differential geometry, a discipline within mathematics, a distribution on a manifold is an assignment of vector subspaces satisfying certain properties. In the most common situations, a distribution is asked to be a vector subbundle of the tangent bundle .

In mathematics, specifically in symplectic geometry, the momentum map is a tool associated with a Hamiltonian action of a Lie group on a symplectic manifold, used to construct conserved quantities for the action. The momentum map generalizes the classical notions of linear and angular momentum. It is an essential ingredient in various constructions of symplectic manifolds, including symplectic (Marsden–Weinstein) quotients, discussed below, and symplectic cuts and sums.

In a field of mathematics known as differential geometry, a Courant geometry was originally introduced by Zhang-Ju Liu, Alan Weinstein and Ping Xu in their investigation of doubles of Lie bialgebroids in 1997. Liu, Weinstein and Xu named it after Courant, who had implicitly devised earlier in 1990 the standard prototype of Courant algebroid through his discovery of a skew symmetric bracket on , called Courant bracket today, which fails to satisfy the Jacobi identity. Both this standard example and the double of a Lie bialgebra are special instances of Courant algebroids.

In mathematics, a diffiety is a geometrical object which plays the same role in the modern theory of partial differential equations that algebraic varieties play for algebraic equations, that is, to encode the space of solutions in a more conceptual way. The term was coined in 1984 by Alexandre Mikhailovich Vinogradov as portmanteau from differential variety.

In mathematics, the Atiyah algebroid, or Atiyah sequence, of a principal -bundle over a manifold , where is a Lie group, is the Lie algebroid of the gauge groupoid of . Explicitly, it is given by the following short exact sequence of vector bundles over :

A differentiable stack is the analogue in differential geometry of an algebraic stack in algebraic geometry. It can be described either as a stack over differentiable manifolds which admits an atlas, or as a Lie groupoid up to Morita equivalence.

A Lie bialgebroid is a mathematical structure in the area of non-Riemannian differential geometry. In brief a Lie bialgebroid are two compatible Lie algebroids defined on dual vector bundles. They form the vector bundle version of a Lie bialgebra.

In mathematics, and especially differential geometry and mathematical physics, gauge theory is the general study of connections on vector bundles, principal bundles, and fibre bundles. Gauge theory in mathematics should not be confused with the closely related concept of a gauge theory in physics, which is a field theory which admits gauge symmetry. In mathematics theory means a mathematical theory, encapsulating the general study of a collection of concepts or phenomena, whereas in the physical sense a gauge theory is a mathematical model of some natural phenomenon.

In mathematics, and especially symplectic geometry, the Thomas–Yau conjecture asks for the existence of a stability condition, similar to those which appear in algebraic geometry, which guarantees the existence of a solution to the special Lagrangian equation inside a Hamiltonian isotopy class of Lagrangian submanifolds. In particular the conjecture contains two difficulties: first it asks what a suitable stability condition might be, and secondly if one can prove stability of an isotopy class if and only if it contains a special Lagrangian representative.

In mathematical physics, the Garnier integrable system, also known as the classical Gaudin model is a classical mechanical system discovered by René Garnier in 1919 by taking the 'Painlevé simplification' or 'autonomous limit' of the Schlesinger equations. It is a classical analogue to the quantum Gaudin model due to Michel Gaudin. The classical Gaudin models are integrable.

References

  1. 1 2 3 Lichnerowicz, A. (1977). "Les variétés de Poisson et leurs algèbres de Lie associées". J. Diff. Geom. 12 (2): 253–300. doi: 10.4310/jdg/1214433987 . MR   0501133.
  2. 1 2 Kosmann-Schwarzbach, Yvette (2022-11-29). "Seven Concepts Attributed to Siméon-Denis Poisson". SIGMA. Symmetry, Integrability and Geometry: Methods and Applications. 18: 092. arXiv: 2211.15946 . doi: 10.3842/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" [On the variation of arbitrary constants in the questions of mechanics]. Journal de l'École polytechnique  [ fr ] (in French). 15e cahier (8): 266–344 via HathiTrust.
  5. 1 2 3 4 Silva, Ana Cannas da; Weinstein, Alan (1999). Geometric models for noncommutative algebras (PDF). Providence, R.I.: American Mathematical Society. ISBN   0-8218-0952-0. OCLC   42433917.
  6. 1 2 3 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. Laurent-Gengoux, C.; Stienon, M.; Xu, P. (2010-07-08). "Holomorphic Poisson Manifolds and Holomorphic Lie Algebroids". International Mathematics Research Notices . 2008. arXiv: 0707.4253 . doi:10.1093/imrn/rnn088. ISSN   1073-7928.
  8. Laurent-Gengoux, Camille; Stiénon, Mathieu; Xu, Ping (2009-12-01). "Integration of holomorphic Lie algebroids". Mathematische Annalen . 345 (4): 895–923. arXiv: 0803.2031 . doi:10.1007/s00208-009-0388-7. ISSN   1432-1807. S2CID   41629.
  9. Broka, Damien; Xu, Ping (2022). "Symplectic realizations of holomorphic Poisson manifolds". Mathematical Research Letters. 29 (4): 903–944. arXiv: 1512.08847 . doi: 10.4310/MRL.2022.v29.n4.a1 . ISSN   1945-001X.
  10. Bailey, Michael (2013-08-01). "Local classification of generalize complex structures". Journal of Differential Geometry . 95 (1). arXiv: 1201.4887 . doi: 10.4310/jdg/1375124607 . ISSN   0022-040X.
  11. 1 2 Coste, A.; Dazord, P.; Weinstein, A. (1987). "Groupoïdes symplectiques" [Symplectic groupoids]. Publications du Département de mathématiques (Lyon) (in French) (2A): 1–62. ISSN   2547-6300.
  12. 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.
  13. Kosmann-Schwarzbach, Yvette (2008-01-16). "Poisson Manifolds, Lie Algebroids, Modular Classes: a Survey". SIGMA. Symmetry, Integrability and Geometry: Methods and Applications. 4: 005. arXiv: 0710.3098 . Bibcode:2008SIGMA...4..005K. doi: 10.3842/SIGMA.2008.005 .
  14. Koszul, Jean-Louis (1985). "Crochet de Schouten-Nijenhuis et cohomologie" [Schouten-Nijenhuis bracket and cohomology]. Astérisque (in French). S131: 257–271.
  15. Weinstein, Alan (1997-11-01). "The modular automorphism group of a Poisson manifold". Journal of Geometry and Physics . 23 (3): 379–394. Bibcode:1997JGP....23..379W. doi:10.1016/S0393-0440(97)80011-3. ISSN   0393-0440.
  16. 1 2 3 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. arXiv: dg-ga/9610008 . doi:10.1093/qjmath/50.200.417.
  17. 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. arXiv: math/0211405v1 . doi: 10.1016/S1631-073X(03)00254-1 . ISSN   1631-073X.
  18. 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. S2CID   122451743.
  19. Fernández, Marisa; Ibáñez, Raúl; León, Manuel de (1996). "Poisson cohomology and canonical homology of Poisson manifolds". Archivum Mathematicum. 032 (1): 29–56. ISSN   0044-8753.
  20. Xu, Ping (1999-02-01). "Gerstenhaber Algebras and BV-Algebras in Poisson Geometry". Communications in Mathematical Physics . 200 (3): 545–560. arXiv: dg-ga/9703001 . Bibcode:1999CMaPh.200..545X. doi:10.1007/s002200050540. ISSN   1432-0916. S2CID   16559555.
  21. 1 2 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. Bibcode:1987IzMat..28..497K. doi:10.1070/im1987v028n03abeh000895. ISSN   0025-5726.
  22. 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.
  23. 1 2 Crainic, Marius; Fernandes, Rui (2004-01-01). "Integrability of Poisson Brackets". Journal of Differential Geometry . 66 (1). arXiv: math/0210152 . doi: 10.4310/jdg/1090415030 . ISSN   0022-040X.
  24. 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.
  25. 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. S2CID   122926678 via Project Euclid.
  26. 1 2 Albert, Claude; Dazord, Pierre (1991). Dazord, Pierre; Weinstein, Alan (eds.). "Groupoïdes de Lie et Groupoïdes Symplectiques" [Lie Groupoids and Symplectic Groupoids]. Symplectic Geometry, Groupoids, and Integrable Systems. Mathematical Sciences Research Institute Publications (in French). 20. New York, NY: Springer US: 1–11. doi:10.1007/978-1-4613-9719-9_1. ISBN   978-1-4613-9719-9.
  27. Liu, Z. -J.; Xu, P. (1996-01-01). "Exact Lie bialgebroids and Poisson groupoids". Geometric & Functional Analysis. 6 (1): 138–145. doi:10.1007/BF02246770. ISSN   1420-8970. S2CID   121836719 via European Digital Mathematics Library.
  28. Crainic, Marius; Fernandes, Rui (2003-03-01). "Integrability of Lie brackets". Annals of Mathematics . 157 (2): 575–620. arXiv: math/0105033 . doi: 10.4007/annals.2003.157.575 . ISSN   0003-486X.
  29. Cattaneo, Alberto S.; Felder, Giovanni (2001). "Poisson sigma models and symplectic groupoids". Quantization of Singular Symplectic Quotients. Progress in Mathematics. Basel: Birkhäuser: 61–93. arXiv: math/0003023 . doi:10.1007/978-3-0348-8364-1_4. ISBN   978-3-0348-8364-1. S2CID   10248666.
  30. Zambon, Marco (2011). Ebeling, Wolfgang; Hulek, Klaus; Smoczyk, Knut (eds.). "Submanifolds in Poisson geometry: a survey". Complex and Differential Geometry. Springer Proceedings in Mathematics. 8. Berlin, Heidelberg: Springer: 403–420. doi:10.1007/978-3-642-20300-8_20. ISBN   978-3-642-20300-8.

Books and surveys