Fact-checked by Grok 2 weeks ago

Symplectic manifold

A symplectic manifold is a pair (M, \omega) consisting of a smooth manifold M of even $2n and a closed nondegenerate 2-form \omega, called the symplectic form, which satisfies d\omega = 0 and induces an isomorphism between the and at every point. The nondegeneracy condition ensures that \omega^n / n! defines a natural on M, providing a and enabling the study of dynamics via vector fields X_H satisfying \iota_{X_H} \omega = -dH for smooth functions H: M \to \mathbb{R}. Symplectic manifolds form the foundational structure of , a field that generalizes to infinite-dimensional settings and connects with and physics. By the Darboux theorem, every symplectic manifold is locally symplectomorphic to the standard symplectic \mathbb{R}^{2n} with form \sum_{i=1}^n dx_i \wedge dy_i, highlighting their uniform local structure despite global complexity. Key properties include the preservation of \omega under symplectomorphisms, which are diffeomorphisms pulling back \omega to itself, and the existence of moment maps for Hamiltonian group actions, linking symmetries to conserved quantities via . Prominent examples include the cotangent bundle T^*Q of any manifold Q, equipped with the canonical symplectic form \sum dq_i \wedge dp_i, which models phase spaces in Lagrangian and Hamiltonian mechanics. Other instances are the 2-sphere S^2 with the form \omega_p(u,v) = \langle p, u \times v \rangle, Kähler manifolds like complex projective space \mathbb{CP}^n with the Fubini-Study metric, and toric varieties arising from torus actions. These structures underpin applications in classical and quantum physics, such as rigid body motion and gauge theories, and in mathematics, including symplectic reduction via the Marsden-Weinstein theorem, which quotients out symmetries to yield new symplectic manifolds. Historically, emerged in the through the study of systems in , with the term "" coined by in the early 20th century for the associated linear group. Modern developments accelerated in the 1960s with Vladimir Arnold's conjectures on fixed points of symplectomorphisms, resolved using , and Alan Weinstein's foundational work on submanifolds and reduction. Further advances, including Mikhail Gromov's introduction of pseudoholomorphic curves in 1985 and convexity theorems by Duistermaat-Heckman and Atiyah-Guillemin-Sternberg, have deepened connections to and .

Introduction and Motivation

Historical Development

The foundations of symplectic geometry trace back to classical mechanics in the early . In 1809, introduced the as a tool for analyzing variations in mechanical systems, providing a on phase space coordinates that later proved essential for understanding Hamiltonian dynamics. This bracket captured the interdependence of position and momentum variables, laying groundwork for the geometric interpretation of phase spaces. Shortly thereafter, in 1834, developed his principal function, which formalized the for mechanical systems and led to Hamilton's , emphasizing the role of phase space as a geometric arena for dynamics. The mathematical formalization of these ideas accelerated in the 20th century through connections with differential geometry and group theory. In 1939, Hermann Weyl coined the term "symplectic" for the symplectic group, drawing from Greek roots to distinguish it from complex structures while highlighting its role in preserving volume in phase space transformations. Post-World War II, the French school advanced the concept significantly: in 1948, Charles Ehresmann and Paulette Libermann initiated the systematic study of symplectic manifolds as even-dimensional spaces equipped with compatible structures, with Libermann proving the Darboux theorem in the early 1950s to establish local canonical forms. Concurrently, in the 1950s, André Lichnerowicz and Jean-Louis Koszul contributed to the integration of symplectic structures with Poisson geometry and algebraic frameworks, formalizing them within broader differential and algebraic geometry contexts. The field experienced a major surge in the 1960s and 1970s, transitioning from to a vibrant area of known as . Vladimir I. Arnold's 1965 paper marked a pivotal moment by demonstrating topological obstructions to certain symplectic embeddings, inspiring applications to dynamical systems and caustics. This era also saw key works by Jean-Marie Souriau, , Victor Guillemin, , Alan Weinstein, and Jerrold Marsden, who developed concepts like moment maps and symplectic reduction, bridging geometry, physics, and . The modern phase began with Mikhail Gromov's 1985 introduction of pseudoholomorphic curves, revolutionizing the study of global symplectic invariants and rigidity phenomena.

Physical and Geometric Motivation

Symplectic manifolds originate from efforts to formalize , with roots in the early 19th-century works of Siméon-Denis Poisson and on systems. In this context, a symplectic manifold serves as the for a mechanical system, a 2n-dimensional space coordinatized by positions q_i and momenta p_i, where the dynamics of the system are governed by Hamilton's equations. This structure naturally encodes the evolution of physical states, such as the motion of particles under conservative forces. A key feature is , which states that the phase flow preserves the volume in this space, ensuring that the density of trajectories remains constant over time and reflecting the incompressibility of under evolution. Geometrically, the symplectic form on such a manifold—a closed, non-degenerate, antisymmetric 2-form—provides an way to measure oriented areas and volumes, distinct from the length and angle measurements afforded by Riemannian metrics. While a Riemannian defines distances and curvatures through a , the symplectic form emphasizes the pairing between positions and momenta, enabling a and derived from powers of the form itself. This invariance under symplectomorphisms preserves essential geometric features, such as the symplectic \omega^n / n!, which contrasts with the more rigid local invariants in . The symplectic structure directly ties to conservation laws in mechanics, as Hamiltonian flows—generated by a Hamiltonian function H—preserve the symplectic form, thereby maintaining the symplectic volume along trajectories. This preservation under the flow implies that quantities like are conserved when the Hamiltonian is time-independent, and it underpins the long-term of dynamical systems by preventing volume collapse or expansion. Such properties ensure that perturbations in mechanical systems do not lead to artificial , a crucial aspect for accurate modeling of reversible processes. Beyond mechanics, symplectic manifolds find applications in optics, where the cotangent bundle models light ray propagation and Hamiltonian ray tracing preserves phase space volumes for lens systems and beam dynamics. In celestial mechanics, they describe planetary orbits and perturbation theories, with symplectomorphisms capturing invariant tori in multi-body problems like the restricted three-body problem. Additionally, the framework bridges to quantum mechanics, where classical phase spaces inform semiclassical approximations and quantization procedures, facilitating transitions between classical and quantum descriptions of systems like the harmonic oscillator.

Definition and Basic Properties

Symplectic Form and Axioms

A symplectic manifold is defined as a pair (M, \omega), where M is a of even $2n and \omega is a smooth 2-form on M. This 2-form \omega serves as a section of the bundle of alternating 2-forms \Lambda^2 T^*M, ensuring compatibility with the of M. The defining axioms for \omega to be a symplectic form are closedness and non-degeneracy. Closedness requires that \omega is a closed differential form, meaning its exterior derivative vanishes: d\omega = 0. Non-degeneracy stipulates that, at every point p \in M, the interior product map v \mapsto \iota_v \omega_p induces an isomorphism from the tangent space T_p M to its dual T_p^* M, or equivalently, for every nonzero vector v \in T_p M, there exists a vector w \in T_p M such that \omega_p(v, w) \neq 0. This condition implies that the nth power of the 2-form, \omega^n, is nonzero at each point p, providing a volume form on M and confirming the even dimensionality. These axioms arise naturally in the context of phase spaces in , where \omega encodes the canonical for and variables.

Non-degeneracy and Closedness

The closedness condition on a symplectic form \omega, meaning d\omega = 0, ensures that \omega represents a well-defined [\omega] in the second group H^2(M; \mathbb{R}) of the manifold M. This class is a global invariant of the symplectic , capturing topological obstructions to the existence of \omega as an exact form on M. Locally, the closedness implies exactness by the Poincaré lemma: on any contractible open set U \subset M, there exists a 1-form \alpha such that \omega|_U = d\alpha. Non-degeneracy of \omega at each point p \in M means that the associated \omega_p: T_p M \times T_p M \to \mathbb{R} is skew-symmetric and invertible, i.e., if \omega_p(v, w) = 0 for all w \in T_p M, then v = 0. This invertibility follows from the map v \mapsto \iota_v \omega_p being an T_p M \to T_p^* M, which requires \dim M = 2n to be even, as skew-symmetric non-degenerate forms exist only on even-dimensional spaces. A key consequence of non-degeneracy is the existence of a compatible almost complex structure J on M, which satisfies J^2 = -\mathrm{Id} and pairs with \omega to define a Riemannian metric g via \omega(X, Y) = g(JX, Y) for vector fields X, Y, with g positive definite. Such a J is \omega-compatible, meaning \omega(JX, JY) = \omega(X, Y) and \omega(X, JX) > 0 for X \neq 0. Together, closedness and non-degeneracy imply that the n-fold wedge product \omega^n is a nowhere-vanishing on the $2n-dimensional manifold M, specifically \frac{\omega^n}{n!} provides a non-zero top-degree form at every point. On a compact symplectic manifold, this yields a finite total volume \int_M \frac{\omega^n}{n!} < \infty, endowing M with a orientation and Liouville measure preserved by symplectomorphisms.

Fundamental Theorems

Darboux Theorem

The Darboux theorem asserts that if (M, \omega) is a symplectic manifold of dimension $2n, then for every point p \in M, there exists a neighborhood Uofpand local coordinates(q^1, \dots, q^n, p_1, \dots, p_n)onU$ such that \omega = \sum_{i=1}^n \, dq^i \wedge dp_i on U. These coordinates are known as Darboux coordinates, and the corresponding symplectomorphism to a standard open set in \mathbb{R}^{2n} with the canonical symplectic form highlights the absence of local invariants in symplectic geometry beyond the manifold's dimension. This result was established by Gaston Darboux in his 1882 memoir addressing the Pfaff problem in differential geometry. Darboux's original proof, set within the context of integrable Pfaffian systems, demonstrated the local normal form for non-degenerate closed 2-forms, laying foundational groundwork for modern symplectic geometry. A sketch of the proof begins with the closedness of \omega (d\omega = 0), which, by the Poincaré lemma on contractible neighborhoods, guarantees the existence of a local 1-form \alpha satisfying \omega = d\alpha. The non-degeneracy of \omega then enables the construction of Darboux coordinates through flows generated by Hamiltonian vector fields: starting from coordinates where \omega matches the standard form at p, time-dependent Hamiltonian flows adjust the 1-form \alpha step-by-step to the canonical \sum p_i \, dq^i, yielding the desired local symplectomorphism. The theorem implies that all symplectic manifolds of the same are locally symplectomorphic to one another, underscoring a profound uniformity in their local structure that contrasts sharply with , where local invariants like the distinguish manifolds. This local indistinguishability facilitates global constructions and analyses in symplectic topology by reducing problems to standard models.

Moser Isotopy Theorem

The Moser isotopy theorem asserts that on a compact manifold M of dimension $2n, if \omega_t for t \in [0,1] is a family of forms such that the class [\omega_t] \in H^2(M; \mathbb{R}) is independent of t, then there exists a of \phi_t: M \to M with \phi_0 = \mathrm{id} satisfying \phi_t^* \omega_t = \omega_0 for all t \in [0,1]. This implies that \phi_1: (M, \omega_1) \to (M, \omega_0) is a , classifying such structures up to global within their class. The proof proceeds via Moser's trick, constructing a time-dependent X_t on M whose time-(t,0)- generates the desired \phi_t. Differentiating \phi_t^* \omega_t = \omega_0 with respect to t yields the equation \frac{d}{dt} (\phi_t^* \omega_t) = \phi_t^* \left( \mathcal{L}_{X_t} \omega_t + \frac{\partial}{\partial t} \omega_t \right) = 0, where \mathcal{L}_{X_t} denotes the along X_t = \frac{d\phi_t}{dt} \circ \phi_t^{-1}. Since each \omega_t is closed, \frac{\partial}{\partial t} \omega_t is also closed, and the constant class ensures [\frac{\partial}{\partial t} \omega_t] = 0, so \frac{\partial}{\partial t} \omega_t = d\mu_t for some 1-form \mu_t obtained via a (e.g., using a on the compact manifold to integrate the closed form). The non-degeneracy of \omega_t then guarantees a unique solution to the equation i_{X_t} \omega_t = -\mu_t, as \omega_t provides an between vector fields and 1-forms. The of M ensures the time-dependent of X_t exists and is , yielding the . A key application is the stability of symplectic structures: if \omega_1 is sufficiently C^\infty-close to \omega_0, the convex path \omega_t = (1-t)\omega_0 + t \omega_1 consists of symplectic forms (as small perturbations preserve non-degeneracy), so the theorem provides a diffeotopy deforming \omega_1 to \omega_0, demonstrating robustness under perturbations. This stability underpins the classification of compact symplectic manifolds up to symplectomorphism by their cohomology class and facilitates techniques like symplectic reduction. The Darboux theorem emerges as a local analogue of this global isotopy result.

Examples

Symplectic Vector Spaces

A is a pair (V, \omega), where V is a finite-dimensional over \mathbb{R} of even $2n, and \omega: V \times V \to \mathbb{R} is a skew-symmetric that is non-degenerate, meaning that if \omega(v, \cdot) = 0 for all v \in V, then v = 0. This non-degeneracy implies that the map v \mapsto \omega(v, \cdot) is an from V to its V^*. Such spaces form the algebraic foundation for the local structure of symplectic manifolds, modeling their spaces at each point. In coordinates, every admits a symplectic basis \{e_1, \dots, e_n, f_1, \dots, f_n\} such that \omega(e_i, f_j) = \delta_{ij} and \omega(e_i, e_j) = \omega(f_i, f_j) = 0. Relative to this basis, the standard form is expressed as \omega = \sum_{i=1}^n e_i^* \wedge f_i^*, where \{e_i^*, f_i^*\} is the dual basis. Identifying V with \mathbb{R}^{2n} via this basis, \omega takes the concrete form \sum_{i=1}^n dq_i \wedge dp_i, where q_i and p_i are the coordinate functions corresponding to the e_i and f_i. This Darboux-like representation highlights the form's role in . The linear automorphisms of V that preserve \omega form the symplectic group \mathrm{Sp}(2n, \mathbb{R}), defined as the subgroup of \mathrm{GL}(2n, \mathbb{R}) consisting of matrices A satisfying A^T J A = J, where J = \begin{pmatrix} 0 & I_n \\ -I_n & 0 \end{pmatrix} is the standard . This group acts transitively on the set of symplectic bases, ensuring that all symplectic vector spaces of the same are isomorphic. A key property of symplectic vector spaces is the existence of subspaces, which are maximal isotropic subspaces of n. A subspace L \subseteq V is isotropic if \omega(u, v) = 0 for all u, v \in L, and it is maximal if no larger subspace satisfies this condition; equivalently, L^\perp = L, where L^\perp = \{ w \in V \mid \omega(w, l) = 0 \ \forall l \in L \}. For example, the subspace spanned by \{e_1, \dots, e_n\} is with respect to the standard form. Every isotropic subspace has at most n, and any two subspaces are related by a .

Cotangent Bundles

The cotangent bundle of any smooth manifold Q of dimension n provides a canonical example of a symplectic manifold of dimension $2n. The total space T^*Q inherits a natural symplectic structure from the bundle construction itself, independent of any additional data on Q. This structure arises from the tautological 1-form \theta on T^*Q, defined intrinsically as follows: for a point p = (q, \xi) \in T^*Q with q \in Q and \xi \in T_q^*Q, and for any tangent vector v \in T_p(T^*Q), \theta_p(v) = \xi(d\pi_p(v)), where \pi: T^*Q \to Q is the bundle projection map. The associated symplectic form is then \omega = -d\theta, which is closed (d\omega = 0) and non-degenerate on T^*Q, endowing it with the required symplectic axioms. In local coordinates on T^*Q, choose coordinates (q^i)_{i=1}^n on an open set in Q; the induced coordinates on the bundle are (q^i, p_i)_{i=1}^n, where the p_i parameterize the cotangent fibers. The tautological 1-form takes the expression \theta = \sum_{i=1}^n p_i \, dq^i, and the symplectic form is \omega = \sum_{i=1}^n dq^i \wedge dp_i. These coordinate expressions confirm the closedness and non-degeneracy locally, as the matrix of \omega with respect to the basis \{dq^i, dp_i\} is the standard block-diagonal form with J = \begin{pmatrix} 0 & I_n \\ -I_n & 0 \end{pmatrix}, which has full rank. The form \omega is globally well-defined and independent of the choice of coordinates, ensuring the symplectic structure is canonical. Key submanifolds of (T^*Q, \omega) exhibit Lagrangian properties. The zero section, embedded as q \mapsto (q, 0) \in T^*Q, is a submanifold diffeomorphic to Q, since the pullback of \omega to it vanishes identically (as \theta restricts to zero there, so d\theta does as well). Each fiber \pi^{-1}(q) \cong T_q^*Q is a affine subspace, as \omega also restricts to zero on tangent vectors within the fiber (which lie in the dp_i directions, where dq^i = 0). These examples illustrate how the of T^*Q naturally accommodates maximal-dimensional isotropic submanifolds. In , T^*Q models the of a with space Q, where points (q, p) represent generalized positions and conjugate momenta. The canonical symplectic form \omega dictates the dynamics through Hamilton's equations for a function H: T^*Q \to \mathbb{R}, such as the kinetic-plus-potential energy H(q, p) = \frac{1}{2} |p|^2 + V(q), yielding \dot{q}^i = \frac{\partial H}{\partial p_i} and \dot{p}_i = -\frac{\partial H}{\partial q^i}. This framework unifies the geometric and variational aspects of systems.

Kähler Manifolds

A is a (M, J) equipped with a Hermitian metric g such that the associated Kähler form \omega(X, Y) = g(JX, Y) is closed, d\omega = 0, thereby endowing M with a structure inherited from its . Locally, in holomorphic coordinates \{z^j\}, the metric g takes the form g = \sum g_{j\bar{k}} dz^j \otimes d\bar{z}^k with g_{j\bar{k}} = \overline{g_{k\bar{j}}}, and the Kähler form is expressed as \omega = \frac{i}{2} \sum_{j,k} g_{j\bar{k}} \, dz^j \wedge d\bar{z}^k, where the coefficients g_{j\bar{k}} form a positive definite , ensuring the form is of type (1,1). The closedness of \omega follows from the relation \omega = \frac{i}{2} \partial \bar{\partial} \phi for some local Kähler potential \phi, as \partial \bar{\partial} + \bar{\partial} \partial = 0. The triple (g, J, \omega) is compatible in the sense that g is a Riemannian metric invariant under J, J is an integrable almost complex structure defining the complex structure on M, and \omega satisfies \omega(JX, JY) = \omega(X, Y) for all vector fields X, Y, with g(X, Y) = \omega(X, JY). This compatibility implies that \omega is non-degenerate, as the positivity of g ensures \omega^n \neq 0 everywhere on the $2n-dimensional manifold.[17] The symplectic nature arises directly from the complex structure, distinguishing Kähler manifolds among symplectic ones by requiring integrability of J$. Prominent examples include complex tori, such as \mathbb{T}^{2n} = \mathbb{C}^n / \Lambda with the flat Hermitian induced from \mathbb{C}^n, yielding a constant symplectic form \omega = \sum dx_i \wedge dy_i. Another key example is complex projective space \mathbb{CP}^n, equipped with the Fubini-Study , whose Kähler form in affine coordinates U_0 = \{[1:z_1:\dots:z_n]\} is \omega_\mathrm{FS} = \frac{i}{2} \partial \bar{\partial} \log(1 + |z|^2), which is invariant under the U(n+1) and restricts to the standard form on linear subspaces. A fundamental property is that the cohomology class [\omega] \in H^2(M, \mathbb{R}) lies in the (1,1)-part of the H^{1,1}(M) and is positive, known as the Kähler class, which plays a central role in the and deformation of the manifold. This class remains fixed under small deformations of the complex structure while preserving the Kähler condition.

Submanifolds

Lagrangian Submanifolds

In a manifold (M^{2n}, \omega) of $2n, a L \subset M is called if it has n and the symplectic form pulls back to zero on L, that is, \iota^*\omega = 0, where \iota: L \to M denotes the . This condition implies that L is isotropic, meaning \omega vanishes on tangent vectors to L, and the dimension condition ensures it is maximal isotropic, or coisotropic as well. Equivalently, L is Lagrangian if, at every point p \in L, the tangent space T_p L is a Lagrangian subspace of the symplectic vector space (T_p M, \omega_p). A subspace V \subset T_p M is Lagrangian if it is isotropic (\omega_p(V, V) = 0) and has dimension n, which forces V to be equal to its symplectic orthogonal complement V^\omega = \{ w \in T_p M \mid \omega_p(v, w) = 0 \ \forall v \in V \}. Standard examples of submanifolds arise in cotangent bundles (T^*Q, \omega_{\rm can}), where \omega_{\rm can} = -d\theta is the canonical symplectic form and \theta is the tautological 1-form. The zero section Z = \{ (q, 0) \in T^*Q \} and each cotangent fiber T^*_q Q are Lagrangian, as \omega_{\rm can} vanishes on their tangent spaces. More generally, the graph of a closed 1-form \sigma: Q \to T^*Q, given by \Gamma_\sigma = \{ (q, \sigma_q) \in T^*Q \}, is Lagrangian whenever d\sigma = 0, since the pullback of \omega_{\rm can} to \Gamma_\sigma equals d\sigma. In particular, graphs of exact 1-forms df for smooth functions f: Q \to \mathbb{R} provide exact submanifolds. A fundamental local structure theorem for Lagrangian submanifolds is Weinstein's Lagrangian neighborhood theorem, which asserts that if L \subset (M, \omega) is a compact Lagrangian submanifold, then there exist neighborhoods U \subset M of L and V \subset T^*L of the zero section in the of L (equipped with its canonical symplectic form), together with a \phi: (V, \omega_{\rm can}) \to (U, \omega) that extends the inclusion of the zero section into T^*L. This result, originally proved in 1971, shows that every Lagrangian submanifold is locally symplectomorphic to the zero section of its own , providing a normal form for neighborhoods of .

Special Lagrangian Submanifolds

In a Calabi-Yau manifold (M, \omega, g, J, \Omega), where M is a compact of complex dimension n with trivial and \Omega a nowhere-vanishing holomorphic normalized so that |\Omega| = 1 with respect to the metric g, a special Lagrangian submanifold is an n-dimensional oriented submanifold L \subset M that is (meaning \omega|_L = 0) and calibrated by the real part of \Omega, satisfying \operatorname{Re}(\Omega)|_L = \operatorname{vol}_g(L), where \operatorname{vol}_g denotes the Riemannian volume form on L induced by g. This calibration condition implies that L is a \phi-submanifold for \phi = \operatorname{Re}(\Omega), a closed n-form of comass 1 on M. Special Lagrangian submanifolds minimize area among all submanifolds homologous to them, as calibrated submanifolds are absolutely volume-minimizing in their class. They are minimal submanifolds, meaning their vanishes, and they remain stable under small deformations within the space of , providing critical points for the area functional. The deformation theory of special Lagrangians is obstructed only by the Maslov class, and their moduli spaces are expected to be of the expected when unobstructed. The Thomas-Yau conjecture posits that in a Calabi-Yau manifold, a is \sigma-stable (for some phase \sigma) if and only if it admits a special representative in its class, and that the starting from such a converges to a special in finite time. This stability condition is analogous to \mu-stability in and governs the existence and uniqueness of unobstructed special Lagrangians, with implications for mirror symmetry and enumerative invariants. Representative examples include the real locus \mathbb{R}^n \subset \mathbb{C}^n, defined by \operatorname{Im}(z_i) = 0 for i=1,\dots,n, which is a flat special calibrated by \operatorname{Re}(dz_1 \wedge \cdots \wedge dz_n). Another example is the Clifford in the Calabi-Yau (\mathbb{C}^n / \Lambda, \operatorname{Re}(\Omega)), given by the product of circles \prod_{i=1}^n \{z_i : |z_i| = 1/\sqrt{n}\} modulo the \Lambda = \mathbb{Z}^{2n}, which is special Lagrangian for the standard holomorphic form.

Other Coisotropic and Isotropic Submanifolds

In , an isotropic submanifold L of a manifold (M^{2n}, \omega) is a such that the symplectic form \omega vanishes when restricted to the at every point, i.e., \omega|_ {T_p L} \equiv 0 for all p \in L, or equivalently, T_p L \subseteq (T_p L)^\omega where (T_p L)^\omega = \{ v \in T_p M \mid \omega(v, w) = 0 \ \forall w \in T_p L \}. This condition implies that the dimension of L satisfies \dim L \leq n. Isotropic submanifolds arise naturally in contexts such as the orbits of torus actions on manifolds. A coisotropic submanifold C satisfies the dual condition (T_p C)^\omega \subseteq T_p C for all p \in C, which is equivalent to the conormal bundle N^* C being contained in T C, or the of T_p C in T_p^* M lying inside T_p C. Consequently, \dim C \geq n. For coisotropic submanifolds, the distribution (T C)^\omega is integrable, defining a characteristic whose leaves are symplectic submanifolds of C, known as symplectic leaves; the quotient space C / ((T C)^\omega) inherits a symplectic structure induced by \omega. Examples of isotropic submanifolds include points in M, where the tangent space is zero-dimensional and thus trivially satisfies \omega \equiv 0, and one-dimensional submanifolds (curves) in any symplectic manifold, as their tangent spaces are one-dimensional and \omega pairs vectors to scalars. In the standard symplectic vector space \mathbb{R}^{2n} with the canonical form, the subspace spanned by the first two standard basis vectors e_1, e_2 is isotropic. For coisotropic submanifolds, codimension-one subspaces (hypersurfaces) in \mathbb{R}^{2n} qualify, as their symplectic orthogonal is one-dimensional and contained within them due to the nondegeneracy of \omega. In low dimensions, such as n=1 where M is a surface, any hypersurface is coisotropic by dimension. Coisotropic and isotropic submanifolds are central to the study of Dirac structures, which provide a unified framework for symplectic and Poisson geometries within generalized complex geometry. A Dirac structure on M is a maximal isotropic subbundle of the Courant algebroid TM \oplus T^*M that is closed under the Courant bracket; in the symplectic case, the graph of \omega defines such a structure. Submanifolds whose normal bundles are isotropic or coisotropic with respect to the tangent Dirac structure induced by this framework characterize Dirac submanifolds, enabling extensions of symplectic reduction and embeddings to more general settings. This connection facilitates the analysis of coisotropic embeddings and their role in Dirac manifolds.

Advanced Structures

Lagrangian Fibrations

A Lagrangian fibration on a symplectic manifold (M, \omega) of dimension $2nis defined as a surjective submersion\pi: M \to Bonto ann-dimensional manifold B, such that the connected components of the regular fibers \pi^{-1}(b)are [Lagrangian](/page/Lagrangian) submanifolds ofM.[26] These fibrations often come equipped with a global section, allowing Mto be viewed as a [fiber bundle](/page/Fiber_bundle) overB$ with fibers. The fibers are typically compact tori in regular regions, reflecting the structure of integrable systems. Lagrangian fibrations are closely tied to completely integrable Hamiltonian systems on (M, \omega). In such systems, a momentum map F = (f_1, \dots, f_n): M \to \mathbb{R}^n consisting of n Poisson-commuting Hamiltonians \{f_i, f_j\} = 0 with linearly independent differentials almost everywhere induces a Lagrangian fibration via the projection to \mathbb{R}^n, where the level sets F^{-1}(c) form the Lagrangian fibers. Near a regular value c \in \mathbb{R}^n with compact connected fiber, the Liouville-Arnold theorem guarantees the existence of action-angle coordinates: local symplectic coordinates (I_1, \dots, I_n, \theta_1, \dots, \theta_n) where the I_k = f_k are the action variables (constants along the flow) and the \theta_k are angle variables parametrizing the n-torus fiber. Prominent examples include the canonical projection of the (T^*B, \Omega_{\mathrm{can}}) \to B for any manifold B, where the zero section provides a natural choice of Lagrangian fibers diffeomorphic to B itself, though typically non-compact unless B is. Another class arises from moment maps in symplectic toric manifolds: for a compact connected $2n-dimensional symplectic manifold admitting an effective Hamiltonian T^n-action, the moment map \mu: M \to \mathbb{R}^nyields a Lagrangian fibration withn$-torus fibers over the interior of the image (a Delzant polytope), classifying such manifolds up to equivariant symplectomorphism. Singularities in Lagrangian fibrations occur at critical values of \pi, disrupting the regular torus structure. A key type is the focus-focus singularity, characterized locally by a fiber consisting of two Lagrangian disks intersecting transversely at a pinch point, as modeled by the map (z,w) \mapsto zw from \mathbb{C}^2 to \mathbb{C}. These singularities induce non-trivial monodromy in the fibration: as one encircles a focus-focus critical value in the base, the homology cycles of nearby torus fibers undergo a shear transformation, obstructing global action-angle coordinates and reflecting quantum mechanical effects like those in the .

Symplectic Reduction

Symplectic reduction is a fundamental technique in for constructing new symplectic manifolds from those possessing symmetries, particularly through the quotienting of level sets of momentum maps associated to group actions. In the context of a symplectic manifold (M, \omega) equipped with a Hamiltonian action of a G, the momentum map \mu: M \to \mathfrak{g}^* encodes the symmetries, allowing the reduction process to eliminate redundant while preserving the symplectic structure. This method, known as Marsden-Weinstein reduction, yields a reduced space that inherits a natural symplectic form, facilitating the study of dynamical systems with conserved quantities. The Marsden-Weinstein reduction theorem states that if G acts freely and properly on the level set \mu^{-1}(0) \subset M, then the reduced space \mu^{-1}(0)/G is a smooth symplectic manifold with the reduced symplectic form \omega_{\mathrm{red}} defined as follows: for \in \mu^{-1}(0)/G and tangent vectors X_{\mathrm{red}}, Y_{\mathrm{red}} at $$, choose representatives \tilde{X}, \tilde{Y} \in T_p M that are tangent to a G-invariant slice through p, and set \omega_{\mathrm{red}}()(X_{\mathrm{red}}, Y_{\mathrm{red}}) = \omega_p(\tilde{X}, \tilde{Y}). This form is well-defined, independent of the choice of slice, and closed, making (\mu^{-1}(0)/G, \omega_{\mathrm{red}}) symplectic. The theorem assumes the action is symplectic and Hamiltonian, with \mu equivariant, ensuring the reduced dynamics correspond to the original via the symplectic quotient. A classic example arises in the , where the is the cotangent bundle T^*\mathbb{R}^3 \setminus \{0\} with the standard symplectic form, and the describes a particle in a $1/r potential under SO(3) . The map \mu corresponds to , and at \mu^{-1}(0) (zero angular momentum, or radial motion) yields a singular reduced space topologically equivalent to a , capturing the radial dynamics of the system. Further SO(4) hidden at non-zero levels produces a spherical S^2 \times S^2, capturing the bounded orbits as flow on this . This reveals the integrability and closed orbit structure of the system. Another prominent example is the of coadjoint orbits as reduced spaces. For a G acting on its dual \mathfrak{g}^* via the coadjoint action, the momentum map for the lifted action on T^*G identifies coadjoint orbits O_\xi = \mathrm{Ad}^*_G \xi \subset \mathfrak{g}^* with \mu^{-1}(\xi)/G_\xi, where G_\xi is the co-stabilizer. The induced Kirillov-Kostant-Souriau symplectic form on O_\xi is given by \omega_\xi(\hat{X}, \hat{Y}) = -\xi([X, Y]) for generators \hat{X}, \hat{Y}, making these orbits the basic building blocks for the orbit method in . More generally, can occur at coadjoint values \xi \in \mathfrak{g}^* where the coadjoint is locally , yielding the reduced space \mu^{-1}(\xi)/G_\xi as a symplectic manifold. Here, G_\xi is the stabilizer subgroup, and the is defined analogously on slices transverse to the G_\xi-orbits, preserving the symplectic structure for arbitrary levels beyond the zero level. This extension applies to systems with non-trivial conserved momenta, such as in or .

Hamiltonian Group Actions

A Lie group G acts symplectically on a symplectic manifold (M, \omega) if the action preserves the symplectic form \omega, meaning that the induced diffeomorphisms satisfy g^* \omega = \omega for all g \in G. Such an action is called if there exists a smooth map \mu: M \to \mathfrak{g}^*, called the momentum map, where \mathfrak{g}^* is the dual of the \mathfrak{g} of G, satisfying the condition d \langle \mu, \xi \rangle = -\iota_{X_\xi} \omega for every \xi \in \mathfrak{g}, with X_\xi denoting the infinitesimal generator of the action corresponding to \xi and \iota the interior product. The momentum map \mu is equivariant with respect to the coadjoint action of G on \mathfrak{g}^*, meaning \mu(g \cdot m) = \mathrm{Ad}^*_g \mu(m) for all g \in G and m \in M, where \mathrm{Ad}^* is the coadjoint representation. This equivariance ensures that the Hamiltonian action integrates the symmetries of the system consistently with the Poisson structure induced by \omega. A prominent example of a Hamiltonian group action is the action of the torus T^n on a compact toric symplectic manifold, where the momentum map \mu: M \to \mathbb{R}^n has an image that is a convex polytope, as established by the Atiyah-Guillemin-Sternberg convexity theorem. Another example is the linear action of the symplectic group \mathrm{Sp}(2n, \mathbb{R}) on \mathbb{R}^{2n} equipped with the standard symplectic form, where the momentum map is given by \langle \mu(x), \xi \rangle = \frac{1}{2} \omega(x, \xi x) for x \in \mathbb{R}^{2n} and \xi \in \mathfrak{sp}(2n, \mathbb{R}), satisfying the defining condition for each element of the Lie algebra. The Duistermaat-Heckman provides a key quantitative result for actions of compact groups on compact manifolds, stating that the under the momentum map of the normalized Liouville measure (defined by \omega^n / n!) is a on \mathfrak{g}^* of at most \dim M / 2 - \dim G, with the volume of the reduced spaces at regular values varying linearly across chambers separated by walls of codimension one. This highlights the nature of the measure on the reduced phase spaces and has implications for quantization and .

Generalizations and Extensions

Almost Symplectic Manifolds

An almost symplectic manifold is a smooth even-dimensional manifold M^{2n} equipped with a non-degenerate 2-form \omega, meaning that for every point p \in M, the interior product map v \mapsto i_v \omega_p: T_p M \to T_p^* M is a isomorphism, but \omega is not required to be closed (d\omega \neq 0). This structure relaxes the closedness axiom of standard symplectic manifolds while preserving non-degeneracy, which ensures that \omega induces a on the tangent spaces with maximal . Such manifolds arise naturally in contexts like nonholonomic , where the failure of closedness reflects constraints that prevent exact . The non-degeneracy of \omega allows the construction of compatible almost complex structures J on M. Specifically, there exists a Riemannian metric g such that g(u, v) = \omega(u, J v), J^2 = -Id, and \omega(J u, J v) = \omega(u, v) for all tangent vectors u, v, forming an almost Hermitian triple (\omega, J, g). The space of such compatible almost complex structures is non-empty and contractible on compact manifolds. Integrability of J—meaning it defines a complex structure—is obstructed by the Nijenhuis tensor N_J, defined by N_J(X, Y) = [X, Y] + J[JX, Y] + J[X, JY] - [JX, JY] for vector fields X, Y; N_J = 0 if and only if J is integrable, by the Newlander–Nirenberg theorem. In the almost Hermitian setting induced by \omega, N_J relates to d\omega through the torsion of the canonical connection: the (3,0) + (0,3) components of d\omega contribute to the skew-symmetric part of N_J, providing an obstruction to both complex and symplectic integrability. A canonical example of almost symplectic manifolds are nearly Kähler manifolds, which are almost Hermitian manifolds satisfying (\nabla_X J)X = 0 for all vector fields X, where \nabla is the Levi-Civita connection. Equivalently, d\omega = 3\lambda \operatorname{Re}(\Omega) for some constant \lambda \neq 0 and fundamental (3,0)-form \Omega, ensuring d\omega \neq 0 while maintaining non-degeneracy; the Nijenhuis tensor is totally skew-symmetric and non-vanishing in non-Kähler cases, such as the standard nearly Kähler structure on S^6. These structures appear in supersymmetry and calibrated geometry, highlighting how the specific form of d\omega encodes obstructions to both closedness and integrability. Almost symplectic structures can be deformed to genuine symplectic ones via averaging techniques when a suitable Lie group action exists, such as averaging \omega over the orbits to produce a closed invariant form in the same cohomology class, provided the action is Hamiltonian-like and d\omega lies in the image of the Lie derivative operator. This method leverages the geometry of group actions to resolve the closedness obstruction while preserving non-degeneracy locally.

Contact Manifolds

A contact manifold is an odd-dimensional smooth manifold N of dimension $2n+1 equipped with a contact structure, which can be defined via a contact 1-form \alpha, a nowhere-vanishing differential 1-form satisfying the non-degeneracy condition \alpha \wedge (d\alpha)^n \neq 0. This condition ensures that the kernel of \alpha, denoted \xi = \ker \alpha, is a hyperplane distribution on which d\alpha restricts to a symplectic form, making contact structures odd-dimensional analogues of symplectic structures. Two contact 1-forms \alpha and \alpha' define the same contact structure if \alpha' = f \alpha for some positive smooth function f: N \to \mathbb{R}^+. Contact manifolds relate to symplectic manifolds through the process of symplectization, which constructs an even-dimensional manifold from a one. Specifically, the symplectization of the manifold (N, \alpha) is the product manifold N \times \mathbb{R} equipped with the symplectic form \omega = d(e^t \alpha), where t is the coordinate on \mathbb{R}; this \omega is conformally symplectic and compatible with the contact structure in the sense that the hyperplanes \xi lift to the kernel of the primitive 1-form e^t \alpha. This construction provides a bridge between and symplectic geometry, allowing techniques from one field to inform the other, such as embedding problems or dynamical properties. A canonical example of a contact manifold is the standard contact structure on \mathbb{R}^{2n+1} with coordinates (x_1, \dots, x_n, y_1, \dots, y_n, z), where the contact 1-form is \alpha = dz - \sum_{i=1}^n y_i \, dx^i. This form satisfies the non-degeneracy condition, as d\alpha = \sum_{i=1}^n dx^i \wedge dy^i restricts to the standard symplectic form on the contact planes \ker \alpha. The standard contact structure on the unit sphere S^{2n+1} \subset \mathbb{R}^{2n+2} is induced by restricting this form after identifying the sphere with the projectivization of the contact planes. Associated to any contact 1-form \alpha is the Reeb vector field, a unique nowhere-vanishing \xi on N characterized by the conditions \alpha(\xi) = 1 and \iota_\xi d\alpha = 0. The Reeb field is transverse to the contact distribution \xi = \ker \alpha and generates a flow that preserves the contact structure, playing a central role in the dynamics of contact manifolds, such as in the study of periodic orbits. For the standard contact form on \mathbb{R}^{2n+1}, the Reeb vector field is simply \partial_z, reflecting the translational invariance in the z-direction.

Poisson Manifolds

A Poisson manifold is a smooth manifold M equipped with a bivector field \pi \in \Gamma(\wedge^2 TM) satisfying [\pi, \pi]_S = 0, where [ \cdot, \cdot ]_S denotes the Schouten-Nijenhuis bracket. This condition ensures that the associated bilinear map \{f, g\} = \pi(df, dg) on smooth functions C^\infty(M) defines a , which is skew-symmetric, satisfies the Leibniz rule \{f, gh\} = g\{f, h\} + h\{f, g\}, and obeys the . The bivector \pi induces a bundle map \pi^\sharp: T^*M \to TM given by \pi^\sharp(\alpha) = \iota_\alpha \pi, and the Poisson bracket generates Hamiltonian vector fields X_f = \pi^\sharp(df). The image of \pi^\sharp defines an integrable distribution on M, and its integral submanifolds, known as , are the connected components where \pi restricts to a non-degenerate . On each L, the rank of \pi is constant and even, and the inverse of the restricted yields a \omega_L such that \pi^\sharp_L = \omega_L^{-1}. Locally, around a point where the rank of \pi is constant, M decomposes as a product of a and a transverse manifold, with the leaves corresponding to the symplectic factor. This foliation generalizes the non-degenerate case of , allowing for degeneracy while preserving the Poisson bracket structure. Prominent examples include the \mathfrak{g}^* of a finite-dimensional \mathfrak{g}, endowed with the Lie- defined by \{f, g\}(\mu) = \langle \mu, [df(\mu), dg(\mu)] \rangle for \mu \in \mathfrak{g}^* and f, g \in C^\infty(\mathfrak{g}^*). Here, the leaves are the coadjoint orbits, which carry the Kirillov-Kostant-Souriau structure. Another example is the zero Poisson structure \pi = 0 on any smooth manifold, such as a , where the bracket vanishes identically \{f, g\} = 0, resulting in zero-dimensional leaves (points) and trivial dynamics. Casimir functions on a M are smooth functions C \in C^\infty(M) that Poisson-commute with every function, i.e., \{C, f\} = 0 for all f \in C^\infty(M), or equivalently, dC lies in the kernel of \pi^\sharp. Such functions are constant along the symplectic leaves, generating the center of the and serving as invariants that foliate M into level sets. In the Lie-Poisson case on \mathfrak{g}^*, the Casimirs are the invariant polynomials on \mathfrak{g}, constant on coadjoint orbits.

References

  1. [1]
    [PDF] Symplectic Manifolds
    Aug 31, 2019 · A symplectic structure on a manifold M is a closed, non-degenerate differential 2-form. The pair (M,ω2) is called a symplectic manifold.
  2. [2]
    [PDF] Lectures on Symplectic Geometry
    Page 11. Part I. Symplectic Manifolds. A symplectic form is a 2-form satisfying an algebraic condition – nondegeneracy – and an analytical condition – ...
  3. [3]
    [PDF] the works of Lagrange and Poisson during the years 1808–1810
    We analyse articles by Lagrange and Poisson written two hundred years ago which are the foundation of present-day symplectic and Poisson geometry. AMS MSC ...Missing: 1927 | Show results with:1927
  4. [4]
    [PDF] arXiv:1711.02440v1 [math.DG] 7 Nov 2017
    Nov 7, 2017 · The development of symplectic geometry since 1941 was dramatic, kept up first by the. French school (Charles Ehresmann, Paulette Libermann, ...
  5. [5]
    [PDF] Symplectic Geometry (Fall 2024)
    A couple of historical remarks:1. Symplectic geometry, as a subject of differential geometry, was developed in the 20th century. The 'symplectic group' was ...
  6. [6]
    [PDF] Early History of Symplectic Geometry
    Currently, symplectic geometry refers to the study of symplectic manifolds. A symplectic manifold is an even dimensional manifold endowed with a closed.
  7. [7]
    [PDF] Hamiltonian Mechanics and Symplectic Geometry
    Hamiltonian mechanics uses a Hamiltonian function to determine time evolution in phase space, and can be formulated on symplectic manifolds.
  8. [8]
    [PDF] Introduction to symplectic mechanics - HAL
    Mar 17, 2022 · Symplectic mechanics relates Hamiltonian mechanics to symplectic geometry, using manifolds with a special symplectic structure.
  9. [9]
    [PDF] Introduction to Symplectic and Hamiltonian Geometry Notes for a ...
    Symplectic geometry uses a closed nondegenerate 2-form. Hamiltonian geometry uses a moment map on symplectic manifolds, with quantities conserved by symmetries.
  10. [10]
    [PDF] Symplectic Geometry - University of Oregon
    A symplectic form on a manifold X is a closed, nondegenerate 2-form. A nondegenerate 2-form is a smoothly varying perfect pairing on the tangent spaces of X ...
  11. [11]
    [PDF] Symplectic Geometry (Fall 2024)
    2The classical Darboux theorem is a result on exterior differential systems. Page 6. 6. 2. Linear symplectic algebra. 2.1. Symplectic vector spaces. Let E be a ...
  12. [12]
    Sur le problème de Pfaff - EuDML
    Sur le problème de Pfaff. G. Darboux · Bulletin des Sciences Mathématiques et Astronomiques (1882). Volume: 6, Issue: 1, page 14-36; ISSN: 1155-8431 ...
  13. [13]
    [PDF] SYMPLECTIC GEOMETRY - Mathematics
    This implies that all intermediate powers ωk ∈ Λ2kV ∗ are also nonzero, 0 ≤ k ≤ n. The volume form ωn/n! is called the Liouville form associated with the.<|control11|><|separator|>
  14. [14]
    [PDF] Symplectic Geometry - arXiv
    May 17, 2005 · The classical proof of Darboux's theorem is by induction on the dimension of ... 2n-dimensional symplectic manifold, by Darboux's theorem (Theorem ...
  15. [15]
    Mathematical Methods of Classical Mechanics | SpringerLink
    In stockIn this text, the author constructs the mathematical apparatus of classical mechanics from the beginning, examining all the basic problems in dynamics.
  16. [16]
    [PDF] • | | SYMPLECTIC GEOMETRY, LECTURE 14 1. Kähler ...
    Let (M, ω, J) be a Kähler manifold, with ω a symplectic form and J an integrable complex structure com patible with ω. Compatibility ω(Ju, Jv) = ω(u, v): note ...
  17. [17]
    [PDF] 1. Kähler manifolds - UChicago Math
    Nov 20, 2013 · On a Kähler manifold, the symplectic form ω is harmonic. Furthermore, for any other harmonic form α, the product ω ∧ α is harmonic. Note ...
  18. [18]
    [PDF] LECTURE 2 1. Symplectic Manifolds 1.1. Basic definitions. 1.1. Recall
    A symplectic manifold is a pair (M,ω) where M is a manifold and ω a symplectic form. Example. Let M = R2n, or an open subset of R2n, with linear coordi- nates ...
  19. [19]
  20. [20]
    [PDF] Symplectic Manifolds and Their Lagrangian Submanifolds 329 - CORE
    The purpose of this paper is to generalize Darboux's theorem in several directions and to give some applications of the generalizations. The first direction of ...<|control11|><|separator|>
  21. [21]
    Calibrated geometries | Acta Mathematica
    Harvey, R., Lawson, H.B. Calibrated geometries. Acta Math 148, 47–157 (1982) ... Lagrangian Submanifolds · Normed Algebra · Lagrangian Plane · Use our pre ...
  22. [22]
    Special Lagrangians, stable bundles and mean curvature flow - arXiv
    Apr 19, 2001 · Abstract: We make a conjecture about mean curvature flow of Lagrangian submanifolds of Calabi-Yau manifolds, expanding on \cite{Th}.
  23. [23]
    [PDF] lecture 2: symplectic manifolds
    Definition 1.1. We call ω a symplectic form on M if. (1) (closeness) ω is a closed 2-form, i.e. dω = 0 (2) (non-degeneracy) for ∀p ∈ M, ωp is a linear ...Missing: primary sources
  24. [24]
  25. [25]
    [PDF] notes on lagrangian fibrations daniele sepe
    Definition 2.1. Let (M,ω) be a symplectic manifold. A Lagrangian fibration is a surjective map π : (M,ω) → B whose ...
  26. [26]
    [PDF] Symplectic Toric Manifolds
    SYMPLECTIC TORIC MANIFOLDS motion is a lagrangian fibration, i.e., it is locally trivial and its fibers are la- grangian submanifolds. The coordinates along ...
  27. [27]
    [PDF] reduction of symplectic manifolds with symmetry
    This symplectic manifold is important in fluid mechanics. See Arnold [2] and Ebin-Marsden [6]. Here the manifolds are Fréchet. Properly, one should use ...
  28. [28]
    [PDF] Symmetry reduction for central force problems* - Massey University
    Jul 6, 2016 · In this paper we take a closer look at Hamiltonian symmetry reduction applied to equations that are intimately familiar to every physicist.
  29. [29]
    [PDF] the very, very basics of hamiltonian actions on symplectic manifolds
    The map Φ is called a moment map for the action A of G on (M, ω) if it satisfies the following two properties. (1) The map Φ is G-equivariant, meaning that for ...
  30. [30]
    [PDF] the atiyah-guillemin-sternberg convexity theorem - UChicago Math
    Sep 9, 2010 · We present a proof, credited to Atiyah, Guillemin, and Sternberg, that investigates the properties of a Hamiltonian action of a torus Lie group, ...Missing: polytope | Show results with:polytope
  31. [31]
    [PDF] 1. Group actions on Symplectic manifolds - UT Math
    A Hamiltonian Tk action on (M,ω) is a Tk action on (M,ω) together with a “moment map” µ : M → Rk = tk ∗ such that: (1) if v ∈ Rk = tk, then the derivative of ...
  32. [32]
    [PDF] On the variation in the cohomology of the symplectic form of the ...
    J.J. Duistermaat and G.J. Heckman finite union of closed symplectic (V-)submanifolds of M~ of codimension >2. Because p~: M'r~ Y~M~ is a principal T ...
  33. [33]
    Hamiltonian vector fields on almost symplectic manifolds - arXiv
    Oct 30, 2012 · Almost symplectic manifolds may have few, non-zero, Hamiltonian vector fields or even none. Therefore, it is important to have examples and it ...
  34. [34]
    [PDF] 1. Almost-complex Structures
    Corollary 1. Any symplectic manifold has compatible almost-complex structures, and the space of such struc tures is path connected.
  35. [35]
    NEARLY KAHLER MANIFOLDS - Project Euclid
    In § 9 we determine differential forms which represent the Chern classes of a nearly Kahler manifold, or more generally any almost Hermitian manifold. Finally ...
  36. [36]
    [PDF] Contact Geometry - Universität zu Köln
    Any of these contact structures is called the standard contact structure on R2n+1. Example 2.10. The standard contact structure on the unit sphere S2n+1 in R2n+ ...
  37. [37]
    [PDF] LECTURE 7 MATH 242 1. Contact manifolds Recall a contact form λ ...
    Let (Y,λ) be a contact manifold. Define the symplectization to be the symplectic manifold. (Rs × Y,ω = d (esλ)) . We now check this ...
  38. [38]
    [PDF] NOTES FOR MATH 599: CONTACT GEOMETRY 1.1. Definitions ...
    INTRODUCTION. 1.1. Definitions and examples. Definition 1.1. A contact manifold (M,ξ) is a (2n + 1)-dimensional manifold M equipped with a smooth maximally ...
  39. [39]
    [PDF] LECTURE 6: CONTACT STRUCTURES MATH 242 1. Definitions ...
    Definition 3. If λ is a contact form, then the Reeb vector field R is the unique vector field such that dλ (R, −) = 0 and λ (R) = 1.
  40. [40]
    The local structure of Poisson manifolds - Project Euclid
    The local structure of Poisson manifolds. Alan Weinstein. Download PDF + Save to My Library. J. Differential Geom. 18(3): 523-557 (1983).
  41. [41]
    [PDF] the local structure of poisson - UC Berkeley math
    A Casimir function is constant along each symplectic leaf, and in a region where the rank is constant, the symplectic leaves are exactly the common level ...