Fact-checked by Grok 2 weeks ago

Hamiltonian vector field

In and physics, a Hamiltonian vector field is a X_H on a (M, \omega) associated to a smooth function H: M \to \mathbb{R}, called the Hamiltonian, and defined by the equation \iota_{X_H} \omega = -dH, or equivalently, \omega(X_H, \cdot) = -dH. This construction arises in the study of Hamiltonian mechanics and symplectic geometry, where the flow generated by X_H describes the time evolution of a physical system conserving both the symplectic structure \omega and the energy function H. In local coordinates on \mathbb{R}^{2n} with the standard symplectic form \omega = \sum_{i=1}^n dq_i \wedge dp_i, the components of X_H are given by \frac{dq_i}{dt} = \frac{\partial H}{\partial p_i} and \frac{dp_i}{dt} = -\frac{\partial H}{\partial q_i}, yielding Hamilton's canonical equations. The Hamiltonian vector field X_H is uniquely determined by the non-degeneracy of \omega, and its integral curves lie on the level sets of H, ensuring that H is constant along the flow: \mathcal{L}_{X_H} H = 0. Moreover, X_H preserves the form, satisfying \mathcal{L}_{X_H} \omega = 0, which classifies it as a . The collection of all Hamiltonian vector fields forms a under the , isomorphic to the of smooth functions on M equipped with the \{f, g\} = X_f(g) = \omega(X_f, X_g) = -X_g(f), satisfying [X_f, X_g] = -X_{\{f,g\}}. Not every is Hamiltonian; on a connected , the Hamiltonian ones correspond to exact symplectic vector fields, with the obstruction lying in the first group H^1(M; \mathbb{R}). Key examples include the on \mathbb{R}^2, where H(q, p) = \frac{p^2}{2m} + \frac{k q^2}{2} generates circular orbits via X_H = \frac{p}{m} \frac{\partial}{\partial q} - k q \frac{\partial}{\partial p}, illustrating volume-preserving and energy-conserving dynamics. In broader contexts, Hamiltonian vector fields underpin integrable systems, , and the study of symplectomorphisms, with applications extending to , , and . Their divergence-free nature in ensures on volume preservation.

Background Concepts

Symplectic Manifolds

A is a pair (M, \omega), where M is a smooth manifold and \omega is a closed nondegenerate 2-form on M. The closedness condition means that the satisfies d\omega = 0, ensuring that \omega defines a cohomology class in the of M. Nondegeneracy implies that for every point p \in M, the map v \mapsto \omega_p(v, \cdot) from the T_p M to its dual T_p^* M is an isomorphism. The nondegeneracy of \omega forces the dimension of M to be even, say \dim M = 2n, as the induced pairing on tangent spaces pairs distinct directions without fixed points. A fundamental local property is given by , which states that around any point in M, there exist coordinates (q^1, \dots, q^n, p_1, \dots, p_n) such that \omega = \sum_{i=1}^n dq^i \wedge dp_i. This canonical local expression highlights the "standard" form of symplectic structures, independent of global topology. In , symplectic manifolds naturally arise as phase spaces, particularly the T^*Q of a manifold Q, equipped with the symplectic form \omega = -d\theta. Here, \theta is the tautological 1-form defined by \theta_{(q,p)}(\delta q, \delta p) = p(\delta q), where (q,p) \in T^*Q and (\delta q, \delta p) \in T_{(q,p)}(T^*Q). This structure captures the of systems with q and momenta p, providing a geometric foundation for . The concept originated in during the late 19th century, with foundational contributions from in his studies of , and was formalized in by pioneers including , who coined the term "" in 1939 to describe the associated linear group.

Hamiltonian Functions

A function on a (M, \omega) is defined as a real-valued function H: M \to \mathbb{R}. Such functions serve as scalar potentials that generate dynamics on the manifold, bridging with . In the context of , the H typically represents the total energy of a system, comprising both kinetic and potential components, within the modeled by the . This formulation generalizes the energy from to a geometric setting, where H encodes the conserved quantities associated with the system's symmetries via . Locally, on any , the Darboux theorem guarantees the existence of (q^1, \dots, q^n, p_1, \dots, p_n) in which H can be expressed as a H(q, p). These coordinates reflect the natural separation into generalized positions q and momenta p, facilitating the analysis of mechanical systems in a coordinate-dependent manner. functions are inherently globally defined as maps on M, but the associated distinguish between locally and globally cases: global Hamiltonians correspond to exact 1-forms dH, while local ones arise when the generating 1-form is merely closed, requiring the first group H^1_{\mathrm{de\ Rham}}(M) = 0 for all symplectic vector fields to be globally .

Definition and Local Expression

Abstract Definition

In the context of a symplectic manifold (M, \omega), where \omega is a closed, non-degenerate 2-form, a Hamiltonian function H: M \to \mathbb{R} defines the associated Hamiltonian vector field X_H as the unique vector field satisfying \iota_{X_H} \omega = dH, with \iota denoting the interior product. The non-degeneracy of \omega ensures the uniqueness of X_H, as it induces a bundle isomorphism \flat_\omega: TM \to T^*M given by v \mapsto \iota_v \omega, which maps vector fields bijectively to exact 1-forms and thus inverts to yield a unique vector field from the exact 1-form dH. Since H is smooth, dH is a smooth 1-form, and the isomorphism \flat_\omega is smooth, it follows that X_H is a smooth vector field. There is a common sign convention variation in the literature, where some texts define X_H via \iota_{X_H} \omega = -dH, often when using the symplectic form \omega = \sum dp_i \wedge dq^i (the negative of the standard \sum dq_i \wedge dp_i), to match conventions in physics. This choice affects the direction of the flow but preserves the underlying symplectic structure and dynamics up to time reversal.

Canonical Coordinates

In , local coordinates that simplify the expression of the symplectic form are known as or Darboux coordinates. On a (M, \omega) of dimension $2n, guarantees the existence of a coordinate chart (U, (q^1, \dots, q^n, p_1, \dots, p_n)) around any point such that the symplectic form takes the standard expression \omega = \sum_{i=1}^n \, dq^i \wedge dp_i. This highlights the pairing between position-like coordinates q^i and momentum-like coordinates p_i, facilitating explicit computations of geometric objects like vector fields. The Hamiltonian vector field X_H associated to a smooth function H: M \to \mathbb{R} admits a concrete local expression in these . Assuming the abstract definition via the interior product \iota_{X_H} \omega = dH, the vector field takes the form X_H = \sum_{i=1}^n \left( \frac{\partial H}{\partial p_i} \frac{\partial}{\partial q^i} - \frac{\partial H}{\partial q^i} \frac{\partial}{\partial p_i} \right). This expression arises from the nondegeneracy of \omega, which uniquely determines X_H for each H. To derive this, substitute the assumed coordinate form of X_H = \sum_{i=1}^n (a_i \partial/\partial q^i + b_i \partial/\partial p_i) into the defining equation and compute the interior product with \omega. The interior product yields \iota_{X_H} \omega = \sum_{i=1}^n \left( a_i \, dp_i - b_i \, dq^i \right), since \omega pairs the basis forms as specified. Setting this equal to dH = \sum_{i=1}^n \left( \frac{\partial H}{\partial q^i} dq^i + \frac{\partial H}{\partial p_i} dp_i \right) and equating coefficients gives a_i = \partial H / \partial p_i (from dp_i) and -b_i = \partial H / \partial q^i so b_i = -\partial H / \partial q^i (from dq^i), confirming the explicit expression. This local computation makes the intrinsic definition tangible and is valid on any Darboux chart. Under transformations, which are preserving \omega, the transforms covariantly as a . Specifically, if \phi: M \to M is a symplectomorphism, then the \phi_* X_H is the Hamiltonian vector field associated to the transformed Hamiltonian H \circ \phi^{-1}. This ensures the geometric structure remains consistent across coordinate systems.

Examples

Simple Mechanical Systems

In , simple systems provide intuitive illustrations of Hamiltonian vector fields, where the is typically the of the configuration space with canonical symplectic structure. For the one-dimensional , the function is given by H(q, p) = \frac{p^2}{2m} + \frac{1}{2} k q^2, where q is the , p is the , m is the , and k is the spring constant. The corresponding Hamiltonian vector field is then X_H = \frac{p}{m} \frac{\partial}{\partial q} - k q \frac{\partial}{\partial p}, which arises from the standard expression in canonical coordinates X_H = \frac{\partial H}{\partial p} \frac{\partial}{\partial q} - \frac{\partial H}{\partial q} \frac{\partial}{\partial p}. This vector field describes oscillatory motion in phase space, with trajectories forming closed ellipses centered at the origin. For a free particle in one dimension, the Hamiltonian simplifies to the kinetic energy term alone, H(q, p) = \frac{p^2}{2m}, since there is no potential. The associated Hamiltonian vector field is X_H = \frac{p}{m} \frac{\partial}{\partial q}, indicating constant velocity motion parallel to the position axis in phase space, as momentum p remains fixed while position q evolves linearly with time. In systems with central forces, such as the Kepler problem describing planetary motion under inverse-square gravitation, polar coordinates in the plane are natural, with radial position r, radial momentum p_r, and conserved angular momentum L. The Hamiltonian becomes H(r, p_r) = \frac{p_r^2}{2m} + \frac{L^2}{2m r^2} + V(r), where V(r) = -\frac{k}{r} for the gravitational potential with k = G m_1 m_2. The Hamiltonian vector field X_H in these coordinates generates bounded elliptical orbits for negative energies, reflecting the effective potential combining centrifugal and attractive terms. In each case, the vector field X_H generates flows along trajectories that conserve the energy represented by H.

Symplectic Geometry Examples

In , coadjoint orbits provide a fundamental example of Hamiltonian vector fields arising from group actions. For a G acting on its coadjoint orbit \mathcal{O}_\lambda \subset \mathfrak{g}^* equipped with the Kostant-Kirillov-Souriau symplectic form, the action is Hamiltonian with moment map \mu: \mathcal{O}_\lambda \to \mathfrak{g}^* given by the inclusion \mu(m) = m. For an element \xi \in \mathfrak{g}, the function H_\xi = \langle \mu, \xi \rangle generates the infinitesimal generator X_{H_\xi} of the action, which is the fundamental vector field X_\xi tangent to the orbit. This vector field satisfies i_{X_{H_\xi}} \omega = -dH_\xi, where \omega is the symplectic form, illustrating how elements produce Hamiltonian flows preserving the orbit's symplectic structure. Another illustrative example occurs on cotangent bundles, which carry a structure. Consider the T^*N of a smooth manifold N with the standard form \omega = -d\theta, where \theta is the tautological 1-form. The H: T^*N \to \mathbb{R} defined by the H(q, p) = \frac{1}{2} g^{ij}(q) p_i p_j, corresponding to a Riemannian metric g on N, generates the flow X_H. The integral curves of X_H project to on N, and X_H satisfies Hamilton's equation i_{X_H} \omega = -dH, demonstrating the nature of dynamics in this geometric setting. Hamiltonian actions of tori or circles on symplectic manifolds yield vector fields that manifest as rotations in suitable coordinates. For instance, the circle group S^1 acts on the 2-sphere S^2 with symplectic form \omega = d\theta \wedge dh (where \theta is the azimuthal angle and h the height function) by rotations e^{it} \cdot (\theta, h) = (\theta + t, h). This action is Hamiltonian with moment map \mu = h, and the generating vector field X_h = \partial/\partial \theta satisfies i_{X_h} \omega = -dh, producing rotational flows around the vertical axis that preserve \omega. More generally, for a torus T^k acting effectively on a compact symplectic manifold (M, \omega) of dimension $2n, the action admits a moment map \mu: M \to \mathbb{R}^k if k \leq n, with each component generating a Hamiltonian vector field corresponding to circle subgroup rotations. A key non-example highlights the structural constraints: on an odd-dimensional manifold, no symplectic form exists, precluding the definition of Hamiltonian vector fields in this context. Specifically, a nondegenerate closed 2-form \omega on a manifold of dimension $2n+1 cannot exist, as the associated map TM \to T^*M induced by \omega would fail to be an due to mismatched s, rendering \omega degenerate. Consequently, the equation i_X \omega = dH admits no unique solution X for arbitrary smooth H, as the nondegeneracy required for well-defined Hamiltonian vector fields is absent.

Properties of Hamiltonian Vector Fields

Hamilton's Equations

The integral curves of a Hamiltonian vector field X_H on a (M, \omega) are smooth curves \gamma: I \to M, where I is an interval in \mathbb{R}, satisfying the \gamma'(t) = X_H(\gamma(t)) for all t \in I. These curves describe the trajectories of the generated by X_H, with the function H: M \to \mathbb{R} determining the direction of motion at each point via the relation \iota_{X_H} \omega = -dH. In (q^i, p_i) on a T^*Q, where the symplectic form takes the standard expression \omega = \sum_i dp_i \wedge dq^i, the components of X_H are given by X_H = \sum_i \left( \frac{\partial H}{\partial p_i} \frac{\partial}{\partial q^i} - \frac{\partial H}{\partial q^i} \frac{\partial}{\partial p_i} \right). To derive Hamilton's equations, consider the action of X_H on the coordinate functions: the derivative along an \gamma(t) = (q^i(t), p_i(t)) yields \frac{d}{dt} q^i(\gamma(t)) = X_H(q^i) = \frac{\partial H}{\partial p_i} and \frac{d}{dt} p_i(\gamma(t)) = X_H(p_i) = -\frac{\partial H}{\partial q^i}, since X_H(f) = \sum_i \left( \frac{\partial H}{\partial p_i} \frac{\partial f}{\partial q^i} - \frac{\partial H}{\partial q^i} \frac{\partial f}{\partial p_i} \right) for any smooth function f. Thus, the integral curves of X_H are precisely the solutions to Hamilton's equations: \begin{aligned} \frac{dq^i}{dt} &= \frac{\partial H}{\partial p_i}, \\ \frac{dp_i}{dt} &= -\frac{\partial H}{\partial q^i}, \end{aligned} which form a of $2n first-order ordinary differential equations on the $2n-dimensional . This equivalence establishes Hamilton's equations as the coordinate manifestation of the generated by X_H.

Conservation Laws

One key property of the Hamiltonian vector field X_H on a symplectic manifold (M, \omega) is that it preserves the Hamiltonian function H along its integral curves. Specifically, the Lie derivative of H with respect to X_H vanishes: \mathcal{L}_{X_H} H = 0. This implies that the time derivative of H along the flow of X_H is zero, \frac{dH}{dt} = 0, ensuring conservation of the Hamiltonian, which often represents the total energy in mechanical systems. The proof follows directly from the definition of the Hamiltonian vector field. By definition, \iota_{X_H} \omega = -dH, so the Lie derivative \mathcal{L}_{X_H} H = X_H(H) = dH(X_H) = -\omega(X_H, X_H). Since the symplectic form \omega is skew-symmetric, \omega(X_H, X_H) = 0, hence \mathcal{L}_{X_H} H = 0. More generally, any smooth function K on M that Poisson commutes with H, i.e., \{K, H\} = 0, is conserved along the flow of X_H. Indeed, the time evolution of K satisfies \frac{dK}{dt} = \{H, K\}, so \{K, H\} = 0 implies \frac{dK}{dt} = 0. Such functions K are called constants of motion or integrals of the system. This conservation arises from symmetries via in the Hamiltonian setting: if a one-parameter group of canonical transformations preserves the H, it generates a conserved quantity corresponding to the infinitesimal generator of the symmetry.

Symplectomorphism Generation

A vector field X_H on a (M, \omega) is defined such that its contraction with the symplectic form satisfies \iota_{X_H} \omega = -dH, where H is the function. This ensures that X_H preserves the symplectic structure infinitesimally, as the Lie derivative of \omega along X_H vanishes: \mathcal{L}_{X_H} \omega = 0. To see this, apply Cartan's magic formula for the of a : \mathcal{L}_{X_H} \omega = \iota_{X_H} d\omega + d(\iota_{X_H} \omega). Since \omega is closed, d\omega = 0, so the first term is zero. The second term simplifies to d(-dH) = -d^2 H = 0, as the of an exact form is zero. Thus, \mathcal{L}_{X_H} \omega = 0, confirming that X_H is a . This condition positions X_H as an infinitesimal generator of symplectomorphisms, tangent to the symplectomorphism group \mathrm{Sympl}(M, \omega) at the . The of all such symplectic vector fields forms a under the Lie bracket, with Hamiltonian vector fields comprising a distinguished . A key consequence is , which states that the flow of X_H preserves the Liouville \frac{\omega^n}{n!} on the $2n-dimensional manifold M. This follows because \mathcal{L}_{X_H} \left( \frac{\omega^n}{n!} \right) = \frac{1}{n!} \left( \mathcal{L}_{X_H} \omega \right) \wedge \omega^{n-1} + \cdots + \frac{1}{n!} \omega^n \wedge \left( \mathcal{L}_{X_H} \omega \right) = 0, since each term involves \mathcal{L}_{X_H} \omega = 0. This volume preservation is fundamental in for understanding incompressibility.

Poisson Brackets

Definition via Vector Fields

In the context of a symplectic manifold (M, \omega), where \omega is a closed non-degenerate 2-form, the Poisson bracket of two smooth functions f, g \in C^\infty(M) is defined using the associated Hamiltonian vector fields X_f and X_g. These vector fields are uniquely determined by the relations \iota_{X_f} \omega = -df and \iota_{X_g} \omega = -dg, where \iota denotes the interior product and d is the exterior derivative. The Poisson bracket is then given by \{f, g\} = \omega(X_f, X_g), which establishes a bilinear operation on the space of smooth functions that encodes the symplectic structure algebraically. This definition highlights the role of Hamiltonian vector fields in translating the geometric data of \omega into an algebraic bracket operation. The Poisson bracket satisfies several fundamental properties derived directly from the symplectic form and the linearity of the interior product. It is bilinear over \mathbb{R}, meaning that for constants a, b \in \mathbb{R} and functions g, h \in C^\infty(M), \{f, ag + bh\} = a \{f, g\} + b \{f, h\}, and similarly in the second argument. It is also skew-symmetric: \{f, g\} = -\{g, f\}, which follows from the antisymmetry of \omega. Additionally, it obeys the Leibniz rule, acting as a derivation in each argument: \{f, gh\} = g \{f, h\} + h \{f, g\}, reflecting the product rule for the underlying vector fields. These properties position the Poisson bracket as a key tool for analyzing the algebraic structure induced by the symplectic geometry. An important equivalence relates the to the action of vector fields as derivations. Specifically, \{f, g\} = X_f(g) = -X_g(f), where X_g(f) denotes the of f along X_g. This identity underscores that the bracket measures how one vector field differentiates another function, providing a coordinate-free of the . This definition arises naturally from the form via the interior product relations. Contracting \iota_{X_g} \omega = -dg with X_f yields \omega(X_g, X_f) = -dg(X_f) = -X_f(g), and since \{f, g\} = X_f(g) by the equivalence above while \omega(X_f, X_g) = - \omega(X_g, X_f), the bracket aligns directly with \omega(X_f, X_g). Symmetrically, applying \iota_{X_f} \omega = -df gives \omega(X_f, X_g) = -df(X_g) = -X_g(f), ensuring the is well-defined and of local coordinates.

Lie Algebra Structure

The Poisson bracket on the space of smooth functions C^\infty(M) on a symplectic manifold (M, \omega) defines a Lie algebra structure, with the bracket satisfying bilinearity, antisymmetry, and the Jacobi identity. Specifically, for any f, g, h \in C^\infty(M), \{f, \{g, h\}\} + \{g, \{h, f\}\} + \{h, \{f, g\}\} = 0. This identity holds because the Poisson bracket is derived from the symplectic form, ensuring the algebraic consistency required for a Lie bracket. The Jacobi identity for the Poisson bracket directly implies a corresponding structure on the Hamiltonian vector fields. The canonical map associating each function f to its Hamiltonian vector field X_f, defined by \iota_{X_f} \omega = -df, yields the relation [X_f, X_g] = -X_{\{f, g\}}, where [ \cdot, \cdot ] denotes the . This equality follows from the definition of the Poisson bracket as \{f, g\} = X_f(g) = \omega(X_f, X_g) and the properties of the interior product and , confirming that the image of Hamiltonian vector fields forms a Lie subalgebra of the of all smooth vector fields on M. The map \phi: C^\infty(M) \to \mathfrak{X}(M), \phi(f) = X_f, is a Lie algebra anti-homomorphism from (C^\infty(M), \{ \cdot, \cdot \}) to the Lie algebra of vector fields, with kernel consisting precisely of the constant functions, as constant functions generate the zero vector field. Thus, it induces a Lie algebra isomorphism between the quotient C^\infty(M)/\mathbb{R} (with the induced bracket) and the Lie algebra \mathfrak{ham}(M, \omega) of Hamiltonian vector fields. The adjoint representation in this Lie algebra acts infinitesimally on functions via \mathrm{ad}_f(g) = \{f, g\}, mirroring the action of the Lie bracket on the vector fields through the identification.

Hamiltonian Flows

Local Flows

The local flow of a Hamiltonian vector field X_H on a (M, \omega) is defined as a one-parameter family of diffeomorphisms \phi_t: U \to M, where U \subset M \times \mathbb{R} is an open set containing the zero section \{(x, 0) \mid x \in M\}, satisfying the initial value problem \frac{d}{dt} \phi_t(x) = X_H(\phi_t(x)), \quad \phi_0(x) = x for all x \in M. This flow describes the integral curves of X_H, which locally evolve points along the direction specified by the Hamiltonian H. The diffeomorphisms \phi_t preserve the symplectic structure locally, as the Lie derivative \mathcal{L}_{X_H} \omega = 0 ensures that each \phi_t pulls back \omega to itself. By the standard Picard-Lindelöf theorem for ordinary differential equations on manifolds, local existence and uniqueness of the flow hold whenever X_H is smooth, which it is since H is assumed smooth and \omega is nondegenerate. Specifically, for each initial point x \in M, there exists a maximal time interval I_x = (a_x, b_x) with $0 \in I_x and b_x > 0 such that the solution \phi_t(x) is defined and unique for t \in I_x, remaining within a compact subset of M where X_H is Lipschitz continuous. The flow is compact in the sense that it is defined only on this maximal interval, beyond which the solution may escape any compact set or approach the boundary of the domain if M is not complete. In the time-dependent case, where the Hamiltonian H = H(t, \cdot) varies with time, the associated vector field X_{H(t)} becomes time-dependent, generating a non-autonomous flow \phi_t satisfying \frac{d}{dt} \phi_t(x) = X_{H(t)}(\phi_t(x)), \quad \phi_0(x) = x. Local existence and uniqueness still follow from standard ODE theory, provided H(t, \cdot) is smooth in its spatial arguments uniformly in t over compact time intervals. The maximal interval of definition depends on both the spatial domain and the time variation of H(t), potentially shortening if the time dependence causes rapid growth in X_{H(t)}. This setup arises naturally in perturbed mechanical systems, where external time-varying forces modify the energy function.

Global Hamiltonian Vector Fields

A Hamiltonian vector field X_H on a (M, \omega) generates a if it is , meaning the \phi_t^H: M \to M is defined for all t \in \mathbb{R} and every point in M. ensures that curves extend indefinitely without singularities or escape in finite time. On symplectic manifolds, every smooth Hamiltonian is . This follows from the general fact that any smooth on a manifold generates a , as bounds the and prevents finite-time blow-up. Since Hamiltonian s arise from smooth Hamiltonians, and manifolds imply bounded Hamiltonians, the s remain confined within the manifold for all time. Cotangent bundles T^*Q of smooth manifolds Q provide prominent examples of spaces supporting Hamiltonian vector fields with global flows. Equipped with the canonical symplectic form \omega = -d\theta, where \theta is the tautological 1-form, smooth functions on T^*Q—such as mechanical Hamiltonians—yield globally defined X_H. For typical mechanical Hamiltonians on T^*Q with compact Q, the resulting X_H often generate complete flows. The Arnold-Liouville theorem illustrates global structure in integrable systems: for a completely integrable on a $2n-dimensional with n independent commuting Hamiltonians, if a regular is compact and connected, it is diffeomorphic to an n-torus, and global action-angle coordinates exist on a saturated neighborhood, linearizing the flows to constant speeds on the tori and ensuring completeness thereon.

References

  1. [1]
    [PDF] Hamiltonian Mechanics and Symplectic Geometry
    Definition 1 (Hamiltonian Vector Field). A vector field X that satisfies. LXω = 0 is called a Hamiltonian vector field and the space of such vector fields on ...
  2. [2]
    [PDF] to be completed. LECTURE NOTES 1. Hamiltonian Mechanics Let ...
    is called the Hamiltonian vector field, denoted by XH. In this case, {(q, p)} = R2n is called the phase space. Any (q, p) is called a state.
  3. [3]
    [PDF] lecture 6: geometry of hamiltonian systems
    Now suppose Ξf is the Hamiltonian vector field associated with f. The following properties are easily seen from the definition. One should be aware of the use ...
  4. [4]
    [PDF] Lectures on Symplectic Geometry
    We will now construct a major class of examples of symplectic forms. The canonical forms on cotangent bundles are relevant for several branches, including.<|separator|>
  5. [5]
    [PDF] Springer - Department of Mathematics | University of Toronto
    TAKEUTIIZARING. Introduction to. 34 SPITZER. Principles of Random Walk. Axiomatic Set Theory. 2nd ed. 2nded. 2 OXTOBV. Measure and Category. 2nd ed.
  6. [6]
    The Symplectic Piece - Ideas | Institute for Advanced Study
    The modern fields of symplectic geometry and dynamical systems originate in Henri Poincaré's work on celestial mechanics. The term “symplectic” was introduced ...
  7. [7]
    [PDF] Hamiltonian Mechanics and Symplectic Geometry
    Definition 1 (Symplectic Manifold). A symplectic manifold (M,ω) is a. 2n ... actually come from a Hamiltonian function. The equation. LXω = (diX + iXd)ω ...
  8. [8]
    Mathematical Methods of Classical Mechanics | SpringerLink
    In stockMathematical Methods of Classical Mechanics. Textbook; © 1989. Latest edition ... Arnold. Pages 3-14. Investigation of the equations of motion. V. I. Arnold.
  9. [9]
    [PDF] 4. The Hamiltonian Formalism - DAMTP
    In this section, we'll present a rather formal, algebraic description of classical dynamics which makes it look almost identical to quantum mechanics! We'll ...
  10. [10]
    [PDF] L03: Kepler problem & Hamiltonian dynamics - MIT Mathematics
    “Every object in the Universe attracts every other object with a force directed along a line of centres for the two objects that is proportional to the product ...
  11. [11]
    [PDF] Symplectic Toric Manifolds
    The motion generated by this vector field is rotation about the vertical axis, which of course preserves both area and height. 2. On the symplectic 2-torus (T2, ...
  12. [12]
  13. [13]
    (PDF) Foundations of Mechanics - Academia.edu
    See full PDF downloadDownload PDF. FOUNDATIONS OF MECHANICS SECOND EDITION RALPH ABRAHAM JERROLD E. MARSDEN AMS CHELSEA PUBLISHING American Mathematical ...
  14. [14]
    [PDF] SYMPLECTIC GEOMETRY Lecture Notes, University of Toronto
    Every Hamiltonian vector field is a symplectic vector field. That is ... What are the generating vector fields? The Lie algebra of the gauge group is ...
  15. [15]
    symplectic vector field in nLab
    Feb 26, 2012 · The flow generated by a symplectic vector field is an auto-symplectomorphism. Relation to Hamiltonian vector fields. By Cartan's magic formula ...Missing: generation | Show results with:generation
  16. [16]
    [PDF] Poisson and Symplectic structures, Hamiltonian action ... - arXiv
    Apr 1, 2020 · As the title suggests, the material covered here includes the Poisson and symplectic structures (Poisson manifolds, Poisson bi-vectors and ...
  17. [17]
    [PDF] symplectic manifolds - Texas Christian University
    Nondegeneracy at a point implies in particular that the dimension of the manifold is even (say dim = 2n). In fact, nondegeneracy is equivalent to βn = β ∧ ... ∧ ...
  18. [18]
    [PDF] Hamiltonian One Parameter Groups
    For example, we have the implicit mapping theorem and the existence and uniqueness theorem for flows of vectorfields. (See LANG [1] or DIEUDONNة [1] ...
  19. [19]
    On time-dependent Hamiltonian realizations of planar and ...
    This essentially removes the explicit time-dependent terms and allows for the construction of a Hamiltonian for the remaining autonomous part [22]. However, ...
  20. [20]
  21. [21]
    On the Completeness of Hamiltonian Vector Fields - jstor
    compact manifold is complete, this being a generalization of the fact that every compact riemannian manifold is complete in the riemannian sense, (the case ...