Skip to main content

Dynamics of Hamiltonian Systems

  • Living reference work entry
  • First Online:

Definition

Hamiltonian systems are a class of dynamical systems which can be characterized by preservation of a symplectic form. This allows to write down the equations of motion in terms of a single function, the Hamiltonian function. They were conceived in the nineteenth century to study physical systems varying from optics to frictionless mechanics in a unified way. This description turned out to be particularly efficient for symmetry reduction and perturbation analysis.

Introduction

The best-known Hamiltonian system is the harmonic oscillator. The second-order differential equation

$$ \ddot{x}+{\varpi}^2x=0 $$

models the motion of a point mass attached to a massless spring (Hooke’s law) and has the general solution

$$ x(t)={x}_0 \cos \varpi t+\frac{y_0}{\varpi } \sin \varpi t $$

with initial conditions \( \left(x(0),\dot{x}(0)\right)=\left({x}_0,{y}_0\right) \). Choosing co-ordinates

$$ \begin{array}{lll}q=\sqrt{\varpi x}\hfill & \mathrm{and}\hfill & p=\frac{y}{\sqrt{\varpi...

This is a preview of subscription content, log in via an institution.

Abbreviations

Action angle variables:

In an integrable Hamiltonian system with n degrees of freedom, the level sets of regular values of the n integrals are Lagrangean submanifolds. In case these level sets are compact, they are (unions of) n-tori. In the neighborhood of such an invariant torus, one can find a Darboux chart with co-ordinates (x, y) such that the integrals (and in particular the Hamiltonian) depend only on the actions y. The values of (x, y) form a product set \( {\mathbb{T}}^n\times \mathbb{Y} \), with \( \mathbb{Y}\subseteq {\mathrm{\mathbb{R}}}^n \) open. Every invariant n-torus y = const in the domain of this chart is parametrized by the angles x.

Canonical 1-form:

The symplectic form (or canonical 2-form) is closed, satisfying dω = 0. If it is furthermore exact, i.e., of the form \( \omega =\mathrm{d}\theta \), then \( \vartheta =-\theta \) is called a canonical 1-form (unique up to addition of df, \( f\in {C}^{\infty}\left(\mathcal{P}\right) \)).

Canonical transformation:

An invertible transformation that preserves the Poisson bracket, turning canonical co-ordinates into canonical co-ordinates. In the symplectic case, it is also called symplectomorphism.

Cantor set, Cantor dust, Cantor family, and Cantor stratification:

Cantor dust is a separable locally compact space that is perfect, i.e., every point is in the closure of its complement, and totally disconnected. This determines Cantor dust up to homeomorphy. The term Cantor set (originally reserved for the specific form of Cantor dust obtained by repeatedly deleting “the middle third”) designates topological spaces that locally have the structure \( {\mathrm{\mathbb{R}}}^n\times Cantor\kern0.5em dust \) for some \( n\in \mathrm{\mathbb{N}} \). Cantor families are parametrized by such Cantor sets. On the real line ℝ, one can define Cantor dust of positive measure by excluding around each rational number p/q an interval of size 2γ/q τ, \( \gamma >0 \), and \( \tau >2 \). Similar Diophantine conditions define Cantor sets in ℝn. Since these Cantor sets have positive measure, their (Hausdorff)-dimension is n. Where the unperturbed system is stratified according to the codimension of occurring (bifurcating) tori, this leads to a Cantor stratification.

Casimir function:

A function \( f:\mathcal{P}\to \mathrm{\mathbb{R}} \) is a Casimir element of the Poisson algebra \( {C}^{\infty}\left(\mathcal{P}\right) \) if {f, g} = 0 for all \( g\in {C}^{\infty}\left(\mathcal{P}\right) \). This induces a Poisson structure on the level sets \( {f}^{-1}(a) \) and \( a\in \mathrm{\mathbb{R}} \). In case every point has a small neighborhood on which the only Casimir elements are the constant functions, the Poisson structure on \( \mathcal{P} \) is nondegenerate, i.e., \( \mathcal{P} \) is a symplectic manifold.

Center:

An equilibrium of a vector field is called a center if all eigenvalues of the linearization are nonzero and on the imaginary axis.

Center-saddle bifurcation:

Under variation of a parameter λ, a center and a saddle meet and vanish.

Chern class:

Let \( \rho :\mathrm{\mathcal{R}}\to C \) be a torus bundle with fiber \( {\mathbb{T}}^n \) and denote by \( \mathcal{G} \) the locally constant sheaf of first homotopy groups \( {\pi}_1\left({\mathbb{T}}^n\right) \) of the fibers. The Chern class of the torus bundle is an element of \( {H}^2\left(C,\mathcal{G}\right) \) that measures the obstruction to the existence of a global section of ρ – such a section exists if and only if the Chern class vanishes. An example with a nonvanishing Chern class is the Hopf fibration \( {S}^3\to {S}^2 \).

Conditionally periodic:

A motion \( t\mapsto \alpha (t)\in \mathcal{P} \) is conditionally periodic if there are frequencies \( {\varpi}_1,\dots, {\varpi}_k\in \mathrm{\mathbb{R}} \) and a smooth embedding \( F:{\mathbb{T}}^k\to \mathcal{P} \) such that \( \alpha (t)=F\left({\mathrm{e}}^{2\pi \mathrm{i}{\varpi}_1t},\dots, {e}^{2\pi \mathrm{i}{\varpi}_kt}\right) \). One can think of the motion as a superposition of the periodic motions \( t\mapsto F\left(1,\dots, 1,{\mathrm{e}}^{2\pi \mathrm{i}{\varpi}_jt},1,\dots, 1\right) \). If the frequencies are rationally independent, the motion \( t\mapsto \alpha (t)\in \mathcal{P} \) lies dense in imF, and this embedded torus is invariant. In case there are resonances among the frequencies, the motion is restricted to a subtorus.

A flow on a torus is parallel or conditionally periodic if there exist co-ordinates in which the vector field becomes constant. In the absence of resonances, the flow is called quasi-periodic.

Conjugate co-ordinates and canonical co-ordinates:

Two co-ordinates Q and P of a Darboux chart are (canonically) conjugate if \( \omega \left({X}_Q,{X}_P\right)=\pm 1 \), i.e., X Q and X P span for every point x in the domain U of the chart a hyperbolic plane in T x U.

Darboux chart, Darboux co-ordinates, and Darboux basis:

In a Darboux basis {e 1,…, e n , f 1,…, f n } of a symplectic vector space (V, ω), the symplectic product takes the simple form \( \omega \left({e}_i,{e}_j\right)=0=\omega \left({f}_i,{f}_j\right) \) and \( \omega \left({e}_i,{f}_j\right)={\delta}_{ij} \). In Darboux co-ordinates (q 1, …, q n , p 1, …, p n ) of a symplectic manifold \( \left(\mathcal{P},\omega \right) \), the symplectic form becomes \( \omega ={\displaystyle \sum \mathrm{d}{q}_i\wedge \mathrm{d}{p}_i} \).

Degree of freedom:

In simple mechanical systems, the phase space is the cotangent bundle of the configuration space, and the dimension of the latter encodes “in how many directions the system can move.” For symplectic manifolds, this notion is immediately generalized to one half of the dimension of the phase space. Poisson spaces are foliated by their symplectic leaves, and the number of degrees of freedom is defined to be one half of the rank of the Poisson structure.

Diophantine condition and Diophantine frequency vector:

A frequency vector \( \varpi \in {\mathrm{\mathbb{R}}}^n \) is called Diophantine if there are constants γ > 0 and τ > n − 1 with \( \begin{array}{ll}\left|\left\langle k\Big|\varpi \right\rangle \right|\ge \frac{\gamma }{\left|k\right|{}^{\tau }}\hfill & \mathrm{f}\mathrm{o}\mathrm{r}\ \mathrm{all}\kern0.5em k\in {\mathbb{Z}}^{n\backslash \left\{0\right\}}\hfill \end{array}. \)

The Diophantine frequency vectors satisfying this condition for fixed γ and τ form a Cantor set of half lines. As the “Diophantine parameter” γ tends to zero (while τ remains fixed), these half lines extend to the origin. The complement in any compact set of frequency vectors satisfying a Diophantine condition with fixed τ has a measure of order O(γ) as γ → 0.

Elliptic:

A periodic orbit/invariant torus is elliptic if all Floquet multipliers/exponents are on the unit circle/imaginary axis. An elliptic equilibrium is called a center.

Energy shell:

The set of points that can be attained given a certain energy. If the phase space is a symplectic manifold, this set is given by the pre-image of that energy value under the Hamiltonian. For more general Poisson spaces, this pre-image has to be intersected with a symplectic leaf.

Focus-focus type:

The (normal) linear behavior generated by a quartet of eigenvalues/Floquet multipliers/exponents.

Frequency-halving bifurcation:

In the supercritical case, an invariant torus loses its stability, giving rise to a stable torus with one of its frequencies halved. In the dual or subcritical case, stability is lost through collision with an unstable invariant torus with one of its frequencies halved. The periodic counterpart is called a period-doubling bifurcation.

Generating function:

A function S = S(q, y) that defines a canonical transformation \( \left(x,y\right)\mapsto \left(q,p\right) \) by means of \( x=\left(\partial S\right)/\left(\partial y\right) \) and \( p=\left(\partial S\right)/\left(\partial q\right) \). The generating function always depends on both the old and the new variables, working with dependence on, e.g., x instead of y introduces a minus sign.

Group action:

A mapping \( \Gamma :G\times \mathcal{P}\to \mathcal{P} \) is a group action of the Lie group G on the phase space \( \mathcal{P} \) if \( {\Gamma}_e=\mathrm{id} \) and \( {\Gamma}_{g\cdot h}={\Gamma}_g\circ {\Gamma}_h \) for all Lie group elements g and h. For \( G=\mathrm{\mathbb{R}} \), one can recover the generating vector field by means of \( X=\frac{\mathrm{d}}{\mathrm{d}t}{\Gamma}_t \).

Hamiltonian Hopf bifurcation:

In a two-degree-of-freedom system, a center loses its stability: the two pairs of purely imaginary eigenvalues meet in a Krein collision. In a three-degree-of-freedom system, the same can happen to two pairs \( {\mathrm{e}}^{\pm \lambda } \) and \( {\mathrm{e}}^{\pm \mu } \) of Floquet multipliers in a one-parameter family of periodic orbits – parametrized by the energy. For a quasi-periodic Hamiltonian Hopf bifurcation, one needs at least four degrees of freedom.

Hamiltonian system:

Newton’s second law states \( F=m\ddot{q} \). Suppose that F is a conservative force, with potential V. Write the equations of motion as a system of first-order differential equations

$$ \dot{q}=\frac{1}{m}p $$
$$ \dot{p}=-\frac{\partial V}{\partial q} $$

that has the total energy \( H\left(q,p\right)=\left\langle p\Big|p\right\rangle /(2m)+V(q) \) as a first integral. This can be generalized. Given a Hamiltonian function H(q, p), one has the Hamiltonian vector field

$$ \begin{array}{c}\hfill \dot{q}=\frac{\partial H}{\partial p}\hfill \\ {}\hfill \dot{p}=-\frac{\partial H}{\partial q}\hfill \end{array} $$

with first integral H. Moreover, one can replace ℝ2n by a symplectic manifold or by a Poisson space.

Hilbert basis:

The (smooth) invariants of an action of a compact group are all given as functions of finitely many “basic” invariants.

Hypoelliptic:

An equilibrium is called hypoelliptic if its linearization has both elliptic and hyperbolic eigenvalues, all nonzero. Similar for periodic orbits and invariant tori.

Integrable system:

A Hamiltonian system with n degrees of freedom is (Liouville)-integrable if it has n functionally independent commuting integrals of motion. Locally, this implies the existence of a (local) torus action.

Integral (of motion):

A conserved quantity.

Iso-energetic Poincaré mapping:

For a Hamiltonian system, a Poincaré mapping leaves the energy shells invariant. Restricting to the intersection \( \varSigma \cap \left\{H=h\right\} \) of the Poincaré section with an energy shell, one obtains an iso-energetic Poincaré mapping.

Isotropic:

For a subspace U < V of a symplectic vector space (V, ω) are equivalent: (i) U is contained in its ω-orthogonal complement, (ii) every basis of U can be extended to a Darboux basis of V. If U satisfies one and thus both of these conditions, it is called an isotropic subspace. A submanifold \( \mathcal{B}\subseteq \mathcal{P} \) of a symplectic manifold \( \left(\mathcal{P},\omega \right) \) is called isotropic if all tangent spaces \( {T}_x\mathcal{B}<{T}_x\mathcal{P} \) are isotropic subspaces.

Krein collision:

Two pairs of purely imaginary Floquet exponents meet in a double pair on the imaginary axis and split off to form a complex quartet \( \pm \mathrm{\Re}\pm \mathrm{i}\Im . \) This is a consequence of a transversality condition on the linear terms at 1:–1 resonance; an additional nondegeneracy condition on the nonlinear part ensures that a ((quasi)-periodic) Hamiltonian Hopf bifurcation takes place.

Lagrangean submanifold:

For a subspace U < V of a symplectic vector space (V, ω) are equivalent: (i) U is a maximal isotropic subspace, (ii) V has a Darboux basis \( \left\{{e}_1,\dots, {e}_n,{f}_1,\dots, {f}_n\right\} \) with span \( \left\{{e}_1,\dots, {e}_n\right\}=U \). If U satisfies one and thus both of these conditions, it is called a Lagrangean subspace. A submanifold \( \mathcal{B}\subseteq \mathcal{P} \) of a symplectic manifold \( \left(\mathcal{P},\omega \right) \) is called Lagrangean if all tangent spaces \( {T}_x\mathcal{B}<{T}_x\mathcal{P} \) are Lagrangean subspaces.

Lie–Poisson structure:

Let \( \mathfrak{g} \) be a Lie algebra with structure constants Γ k ij . Then \( \left\{{\mu}_i,{\mu}_j\right\}=\pm {\displaystyle \sum {\Gamma}_{ij}^k{\mu}_k} \) defines two Poisson structures on the dual space \( \mathfrak{g}* \).

Local bifurcation:

Bifurcations of equilibria can be studied within a small neighborhood in the product of phase space and parameter space. The same is true for fixed points of a discrete dynamical system, but when suspended to a flow, the corresponding bifurcating periodic orbits obtain a semi-local character.

Momentum mapping:

Let \( \Gamma :G\times \mathcal{P}\to \mathcal{P} \) be a symplectic action of the Lie group G on the symplectic manifold \( \mathcal{P} \). A mapping \( J:\mathcal{P}\to \mathfrak{g}* \) into the dual space of the Lie algebra of G is a momentum mapping for the action if \( {X}_{J^{\xi }}={\xi}_{\mathcal{P}} \) for all \( \xi \in \mathfrak{g} \). Here, \( {J}^{\xi }:\mathcal{P}\to \mathrm{\mathbb{R}} \) is defined by \( {J}^{\xi }(z)=J(z)\cdot \xi \) and \( {\xi}_{\mathcal{P}} \) is the infinitesimal generator of the action corresponding to ξ. The momentum mapping J is called Ad*-equivariant provided that \( J\circ {\Gamma}_g={\mathrm{Ad}}_{g^{-1}}^{*}\circ J \).

Monodromy:

The group homomorphism \( \mathrm{\mathcal{M}}:{\pi}_1(C)\to S{L}_n\left(\mathrm{\mathbb{Z}}\right) \) that measures how much the n-torus bundle \( \mathrm{\mathcal{R}}\to C \) deviates from a principal torus bundle.

Nondegenerate integrable Hamiltonian system, function:

In action angle variables (x, y), the Hamiltonian H only depends on the action variables y, and the equations of motion become \( \dot{x}=\varpi (y) \) and \( \dot{y}=0 \), with frequencies \( \varpi (y)=\left(\partial H\right)/\left(\partial y\right) \). The system is nondegenerate at the invariant torus {y = y 0} if D 2 H(y 0) is invertible. In this case, ϖ defines near {y = y 0} an isomorphism between the actions and the angular velocities. Other conditions ensuring that most tori have Diophantine frequency vectors are iso-energetic nondegeneracy or Rüssmann-like conditions on higher derivatives.

Normal frequency:

Given an elliptic invariant torus of a Hamiltonian system, one can define the normal linearization on the symplectic normal bundle. The eigenvalues of the normal linearization being \( \pm {\mathrm{i}\Omega}_1,\dots, \pm {\mathrm{i}\Omega}_m \), one calls the Ω j the normal frequencies. Under the exponential mapping, the eigenvalues of the normal linearization of a periodic orbit are mapped to Floquet multipliers.

Normal linearization:

The linearization within the normal bundle of an invariant submanifold. In the Hamiltonian context, these are often isotropic, and one further restricts to the symplectic normal linear behavior, e.g., to identify elliptic and hyperbolic invariant tori.

Normal modes:

A family of periodic orbits parametrized by the energy value h, as h → h 0 shrinking down to an equilibrium that has a pair of eigenvalues \( \pm 2\pi \mathrm{i}\underset{h\to {h}_0}{ \lim }{T}_h^{-1} \), where T h denotes the period(s).

Parabolic:

An equilibrium of a one-degree-of-freedom system is called parabolic if its linearization is nilpotent but nonzero. An invariant torus is parabolic if its symplectic normal linearization has a parabolic equilibrium. In particular, the four Floquet multipliers of a parabolic periodic orbit in two degrees of freedom are all equal to 1.

Phase space:

By Newton’s second law, the equations of motion are second-order differential equations. The trajectory is completely determined by the initial positions and the initial velocities, or, equivalently, the initial momenta. The phase space is the set of all possible combinations of initial positions and initial momenta.

Pinched torus:

The compact (un)stable manifold of a saddle in two degrees of freedom with a quartet \( \pm \Re \pm \mathrm{i}\Im \) of hyperbolic eigenvalues resembles a torus \( {\mathbb{T}}^2={S}^1\times {S}^1 \) with one of the fibers \( \left\{x\right\}\times {S}^1 \) reduced to a point.

Poisson space, Poisson bracket, Poisson structure:

A Poisson algebra \( \mathcal{A} \) is a real Lie algebra that is also a commutative ring with unit. These two structures are related by Leibniz’ rule \( \left\{f\cdot g,h\right\}=f\cdot \left\{g,h\right\}+g\cdot \left\{f,h\right\} \). A Poisson manifold \( \mathcal{P} \) has a Poisson bracket on \( {C}^{\infty}\left(\mathcal{P}\right) \) that makes \( {C}^{\infty}\left(\mathcal{P}\right) \) a Poisson algebra. If there are locally no Casimir elements other than constant functions, this leads to a symplectic structure on \( \mathcal{P} \).

Poisson spaces naturally arise in singular reduction; this motivates to allow varieties \( \mathcal{P} \) where the Poisson bracket is defined on a suitable subalgebra \( \mathcal{A} \) of \( C\left(\mathcal{P}\right) \).

Given a Hamiltonian function \( H\in \mathcal{A} \), one obtains for \( f\in \mathcal{A} \) the equations of motion \( \frac{\mathrm{d}}{\mathrm{d}t}f=\left\{f,H\right\} \). For canonically conjugate co-ordinates (q, p) on \( \mathcal{P} \), i.e., with \( \left\{{q}_i,{q}_j\right\}=0=\left\{{p}_i,{p}_j\right\} \) and \( \left\{{q}_i,{p}_j\right\}={\delta}_{ij} \), this amounts to

$$ \begin{array}{c}\hfill \dot{q}=\frac{\partial H}{\partial p}\hfill \\ {}\hfill \dot{p}=-\frac{\partial H}{\partial q}.\hfill \end{array} $$
Poisson symmetry:

A symmetry that preserves the Poisson structure.

Principal fiber bundle:

A fiber bundle \( F:\mathcal{P}\to C \) admitting a free Lie group action \( G\times \mathcal{P}\to \mathcal{P} \) that acts transitively on the fibers \( {F}^{-1}(c)\cong G \).

Proper degeneracy:

For the application of the KAM theorem to a perturbation of an integrable system, it is necessary that the integrable system is nondegenerate, so that the frequencies have maximal rank as function of the actions. If there are global conditions relating the frequencies, so that the conditionally periodic motion can be described by a smaller number of these, the system is properly degenerate. Superintegrable systems are a particular example.

Quasi-periodic:

A conditionally periodic motion that is not periodic. The closure of the trajectory is an invariant k-torus with k ≥ 2.

A parallel or conditionally periodic flow on a k-torus is called quasi-periodic if the frequencies ϖ 1, …, ϖ k are rationally independent.

Ramified torus bundle:

Let a differentiable mapping \( f:\mathcal{P}\to {\mathrm{\mathbb{R}}}^m \) be given. According to Sard’s Lemma, almost all values \( a\in {\mathrm{\mathbb{R}}}^m \) are regular. The connected components of the sets \( {f}^{-1}(a) \) and \( a\in \mathrm{i}\mathrm{m}\;f \) regular, define a foliation of an open subset ℛ of the (n + m)-dimensional manifold \( \mathcal{P} \).

In the present settings, the components of f are the first integrals of an integrable Hamiltonian system with compact level sets. Then the regular fibers are n-tori, and their union is a torus bundle. The topology of this bundle depends on the topology of the base space C, the monodromy, and the Chern class of the bundle. In many examples that one encounters, the connected components of C are contractible, whence monodromy and Chern classes are trivial.

For the geometry of the bundle, one also wants to know how the singular fibers are distributed: where n-tori shrink to normally elliptic (n − 1)-tori and where they are separated by stable and unstable manifolds of normally hyperbolic (n − 1)-tori. The singular fibers are not necessarily manifolds but may be stratified into X H -invariant strata which are possibly noncompact.

One can continue and look for the singularities of these “regular singular leaves,” the tori of dimension ≤n − 2 and the normally parabolic (n − 1)-tori, in which normally elliptic and normally hyperbolic (n − 1)-tori meet in a quasi-periodic center-saddle bifurcation or a frequency-halving bifurcation. The next “layer” is given by quasi-periodic Hamiltonian Hopf bifurcations and bifurcations of higher codimension.

Reduced phase space:

Let \( \Gamma :G\times \mathcal{P}\to \mathcal{P} \) be a symplectic action of the Lie group G on the symplectic manifold \( \mathcal{P} \) with Ad*-equivariant momentum mapping \( J:\mathcal{P}\to \mathfrak{g}* \). For a regular value \( \mu \in \mathfrak{g}* \), let the action of the isotropy group \( {G}_{\mu }=\left\{g\in G\Big|{\mathrm{Ad}}_{g^{-1}}^{*}\left(\mu \right)=\mu \right\} \) on \( {J}^{-1}\left(\mu \right) \) be free and proper. Then the quotient \( {J}^{-1}\left(\mu \right)/{G}_{\mu } \) is again a symplectic manifold, the reduced phase space. A Γ-invariant Hamiltonian function H on \( \mathcal{P} \) leads to a reduced Hamiltonian function H μ on the reduced phase space.

Relative equilibrium:

Let H be a Hamiltonian function invariant under the symplectic action \( G\times \mathcal{P}\to \mathcal{P} \), and let \( \mu \in \mathfrak{g}* \) be a regular value of the Ad*-equivariant momentum mapping \( J:\mathcal{P}\to \mathfrak{g}* \). Also assume that the isotropy group G μ under the Ad* action on \( \mathfrak{g}* \) acts freely and properly on \( {J}^{-1}\left(\mu \right) \). Then X H induces a Hamiltonian flow on the reduced phase space \( {\mathcal{P}}_{\mu }={J}^{-1}\left(\mu \right)/{G}_{\mu } \). The phase curves of the given Hamiltonian system on \( \mathcal{P} \) with momentum constant \( J=\mu \) that are taken by the projection \( {J}^{-1}\left(\mu \right)\to {\mathcal{P}}_{\mu } \) into equilibrium positions of the reduced Hamiltonian system are called relative equilibria or stationary motions (of the original system).

Remove the degeneracy:

A perturbation of a superintegrable system removes the degeneracy if it is sufficiently mild to define an intermediate system that is still integrable and sufficiently wild to make that intermediate system nondegenerate.

Resonance:

If the frequencies of an invariant torus with conditionally periodic flow are rationally dependent, this torus divides into invariant subtori. Such resonances \( \left\langle h\Big|\varpi \right\rangle =0 \) and \( h\in {\mathrm{\mathbb{Z}}}^k \) define hyperplanes in ϖ space and, by means of the frequency mapping, also in phase space. The smallest number \( \left|h\right|=\left|{h}_1\right|+\cdots +\left|{h}_k\right| \) is the order of the resonance. Diophantine conditions describe a measure-theoretically large complement of a neighborhood of the (dense!) set of all resonances.

Saddle:

An equilibrium of a vector field is called a saddle if the linearization has no eigenvalues on the imaginary axis. On a small neighborhood of a saddle, the flow is topologically conjugate to its linearization.

Semi-local bifurcation:

Bifurcations of n-tori can be studied in a tubular neighborhood. For n = 1, a Poincaré section turns the periodic orbit into a fixed point of the Poincaré mapping, and the bifurcation obtains a local character.

Simple mechanical system:

A quintuple \( \left({T}^{*}M,\left\langle ..\Big|..\right\rangle, V,T,\omega \right) \) consisting of the cotangent bundle of a Riemannian manifold \( \left(M,\left\langle ..\Big|..\right\rangle \right) \), a potential function \( V:M\to \mathrm{\mathbb{R}} \), the kinetic energy \( T\left(\alpha \right)=\left\langle \alpha \Big|\alpha \right\rangle \), and the symplectic form \( \omega =-\mathrm{d}\vartheta \) derived from the canonical 1-form on the cotangent bundle T*M.

Singular reduction:

If \( \Gamma :G\times \mathcal{P}\to \mathcal{P} \) is a Poisson symmetry, then the group action on \( \mathcal{P} \) makes \( \mathcal{B}=\mathcal{P}/G \) a Poisson space as well. Fixing the values of the resulting Casimir functions yields the reduced phase space, which turns out to have singular points where the action Γ is not free.

Solenoid:

Given a sequence \( {f}_j:{S}^1\to {S}^1 \) of coverings \( {f}_j\left(\zeta \right)={\zeta}^{\alpha_j} \) of the circle S 1, the solenoid \( {\varSigma}_a\subseteq {\left({S}^1\right)}^{{\mathrm{\mathbb{N}}}_0} \), \( a={\left({\alpha}_j\right)}_{j\in {\mathrm{\mathbb{N}}}_0} \) consists of all \( z={\left({\zeta}_j\right)}_{j\in {\mathrm{\mathbb{N}}}_0} \) with \( {\zeta}_j={f}_j\left({\zeta}_{j+1}\right) \) for all \( j\in {\mathrm{\mathbb{N}}}_0 \).

Stratification:

The decomposition of a topological space into smaller pieces satisfying certain boundary conditions.

Superintegrable system:

A Hamiltonian system with n degrees of freedom is superintegrable if it has n + 1 functionally independent integrals of motion such that each of the first n − 1 of them commutes with all n + 1. Such a properly degenerate system admits generalized action angle co-ordinates \( \left({x}_1,\dots, {x}_{n-1},{y}_1,\dots, {y}_{n-1},q,p\right) \). In case the nondegeneracy condition det \( {D}^2H(y)\ne 0 \) is satisfied almost everywhere, the system is “minimally superintegrable.” In the other extreme of a “maximally superintegrable” system, all motions are periodic.

Symplectic manifold:

A 2n-dimensional manifold \( \mathcal{P} \) with a nondegenerate closed two-form ω, i.e., dω = 0 and ω(u, v) = 0 for all \( v\in T\mathcal{P}\Rightarrow u=0 \). A diffeomorphism ψ of symplectic manifolds that respects the two-form(s) is called a symplectomorphism. Given a Hamiltonian function \( H\in {C}^{\infty}\left(\mathcal{P}\right) \), one obtains through \( \omega \left({X}_H,\dots \right)=\mathrm{d}H \) the Hamiltonian vector field \( {X}_H \). For every \( x\in \mathcal{P} \), there are co-ordinates (q, p) around x with \( \omega =\mathrm{d}q\wedge \mathrm{d}p \). In these Darboux co-ordinates \( {X}_H \) reads

$$ \begin{array}{c}\hfill \dot{q}=\frac{\partial H}{\partial p}\hfill \\ {}\hfill \dot{p}=-\frac{\partial H}{\partial q}.\hfill \end{array} $$
Symplectic form:

A nondegenerate closed two-form.

Syzygy:

A constraining equation that is identically fulfilled by the elements of a Hilbert basis.

(Un)stable manifold:

In Hamiltonian systems with one degree of freedom, the stable manifold and the unstable manifold of an equilibrium often coincide and thus consist of homoclinic orbits. In such a case, it is called an (un)stable manifold. This carries over to the stable and the unstable manifold of a periodic orbit or an invariant torus in higher degrees of freedom if the system is integrable.

Bibliography

Primary Literature

  • Abraham R, Marsden JE (1978) Foundations of mechanics, 2nd edn. Benjamin

    Google Scholar 

  • Arnol’d VI (1989) Mathematical methods of classical mechanics, 2nd edn. Springer

    Book  MATH  Google Scholar 

  • Arnol’d VI, Avez A (1968) Ergodic problems of classical mechanics. Benjamin

    MATH  Google Scholar 

  • Birkhoff BD (1935) Nouvelles recherches sur les systèmes dynamiques. Mem Pont Acad Sci Novi Lyncaei, Ser3 1:85–216

    Google Scholar 

  • Broer HW, Cushman R, Fassò F, Takens F (2007) Geometry of KAM tori for nearly integrable Hamiltonian systems. Ergod Theory Dyn Syst 27(3):725–741

    Article  MATH  Google Scholar 

  • Broer HW, Dumortier F, van Strien SJ, Takens F (1991) Structures in dynamics. Finite-dimensional deterministic studies. Studies in mathematical physics, vol 2. North-Holland

    MATH  Google Scholar 

  • Broer HW, Huitema GB, Sevryuk MB (1996) Quasi-periodic motions in families of dynamical systems: order amidst chaos. Lecture notes in mathematics, vol 1645. Springer

    Google Scholar 

  • Cushman RH, Bates LM (1997) Global aspects of classical integrable systems. Birkhäuser

    Book  MATH  Google Scholar 

  • Delshams A, de la Llave R, Martínez-Seara T (2006) A geometric mechanism for diffusion in Hamiltonian systems overcoming the large gap problem: heuristics and rigorous verification on a model. Mem Am Math Soc 179(844):1–141

    Google Scholar 

  • Duistermaat JJ (1980) On global action-angle co-ordinates. Commun Pure Appl Math 33:687–706

    Article  MATH  MathSciNet  Google Scholar 

  • Fassò F (2005) Superintegrable Hamiltonian systems: geometry and perturbation. In: Gaeta G (ed) Symmetry and perturbation theory, Cala Gonone. Acta applicandae mathematicae 87:93–121

    Google Scholar 

  • Goldstein H (1980) Classical mechanics. Addison-Wesley

    MATH  Google Scholar 

  • Hanßmann H (2007) Local and semi-local bifurcations in Hamiltonian dynamical systems – results and examples. Lecture notes in mathematics, vol 1893. Springer

    MATH  Google Scholar 

  • Kappeler T, Pöschel J (2003) KdV & KAM, vol 45. Ergebnisse der Mathematik und ihrer Grenzgebiete, 3. Springer

    Book  MATH  Google Scholar 

  • Kuksin SB (1993) Nearly integrable infinite-dimensional Hamiltonian systems. Lecture notes in mathematics, vol 1556. Springer

    MATH  Google Scholar 

  • Landau LD, Lifshitz EM (1960) Mechanics. Course of theoretical physics, vol 1. Pergamon

    MATH  Google Scholar 

  • Litvak-Hinenzon A, Rom-Kedar V (2002) Parabolic resonances in 3 degree of freedom near-integrable Hamiltonian systems. Physica D 164:213–250

    Article  MATH  MathSciNet  ADS  Google Scholar 

  • Lukina O, Takens F, Broer HW (2008) Global properties of integrable Hamiltonian systems. Regul Chaotic Dyn 13(6):602–644

    Article  MATH  MathSciNet  ADS  Google Scholar 

  • MacKay RS (1993) Renormalisation in area-preserving maps. World Scientific

    MATH  Google Scholar 

  • Markus L, Meyer KR (1980) Periodic orbits and solenoids in generic Hamiltonian dynamical systems. Am J Math 102(1):25–92

    Article  MATH  MathSciNet  Google Scholar 

  • Marsden JE, Raţiu TS (1994) Introduction to mechanics and symmetry. Springer

    Book  MATH  Google Scholar 

  • Meyer KR, Hall GR (1992) Introduction to Hamiltonian dynamical systems and the N-body problem. Applied mathematical sciences, vol 90. Springer

    MATH  Google Scholar 

  • Montaldi J, Raţiu TS (2005) Geometric mechanics and symmetry: the Peyresq lectures. Cambridge University Press

    Book  Google Scholar 

  • Moser J (1973) Stable and random motions in dynamical systems: with special emphasis on celestial mechanics. Princeton University Press

    MATH  Google Scholar 

  • Moser J, Zehnder EJ (2005) Notes on dynamical systems. Courant lecture notes in mathematics, vol 12. AMS

    Google Scholar 

  • Ortega J-P, Raţiu TS (2004) Momentum maps and Hamiltonian reduction. Progress in mathematics, vol 222. Birkhäuser, Boston

    Book  MATH  Google Scholar 

  • Ratiu TS, Tudoran R, Sbano L, Sousa Dias E, Terra G (2005) A crash course in geometric mechanics. In: Montaldi J, Ratiu TS (eds) Geometric mechanics and symmetry: the Peyresq lectures. Cambridge University Press pp 23–156

    Chapter  Google Scholar 

  • Sanders JA, Verhulst F, Murdock J (2007) Averaging methods in nonlinear dynamical systems, 2nd edn. Springer

    MATH  Google Scholar 

  • Sevryuk MB (2003) The classical KAM theory at the dawn of the twenty-first century. Mosc Math J 3(3):1113–1144

    MATH  MathSciNet  Google Scholar 

  • Siegel CL, Moser JK (1971) Lectures on celestial mechanics. Springer

    Book  MATH  Google Scholar 

Books and Reviews

  • Arnol’d VI, Givental’ AB (2001) Symplectic geometry. In: Novikov SP (ed) Dynamical systems IV, symplectic geometry and its applications. Springer pp 1–138

    Google Scholar 

  • Arnol’d VI, Kozlov VV, Neishtadt AI (1988) Mathematical aspects of classical and celestial mechanics. In: Arnol’d VI (ed) Dynamical systems III. Springer

    Chapter  Google Scholar 

  • Arnol’d VI, Novikov SP (1994) Dynamical systems VII, integrable systems. Springer

    Book  MATH  Google Scholar 

  • Bartsch T, Szulkin A (2005) Hamiltonian systems: periodic and homoclinic solutions by variational methods. In: Cañada A, Drábek P, Fonda A (eds) Handbook of differential equations: ordinary differential equations, vol II. Elsevier pp 77–146

    Google Scholar 

  • Bibikov YN (1979) Local theory of nonlinear analytic ordinary differential equations. Lecture notes in mathematics, vol 702. Springer

    MATH  Google Scholar 

  • Birkhoff BD (1927) Dynamical systems. AMS, Reprint: AMS (1966)

    MATH  Google Scholar 

  • Bolsinov AV, Fomenko AT (2004) Integrable Hamiltonian systems – geometry, topology, classification. Chapman & Hall/CRC

    MATH  Google Scholar 

  • Bridges TJ, Furter JE (1993) Singularity theory and equivariant symplectic maps. Lecture notes in mathematics, vol 1558. Springer

    MATH  Google Scholar 

  • Broer HW, Hasselblatt B, Takens F (2010) Handbook of dynamical systems, vol 3. North-Holland

    MATH  Google Scholar 

  • Broer HW, Hoveijn I, Lunter G, Vegter G (2003) Bifurcations in Hamiltonian systems: computing singularities by Gröbner bases. Lecture notes in mathematics, vol 1806. Springer

    Book  Google Scholar 

  • Broer HW, Sevryuk MB (2010) KAM theory: quasi-periodicity in dynamical systems. In: Broer HW, Hasselblatt B, Takens F (eds) Handbook of dynamical systems, vol 3. North-Holland, pp 249–344

    Google Scholar 

  • Buono P-L, Laurent-Polz F, Montaldi J (2005) Symmetric Hamiltonian bifurcations. In: Montaldi J, Ratiu TS (eds) Geometric mechanics and symmetry: the Peyresq lectures. Cambridge University Press, Cambridge, pp 357–402

    Chapter  Google Scholar 

  • Cannas da Silva A, Weinstein A (1999) Geometric models for noncommutative algebras. Berkeley mathematics lecture notes, vol 10. AMS

    MATH  Google Scholar 

  • Ciocci MC, Litvak-Hinenzon A, Broer HW (2005) Survey on dissipative KAM theory including quasi-periodic bifurcation theory. In: Montaldi J, Ratiu TS (eds) Geometric mechanics and symmetry: the Peyresq lectures. Cambridge University Press, Cambridge, pp 303–355

    Chapter  Google Scholar 

  • Cushman R (1992) A survey of normalization techniques applied to perturbed Keplerian systems. In: Jones K et al (eds) Dynamics reported, new series 1:54–112

    Google Scholar 

  • Dubrovin BA, Krichever IM, Novikov SP (2001) Integrable systems I. In: Novikov SP (ed) Dynamical systems IV, symplectic geometry and its applications. Springer, Berlin, pp 177–332

    Google Scholar 

  • Duistermaat JJ (1984) Bifurcations of periodic solutions near equilibrium points of Hamiltonian systems. In: Salvadori L (ed) Bifurcation theory and applications, Montecatini 1983. Lecture notes in mathematics, vol 1057. Springer, pp 57–105

    Chapter  Google Scholar 

  • Efstathiou K (2005) Metamorphoses of Hamiltonian systems with symmetries. Lecture notes in mathematics, vol 1864. Springer

    MATH  Google Scholar 

  • Efstathiou K, Sadovskií D (2005) No polar co-ordinates. In: Montaldi J, Ratiu TS (eds) Geometric mechanics and symmetry: the Peyresq lectures. Cambridge University Press, Cambridge, pp 211–301

    Chapter  Google Scholar 

  • Eliasson LH, Kuksin SB, Marmi S, Yoccoz J-C (2002) Dynamical systems and small divisors, Cetraro 1998 Lecture notes in mathematics, vol 1784. Springer

    Google Scholar 

  • Gallavotti G (1983) The elements of mechanics. Springer

    Book  MATH  Google Scholar 

  • Gallavotti G (1994) Twistless KAM tori, quasi flat homoclinic intersections, and other cancellations in the perturbation series of certain completely integrable Hamiltonian systems. A review. Rev Math Phys 6(3):343–411

    Article  MATH  MathSciNet  Google Scholar 

  • Guillemin V, Sternberg S (1984) Symplectic techniques in physics. Cambridge University Press

    MATH  Google Scholar 

  • Guillemin V, Sternberg S (1990) Variations on a theme by Kepler. American Mathematical Society Colloquium publications, vol 42. AMS

    MATH  Google Scholar 

  • Haller G (1999) Chaos near resonance. Applied mathematical sciences, vol 138. Springer

    MATH  Google Scholar 

  • Hitchin NJ, Segal GB, Ward RS (1999) Integrable systems. Twistors, loop groups, and Riemann surfaces, Oxford 1997. Clarendon/Oxford University Press

    Google Scholar 

  • Hofer H, Zehnder E (1994) Symplectic invariants and Hamiltonian dynamics. Birkhäuser

    Book  Google Scholar 

  • Holm DD (2005) The Euler-Poincaré variational framework for modeling fluid dynamics. In: Montaldi J, Ratiu TS (eds) Geometric mechanics and symmetry: the Peyresq lectures. Cambridge University Press, Cambridge, pp 157–209

    Chapter  Google Scholar 

  • Karasev MV, Maslov VP (1993) Nonlinear Poisson brackets: geometry and quantization. Translations of mathematical monographs, vol 119. AMS

    MATH  Google Scholar 

  • Kuksin SB (2000) Analysis of Hamiltonian PDEs. Oxford lecture series in mathematics and its applications, vol 19. Oxford University Press

    Google Scholar 

  • Libermann P, Marle C-M (1987) Symplectic geometry and analytical mechanics. Reidel

    Book  MATH  Google Scholar 

  • Markus L, Meyer KR (1974) Generic Hamiltonian dynamical systems are neither integrable nor ergodic, Mem Am Math Soc 144:1–52

    Google Scholar 

  • Marsden JE (1992) Lectures on mechanics. Cambridge University Press

    Book  MATH  Google Scholar 

  • Marsden J, Montgomery R, Raţiu T (1990) Reduction, symmetry, and phases in mechanics, Mem Am Math Soc 88(436):1–110

    Google Scholar 

  • Marsden JE, Misiołek G, Ortega JP, Perlmutter M, Raţiu TS (2007) Hamiltonian reduction by stages. Lecture notes in mathematics, vol 1913. Springer

    Google Scholar 

  • Mawhin J, Willem M (1989) Critical point theory and Hamiltonian systems, 2nd edn. Applied mathematical sciences, vol 74. Springer

    Book  MATH  Google Scholar 

  • McDuff D, Salamon D (1995) Introduction to symplectic topology. Clarendon/Oxford University Press

    MATH  Google Scholar 

  • Meyer KR (1999) Periodic solutions of the N–body problem. Lecture notes in mathematics, vol 1719. Springer

    MATH  Google Scholar 

  • Mielke A (1991) Hamiltonian and Lagrangian flows on center manifolds – with applications to elliptic variational problems. Lecture notes in mathematics, vol 1489. Springer

    Book  MATH  Google Scholar 

  • Novikov SP (2001) Dynamical systems IV, symplectic geometry and its applications. Springer

    MATH  Google Scholar 

  • Olshanetsky MA, Perelomov AM, Reyman AG, Semenov-Tian-Shansky MA (1994) Integrable systems II. In: Arnol’d VI, Novikov SP (eds) Dynamical systems VII, integrable systems. Springer, Berlin, pp 87–259

    Chapter  Google Scholar 

  • Oxtoby JC (1971) Measure and category. A survey of the analogies between topological and measure spaces. Graduate texts in mathematics, vol 2. Springer

    MATH  Google Scholar 

  • Poincaré H (1892/1893/1899) Les méthodes nouvelles de la mécanique céleste. Gauthier–Villars. Reprint: Dover (1957)

    Google Scholar 

  • Rabinowitz PH (2002) Variational methods for Hamiltonian systems. In: Hasselblatt B, Katok A (eds) Handbook of dynamical systems, vol 1A. North-Holland, pp 1091–1127

    Google Scholar 

  • Rink B, Tuwankotta T (2005) Stability in Hamiltonian systems. In: Montaldi J, Ratiu TS (eds) Geometric mechanics and symmetry: the Peyresq lectures. Cambridge University Press, Cambridge, pp 1–22

    Chapter  Google Scholar 

  • Souriau J-M (1970) Structure des systèmes dynamiques. Dunod, Paris. English edition: Souriau J-M (1997) Structure of dynamical systems, a symplectic view of physics. Progress in mathematics, vol 149. Birkhäuser

    Google Scholar 

  • Thirring W (1978) A course in mathematical physics I. Classical dynamical systems. Springer

    Book  Google Scholar 

  • Trofimov VV, Fomenko AT (1994) Geometric and algebraic mechanisms of the integrability of Hamiltonian systems on homogeneous spaces and Lie algebras. In: Arnol’d VI, Novikov SP (eds) Dynamical systems VII, integrable systems. Springer, Berlin, pp 261–333

    Google Scholar 

  • van der Meer JC (1985) The Hamiltonian Hopf bifurcation. Lecture notes in mathematics, vol 1160. Springer

    Book  MATH  Google Scholar 

Download references

Acknowledgments

The author thanks Henk Broer, Marius Crainic, Hans Duistermaat, Luuk Hoevenaars, Hessel Posthuma, and Ferdinand Verhulst for fruitful discussions and helpful remarks.

Author information

Authors and Affiliations

Authors

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer Science+Business Media New York

About this entry

Cite this entry

Hanßmann, H. (2015). Dynamics of Hamiltonian Systems. In: Meyers, R. (eds) Encyclopedia of Complexity and Systems Science. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-27737-5_141-2

Download citation

  • DOI: https://doi.org/10.1007/978-3-642-27737-5_141-2

  • Received:

  • Accepted:

  • Published:

  • Publisher Name: Springer, Berlin, Heidelberg

  • Online ISBN: 978-3-642-27737-5

  • eBook Packages: Springer Reference Physics and AstronomyReference Module Physical and Materials ScienceReference Module Chemistry, Materials and Physics

Publish with us

Policies and ethics