Hamiltonian vector field

In mathematics and physics, a Hamiltonian vector field on a symplectic manifold is a vector field defined for any energy function or Hamiltonian. Named after the physicist and mathematician Sir William Rowan Hamilton, a Hamiltonian vector field is a geometric manifestation of Hamilton's equations in classical mechanics. The integral curves of a Hamiltonian vector field represent solutions to the equations of motion in the Hamiltonian form. The diffeomorphisms of a symplectic manifold arising from the flow of a Hamiltonian vector field are known as canonical transformations in physics and (Hamiltonian) symplectomorphisms in mathematics.[1]

Hamiltonian vector fields can be defined more generally on an arbitrary Poisson manifold. The Lie bracket of two Hamiltonian vector fields corresponding to functions f and g on the manifold is itself a Hamiltonian vector field, with the Hamiltonian given by the Poisson bracket of f and g.

Definition

Suppose that (M, ω) is a symplectic manifold. Since the symplectic form ω is nondegenerate, it sets up a fiberwise-linear isomorphism

ω : T M T M , {\displaystyle \omega :TM\to T^{*}M,}

between the tangent bundle TM and the cotangent bundle T*M, with the inverse

Ω : T M T M , Ω = ω 1 . {\displaystyle \Omega :T^{*}M\to TM,\quad \Omega =\omega ^{-1}.}

Therefore, one-forms on a symplectic manifold M may be identified with vector fields and every differentiable function H: MR determines a unique vector field XH, called the Hamiltonian vector field with the Hamiltonian H, by defining for every vector field Y on M,

d H ( Y ) = ω ( X H , Y ) . {\displaystyle \mathrm {d} H(Y)=\omega (X_{H},Y).}

Note: Some authors define the Hamiltonian vector field with the opposite sign. One has to be mindful of varying conventions in physical and mathematical literature.

Examples

Suppose that M is a 2n-dimensional symplectic manifold. Then locally, one may choose canonical coordinates (q1, ..., qn, p1, ..., pn) on M, in which the symplectic form is expressed as:[2] ω = i d q i d p i , {\displaystyle \omega =\sum _{i}\mathrm {d} q^{i}\wedge \mathrm {d} p_{i},}

where d denotes the exterior derivative and denotes the exterior product. Then the Hamiltonian vector field with Hamiltonian H takes the form:[1] X H = ( H p i , H q i ) = Ω d H , {\displaystyle \mathrm {X} _{H}=\left({\frac {\partial H}{\partial p_{i}}},-{\frac {\partial H}{\partial q^{i}}}\right)=\Omega \,\mathrm {d} H,}

where Ω is a 2n × 2n square matrix

Ω = [ 0 I n I n 0 ] , {\displaystyle \Omega ={\begin{bmatrix}0&I_{n}\\-I_{n}&0\\\end{bmatrix}},}

and

d H = [ H q i H p i ] . {\displaystyle \mathrm {d} H={\begin{bmatrix}{\frac {\partial H}{\partial q^{i}}}\\{\frac {\partial H}{\partial p_{i}}}\end{bmatrix}}.}

The matrix Ω is frequently denoted with J.

Suppose that M = R2n is the 2n-dimensional symplectic vector space with (global) canonical coordinates.

  • If H = p i {\displaystyle H=p_{i}} then X H = / q i ; {\displaystyle X_{H}=\partial /\partial q^{i};}
  • if H = q i {\displaystyle H=q_{i}} then X H = / p i ; {\displaystyle X_{H}=-\partial /\partial p^{i};}
  • if H = 1 / 2 ( p i ) 2 {\displaystyle H=1/2\sum (p_{i})^{2}} then X H = p i / q i ; {\displaystyle X_{H}=\sum p_{i}\partial /\partial q^{i};}
  • if H = 1 / 2 a i j q i q j , a i j = a j i {\displaystyle H=1/2\sum a_{ij}q^{i}q^{j},a_{ij}=a_{ji}} then X H = a i j q i / p j . {\displaystyle X_{H}=-\sum a_{ij}q_{i}\partial /\partial p^{j}.}

Properties

  • The assignment fXf is linear, so that the sum of two Hamiltonian functions transforms into the sum of the corresponding Hamiltonian vector fields.
  • Suppose that (q1, ..., qn, p1, ..., pn) are canonical coordinates on M (see above). Then a curve γ(t) = (q(t),p(t)) is an integral curve of the Hamiltonian vector field XH if and only if it is a solution of Hamilton's equations:[1] q ˙ i = H p i {\displaystyle {\dot {q}}^{i}={\frac {\partial H}{\partial p_{i}}}}
p ˙ i = H q i . {\displaystyle {\dot {p}}_{i}=-{\frac {\partial H}{\partial q^{i}}}.}
  • The Hamiltonian H is constant along the integral curves, because d H , γ ˙ = ω ( X H ( γ ) , X H ( γ ) ) = 0 {\displaystyle \langle dH,{\dot {\gamma }}\rangle =\omega (X_{H}(\gamma ),X_{H}(\gamma ))=0} . That is, H(γ(t)) is actually independent of t. This property corresponds to the conservation of energy in Hamiltonian mechanics.
  • More generally, if two functions F and H have a zero Poisson bracket (cf. below), then F is constant along the integral curves of H, and similarly, H is constant along the integral curves of F. This fact is the abstract mathematical principle behind Noether's theorem.[nb 1]
  • The symplectic form ω is preserved by the Hamiltonian flow. Equivalently, the Lie derivative L X H ω = 0. {\displaystyle {\mathcal {L}}_{X_{H}}\omega =0.}

Poisson bracket

The notion of a Hamiltonian vector field leads to a skew-symmetric bilinear operation on the differentiable functions on a symplectic manifold M, the Poisson bracket, defined by the formula

{ f , g } = ω ( X g , X f ) = d g ( X f ) = L X f g {\displaystyle \{f,g\}=\omega (X_{g},X_{f})=dg(X_{f})={\mathcal {L}}_{X_{f}}g}

where L X {\displaystyle {\mathcal {L}}_{X}} denotes the Lie derivative along a vector field X. Moreover, one can check that the following identity holds:[1] X { f , g } = [ X f , X g ] , {\displaystyle X_{\{f,g\}}=[X_{f},X_{g}],}

where the right hand side represents the Lie bracket of the Hamiltonian vector fields with Hamiltonians f and g. As a consequence (a proof at Poisson bracket), the Poisson bracket satisfies the Jacobi identity:[1] { { f , g } , h } + { { g , h } , f } + { { h , f } , g } = 0 , {\displaystyle \{\{f,g\},h\}+\{\{g,h\},f\}+\{\{h,f\},g\}=0,}

which means that the vector space of differentiable functions on M, endowed with the Poisson bracket, has the structure of a Lie algebra over R, and the assignment fXf is a Lie algebra homomorphism, whose kernel consists of the locally constant functions (constant functions if M is connected).

Remarks

  1. ^ See Lee (2003, Chapter 18) for a very concise statement and proof of Noether's theorem.

Notes

  1. ^ a b c d e Lee 2003, Chapter 18.
  2. ^ Lee 2003, Chapter 12.

Works cited

  • Abraham, Ralph; Marsden, Jerrold E. (1978). Foundations of Mechanics. London: Benjamin-Cummings. ISBN 978-080530102-1.See section 3.2.
  • Arnol'd, V.I. (1997). Mathematical Methods of Classical Mechanics. Berlin etc: Springer. ISBN 0-387-96890-3.
  • Frankel, Theodore (1997). The Geometry of Physics. Cambridge University Press. ISBN 0-521-38753-1.
  • Lee, J. M. (2003), Introduction to Smooth manifolds, Springer Graduate Texts in Mathematics, vol. 218, ISBN 0-387-95448-1
  • McDuff, Dusa; Salamon, D. (1998). Introduction to Symplectic Topology. Oxford Mathematical Monographs. ISBN 0-19-850451-9.

External links

  • Hamiltonian vector field on nLab