Hamiltonian mechanics

See also classical Hamiltonian system.
The states of a system of Classical Mechanics are determined by generalized variables and their derivatives. In other words, we work in the tangent bundle of the configuration space, or the jet bundle of order 1. That is part of the formulation of Lagrangian Mechanics. The Legendre transform allows us to move everything to the cotangent bundle (phase space). The states of the system are determined by the original variables and generalized momenta. It may seem cumbersome, but it has advantages in geometric interpretation. The systems (dynamical systems) with this origin are called classical Hamiltonian systems. Some of them are called integrable systems. Important example: harmonic oscillator.

The Hamiltonian mechanics formalism can be summarized as follows (modified from "Review of CM and QM by Terence Tao"):

  1. The physical system has a phase space Ω of states x (often parameterized by position variables q and momentum variables p), which mathematically corresponds to a symplectic manifold with a symplectic form ω (e.g., ω=dpdq for position and momentum coordinates).
  2. The complete state of the system at any time t is represented by a point x(t) in the phase space Ω.
  3. Every physical observable A (e.g., energy, momentum, position) is associated with a function A that maps the phase space Ω to the range of the observable (e.g., for real observables, A maps Ω to R). Measuring the observable A at time t yields the measurement A(x(t)).
  4. The Hamiltonian H:ΩR is a special observable that governs the evolution of the state x(t) over time through Hamiltonian equations of motion. In terms of position and momentum coordinates x(t)=(qi(t),pi(t))i=1n, these equations are given by:
pit=Hqi,qit=Hpi.

More abstractly, using the symplectic form ω, the equations of motion can be written as:

x(t)t=ωH(x(t)),(2)

where ωH is the symplectic gradient of H.

Hamilton's equations of motion can also be expressed in a dual form using observables A as Poisson's equations of motion:

A(x(t))t={H,A}(x(t)),

where {H,A}=ωHA is the induced Poisson bracket (here we consider an arbitrary Riemannian metric defined on Ω in order to define A and the dot product. See relation symplectic form and Riemannian metric). In a more abstract form, Poisson's equation can be written as:

At={H,A},(3)

where {H,A} represents the Poisson bracket. This has to do with Heisenberg vs Schrodinger picture (?).

Variational principle for Hamiltonian mechanics

(From C. Rovelli "Forget time")
Denote γ~ as a curve in the phase space Ω (observables and momenta) and γ its restriction to C (observables alone). The Hamiltonian H:ΩR determines the physical motions via the following variational principle:
A curve γ connecting the events qa1 and qa2 is a physical motion if γ~ extremizes the action

S[γ~]=γ~padqa

in the class of the curves γ~ satisfying H(qa,pa)=0 whose restriction γ to C connects qa1 and qa2. The action is the integration of the tautological 1-form along the curve γ~.

All known physical (relativistic and nonrelativistic) Hamiltonian systems can be formulated in this manner.
See relativistic Hamiltonian mechanics.

This is equivalent to the variational principle of Lagrangian Mechanics. The Hamiltonian is given by:

H(qa,pa)=paq˙aL(qa,q˙a),

so the constraint H(qa,pa)=0 becomes paq˙a=L(qa,q˙a).
Therefore,

S[γ~]=γ~padqa=paq˙adt.

and substituting paq˙a=L(qa,q˙a) from the Hamiltonian constraint:

S[γ~]=L(qa,q˙a)dt.

Thus, the action in the Hamiltonian formalism reduces to the action in the Lagrangian formalism:

S[γ]=L(qa,q˙a)dt.

Mixed states (statistical mechanics (?))

In the formalism above, we assume the system is in a pure state at each time t, occupying a single point x(t) in phase space. However, mixed states can also be considered, where the state of the system at time t is described by a probability distribution ρ(t,x)dx on the phase space. Measuring an observable A at time t becomes a random variable, and its expectation A is given by:

A(t)=ΩA(x)ρ(t,x)dx,(4)

The equation of motion for a mixed state ρ is given by the advection equation:

ρt=div(ρωH)

using the same vector field ωH as in equation (2). This equation can also be derived from equations (3), (4), and a duality argument.

Pure states can be seen as a special case of mixed states, where the probability distribution ρ(t,x)dx is a Dirac delta δx(t)(x). Mixed states can be thought of as continuous averages of pure states, or equivalently, the space of mixed states is the convex hull of the space of pure states.

Consider a system of 2 particles with a joint phase space Ω=Ω1×Ω2, where Ω1 and Ω2 are the individual one-particle phase spaces. A pure joint state is represented by a point x=(x1,x2) in Ω, where x1 and x2 represent the states of the first and second particles, respectively. If the joint Hamiltonian H:ΩR splits as:

H(x1,x2)=H1(x1)+H2(x2),

then the equations of motion for the first and second particles are completely decoupled, with no interactions between them. However, in practice, the joint Hamiltonian contains coupling terms between x1 and x2 that prevent total decoupling. For instance, the joint Hamiltonian may be:

H(x1,x2)=|p1|22m1+|p2|22m2+V(q1q2),

where x1=(q1,p1) and x2=(q2,p2) are position and momentum coordinates, m1 and m2 are mass constants, and V(q1q2) represents the interaction potential depending on the spatial separation q1q2 between the particles.

Similarly, a mixed joint state is a joint probability distribution ρ(x1,x2)dx1dx2 on the product state space. To obtain the (mixed) state of an individual particle, we consider marginal distributions such as:

ρ1(x1)=Ω2ρ(x1,x2)dx2

for the first particle or

ρ2(x2)=Ω1ρ(x1,x2)dx1

for the second particle. For N-particle systems, if the joint distribution of N distinct particles is given by ρ(x1,,xN)dx1dxN, then the distribution of the first particle is:

ρ1(x1)=Ω2××ΩNρ(x1,x2,,xN)dx2dxN,

and the distribution of the first two particles is:

ρ12(x1,x2)=Ω3××ΩNρ(x1,x2,,xN)dx3dxN,

and so on.

Old stuff

Symplectic form

See symplectic form.

Hamiltonian vector fields

See Hamiltonian vector fields.
This symplectic form has a main use: to convert the differential of the hamiltonian, dH, into a vector field called the symplectic gradient: the only vector field XH such that ω(XH,)=dH. If we have also a Riemannian metric g, it turns out that this symplectic gradient and the usual gradient are orthogonal: remember that given a function H the gradient is the only vector field H such that dH=g(H,). Therefore

g(H,XH)=dH(XH)=ω(XH,XH)=0

It's important to observe that:

  1. iXHω=dH
  2. LXHω=0. It is proven by Cartan's formula.
  3. LXHωn=0. This 2n- form is a volume form. So the flow over the Hamiltonian vector field conserve the volume. This fact is known as Liouville theorem.

More things about:

X(H)=dH(X)=iX(ω)(X)=0

(maybe with a minus sign?)

So H is constant along the ϕ curves:

X(H)=ddtHϕ D=ddt(Fϕ)=X(F)=dF(X)

But observe that if we take Y such that iYω=dF then

dF(X)=iYω(X)=ω(Y,X)==ω(X,Y)=iXω(Y)=dH(Y)=Y(H)

So the interesting quantity D can be computed in two ways

X(F)=Y(H)

This motivate a new definition. Remove, temporarily, the special paper to H. For the functions F,H we define the Poisson brackets as the new function:

{F,H}=X(F)=Y(H)=ω(X,Y)

It is satisfied:

Xf={,f} {f,{g,h}}={{f,h},g}+{h,{f,g}}

for any three functions f,g,h.
Leaving an slot instead of f we get

{,{g,h}}=[{,g},{,h}]

expression that can be rewritten as

X{g,h}=[Xg,Xh]=Xω(Xg,Xh) {F,pi}=Xpi(F)

but Xpi is just qi

{f,g}=Xg(f)=df(Xg)=fXg Xf={,f}

Then, the flow theorem for vector fields let you assure that there exists the local flow

ϕs:TQTQ

such that

  1. ϕ0=id

  2. ddsϕs(m)|s=0=Xf

  3. ϕsϕt=ϕs+t

This can be seen also as a flow of observables in CM.