Hamiltonian mechanics

Sir William Rowan Hamilton

In physics, Hamiltonian mechanics is a reformulation of Lagrangian mechanics that emerged in 1833. Introduced by Sir William Rowan Hamilton,[1] Hamiltonian mechanics replaces (generalized) velocities used in Lagrangian mechanics with (generalized) momenta. Both theories provide interpretations of classical mechanics and describe the same physical phenomena.

Hamiltonian mechanics has a close relationship with geometry (notably, symplectic geometry and Poisson structures) and serves as a link between classical and quantum mechanics.

Overview[edit]

Phase space coordinates (p, q) and Hamiltonian H[edit]

Let be a mechanical system with the configuration space and the smooth Lagrangian Select a standard coordinate system on The quantities are called momenta. (Also generalized momenta, conjugate momenta, and canonical momenta). For a time instant the Legendre transformation of is defined as the map which is assumed to have a smooth inverse For a system with degrees of freedom, the Lagrangian mechanics defines the energy function

The Legendre transform of turns into a function known as the Hamiltonian. The Hamiltonian satisfies

which implies that
where the velocities are found from the (-dimensional) equation which, by assumption, is uniquely solvable for . The (-dimensional) pair is called phase space coordinates. (Also canonical coordinates).

From Euler–Lagrange equation to Hamilton's equations[edit]

In phase space coordinates , the (-dimensional) Euler–Lagrange equation

becomes Hamilton's equations in dimensions

Proof

The Hamiltonian is the Legendre transform of the Lagrangian , thus one has

and thus

Besides, since the Euler-Lagrange equations yield

From stationary action principle to Hamilton's equations[edit]

Let be the set of smooth paths for which and The action functional is defined via

where , and (see above). A path is a stationary point of (and hence is an equation of motion) if and only if the path in phase space coordinates obeys the Hamilton's equations.

Basic physical interpretation[edit]

A simple interpretation of Hamiltonian mechanics comes from its application on a one-dimensional system consisting of one nonrelativistic particle of mass m. The value of the Hamiltonian is the total energy of the system, in this case the sum of kinetic and potential energy, traditionally denoted T and V, respectively. Here p is the momentum mv and q is the space coordinate. Then

T is a function of p alone, while V is a function of q alone (i.e., T and V are scleronomic).

In this example, the time derivative of q is the velocity, and so the first Hamilton equation means that the particle's velocity equals the derivative of its kinetic energy with respect to its momentum. The time derivative of the momentum p equals the Newtonian force, and so the second Hamilton equation means that the force equals the negative gradient of potential energy.

Example[edit]

A spherical pendulum consists of a mass m moving without friction on the surface of a sphere. The only forces acting on the mass are the reaction from the sphere and gravity. Spherical coordinates are used to describe the position of the mass in terms of (r, θ, φ), where r is fixed, r = .

Spherical pendulum: angles and velocities.

The Lagrangian for this system is[2]

Thus the Hamiltonian is

where
and
In terms of coordinates and momenta, the Hamiltonian reads
Hamilton's equations give the time evolution of coordinates and conjugate momenta in four first-order differential equations,
Momentum , which corresponds to the vertical component of angular momentum , is a constant of motion. That is a consequence of the rotational symmetry of the system around the vertical axis. Being absent from the Hamiltonian, azimuth is a cyclic coordinate, which implies conservation of its conjugate momentum.

Deriving Hamilton's equations[edit]

Hamilton's equations can be derived by a calculation with the Lagrangian , generalized positions qi, and generalized velocities qi, where .[3] Here we work off-shell, meaning , , are independent coordinates in phase space, not constrained to follow any equations of motion (in particular, is not a derivative of ). The total differential of the Lagrangian is:

The generalized momentum coordinates were defined as , so we may rewrite the equation as:

After rearranging, one obtains:

The term in parentheses on the left-hand side is just the Hamiltonian defined previously, therefore:

One may also calculate the total differential of the Hamiltonian with respect to coordinates , , instead of , , , yielding:

One may now equate these two expressions for , one in terms of , the other in terms of :

Since these calculations are off-shell, one can equate the respective coefficients of , , on the two sides:

On-shell, one substitutes parametric functions which define a trajectory in phase space with velocities , obeying Lagrange's equations:

Rearranging and writing in terms of the on-shell gives:

Thus Lagrange's equations are equivalent to Hamilton's equations:

In the case of time-independent and , i.e. , Hamilton's equations consist of 2n first-order differential equations, while Lagrange's equations consist of n second-order equations. Hamilton's equations usually do not reduce the difficulty of finding explicit solutions, but important theoretical results can be derived from them, because coordinates and momenta are independent variables with nearly symmetric roles.

Hamilton's equations have another advantage over Lagrange's equations: if a system has a symmetry, so that some coordinate does not occur in the Hamiltonian (i.e. a cyclic coordinate), the corresponding momentum coordinate is conserved along each trajectory, and that coordinate can be reduced to a constant in the other equations of the set. This effectively reduces the problem from n coordinates to (n − 1) coordinates: this is the basis of symplectic reduction in geometry. In the Lagrangian framework, the conservation of momentum also follows immediately, however all the generalized velocities still occur in the Lagrangian, and a system of equations in n coordinates still has to be solved.[4]

The Lagrangian and Hamiltonian approaches provide the groundwork for deeper results in classical mechanics, and suggest analogous formulations in quantum mechanics: the path integral formulation and the Schrödinger equation.

Properties of the Hamiltonian[edit]

  • The value of the Hamiltonian is the total energy of the system if and only if the energy function has the same property. (See definition of ).[clarification needed]
  • when , form a solution of Hamilton's equations.
    Indeed, and everything but the final term cancels out.
  • does not change under point transformations, i.e. smooth changes of space coordinates. (Follows from the invariance of the energy function under point transformations. The invariance of can be established directly).
  • (See § Deriving Hamilton's equations).
  • . (Compare Hamilton's and Euler-Lagrange equations or see § Deriving Hamilton's equations).
  • if and only if .
    A coordinate for which the last equation holds is called cyclic (or ignorable). Every cyclic coordinate reduces the number of degrees of freedom by , causes the corresponding momentum to be conserved, and makes Hamilton's equations easier to solve.

Hamiltonian as the total system energy[edit]

In its application to a given system, the Hamiltonian is often taken to be

where is the kinetic energy and is the potential energy. Using this relation can be simpler than first calculating the Lagrangian, and then deriving the Hamiltonian from the Lagrangian. However, the relation is not true for all systems.

The relation holds true for nonrelativistic systems when all of the following conditions are satisfied[5][6]

where is time, is the number of degrees of freedom of the system, and each is an arbitrary scalar function of .

In words, this means that the relation holds true if does not contain time as an explicit variable (it is scleronomic), does not contain generalised velocity as an explicit variable, and each term of is quadratic in generalised velocity.

Proof[edit]

Preliminary to this proof, it is important to address an ambiguity in the related mathematical notation. While a change of variables can be used to equate , it is important to note that . In this case, the right hand side always evaluates to 0. To perform a change of variables inside of a partial derivative, the multivariable chain rule should be used. Hence, to avoid ambiguity, the function arguments of any term inside of a partial derivative should be stated.

Additionally, this proof uses the notation to imply that .

Proof

Starting from definitions of the Hamiltonian, generalized momenta, and Lagrangian for an degrees of freedom system

Substituting the generalized momenta into the Hamiltonian gives

Substituting the Lagrangian into the result gives

Now assume that

and also assume that

Applying these assumptions results in

Next assume that T is of the form

where each is an arbitrary scalar function of .

Differentiating this with respect to , , gives

Splitting the summation, evaluating the partial derivative, and rejoining the summation gives

Summing (this multiplied by ) over results in

This simplification is a result of Euler's homogeneous function theorem.

Hence, the Hamiltonian becomes

Application to systems of point masses[edit]

For a system of point masses, the requirement for to be quadratic in generalised velocity is always satisfied for the case where , which is a requirement for anyway.

Proof

Consider the kinetic energy for a system of N point masses. If it is assumed that , then it can be shown that (See Scleronomous § Application). Therefore, the kinetic energy is

The chain rule for many variables can be used to expand the velocity

Resulting in

This is of the required form.

Conservation of energy[edit]

If the conditions for are satisfied, then conservation of the Hamiltonian implies conservation of energy. This requires the additional condition that does not contain time as an explicit variable.

With respect to the extended Euler-Lagrange formulation (See Lagrangian mechanics § Extensions to include non-conservative forces), the Rayleigh dissipation function represents energy dissipation by nature. Therefore, energy is not conserved when . This is similar to the velocity dependent potential.

In summary, the requirements for to be satisfied for a nonrelativistic system are[5][6]

  1. is a homogeneous quadratic function in

Hamiltonian of a charged particle in an electromagnetic field[edit]

A sufficient illustration of Hamiltonian mechanics is given by the Hamiltonian of a charged particle in an electromagnetic field. In Cartesian coordinates the Lagrangian of a non-relativistic classical particle in an electromagnetic field is (in SI Units):

where q is the electric charge of the particle, φ is the electric scalar potential, and the Ai are the components of the magnetic vector potential that may all explicitly depend on and .

This Lagrangian, combined with Euler–Lagrange equation, produces the Lorentz force law

and is called minimal coupling.

The canonical momenta are given by:

The Hamiltonian, as the Legendre transformation of the Lagrangian, is therefore:

This equation is used frequently in quantum mechanics.

Under gauge transformation:

where f(r, t) is any scalar function of space and time. The aforementioned Lagrangian, the canonical momenta, and the Hamiltonian transform like:
which still produces the same Hamilton's equation:

In quantum mechanics, the wave function will also undergo a local U(1) group transformation[7] during the Gauge Transformation, which implies that all physical results must be invariant under local U(1) transformations.

Relativistic charged particle in an electromagnetic field[edit]

The relativistic Lagrangian for a particle (rest mass and charge ) is given by:

Thus the particle's canonical momentum is

that is, the sum of the kinetic momentum and the potential momentum.

Solving for the velocity, we get

So the Hamiltonian is

This results in the force equation (equivalent to the Euler–Lagrange equation)

from which one can derive