Strahler number

From formulasearchengine
Revision as of 23:28, 18 October 2013 by en>Anrnusna (→‎References: journal name, replaced: The European Physical Journal B - Condensed Matter and Complex Systems → European Physical Journal B using AWB)
Jump to navigation Jump to search

In classical mechanics, action-angle coordinates are a set of canonical coordinates useful in solving many integrable systems. The method of action-angles is useful for obtaining the frequencies of oscillatory or rotational motion without solving the equations of motion. Action-angle coordinates are chiefly used when the Hamilton–Jacobi equations are completely separable. (Hence, the Hamiltonian does not depend explicitly on time, i.e., the energy is conserved.) Action-angle variables define an invariant torus, so called because holding the action constant defines the surface of a torus, while the angle variables provide the coordinates on the torus.

The Bohr–Sommerfeld quantization conditions, used to develop quantum mechanics before the advent of wave mechanics, state that the action must be an integral multiple of Planck's constant; similarly, Einstein's insight into EBK quantization and the difficulty of quantizing non-integrable systems was expressed in terms of the invariant tori of action-angle coordinates.

Action-angle coordinates are also useful in perturbation theory of Hamiltonian mechanics, especially in determining adiabatic invariants. One of the earliest results from chaos theory, for the non-linear perturbations of dynamical systems with a small number of degrees of freedom is the KAM theorem, which states that the invariant tori are stable under small perturbations.

The use of action-angle variables was central to the solution of the Toda lattice, and to the definition of Lax pairs, or more generally, the idea of the isospectral evolution of a system.

Derivation

Action angles result from a type-2 canonical transformation where the generating function is Hamilton's characteristic function W(q) (not Hamilton's principal function S). Since the original Hamiltonian does not depend on time explicitly, the new Hamiltonian K(w,J) is merely the old Hamiltonian H(q,p) expressed in terms of the new canonical coordinates, which we denote as w (the action angles, which are the generalized coordinates) and their new generalized momenta J. We will not need to solve here for the generating function W itself; instead, we will use it merely as a vehicle for relating the new and old canonical coordinates.

Rather than defining the action angles w directly, we define instead their generalized momenta, which resemble the classical action for each original generalized coordinate

Jkpkdqk

where the integration path is implicitly given by the constant energy function E=E(qk,pk). Since the actual motion is not involved in this integration, these generalized momenta Jk are constants of the motion, implying that the transformed Hamiltonian K does not depend on the conjugate generalized coordinates wk

ddtJk=0=Kwk

where the wk are given by the typical equation for a type-2 canonical transformation

wkWJk

Hence, the new Hamiltonian K=K(J) depends only on the new generalized momenta J.

The dynamics of the action angles is given by Hamilton's equations

ddtwk=KJkνk(J)

The right-hand side is a constant of the motion (since all the J's are). Hence, the solution is given by

wk=νk(J)t+βk

where βk is a constant of integration. In particular, if the original generalized coordinate undergoes an oscillation or rotation of period T, the corresponding action angle wk changes by Δwk=νk(J)T.

These νk(J) are the frequencies of oscillation/rotation for the original generalized coordinates qk. To show this, we integrate the net change in the action angle wk over exactly one complete variation (i.e., oscillation or rotation) of its generalized coordinates qk

Δwkwkqkdqk=2WJkqkdqk=ddJkWqkdqk=ddJkpkdqk=dJkdJk=1

Setting the two expressions for Δwk equal, we obtain the desired equation

νk(J)=1T

The action angles w are an independent set of generalized coordinates. Thus, in the general case, each original generalized coordinate qk can be expressed as a Fourier series in all the action angles

qk=s1=s2=sN=As1,s2,,sNkei2πs1w1ei2πs2w2ei2πsNwN

where As1,s2,,sNk is the Fourier series coefficient. In most practical cases, however, an original generalized coordinate qk will be expressible as a Fourier series in only its own action angles wk

qk=sk=ei2πskwk

Summary of basic protocol

The general procedure has three steps:

  1. Calculate the new generalized momenta Jk
  2. Express the original Hamiltonian entirely in terms of these variables.
  3. Take the derivatives of the Hamiltonian with respect to these momenta to obtain the frequencies νk

Degeneracy

In some cases, the frequencies of two different generalized coordinates are identical, i.e., νk=νl for kl. In such cases, the motion is called degenerate.

Degenerate motion signals that there are additional general conserved quantities; for example, the frequencies of the Kepler problem are degenerate, corresponding to the conservation of the Laplace–Runge–Lenz vector.

Degenerate motion also signals that the Hamilton–Jacobi equations are completely separable in more than one coordinate system; for example, the Kepler problem is completely separable in both spherical coordinates and parabolic coordinates.

See also

References

  • L. D. Landau and E. M. Lifshitz, (1976) Mechanics, 3rd. ed., Pergamon Press. ISBN 0-08-021022-8 (hardcover) and ISBN 0-08-029141-4 (softcover).
  • H. Goldstein, (1980) Classical Mechanics, 2nd. ed., Addison-Wesley. ISBN 0-201-02918-9