Classical limit

Source: Wikipedia, the free encyclopedia.

The classical limit or correspondence limit is the ability of a physical theory to approximate or "recover" classical mechanics when considered over special values of its parameters.[1] The classical limit is used with physical theories that predict non-classical behavior.

Quantum theory

A heuristic postulate called the correspondence principle was introduced to quantum theory by Niels Bohr: in effect it states that some kind of continuity argument should apply to the classical limit of quantum systems as the value of the Planck constant normalized by the action of these systems becomes very small. Often, this is approached through "quasi-classical" techniques (cf. WKB approximation).[2]

More rigorously,

Weyl quantization.) Thus typically, quantum commutators (equivalently, Moyal brackets) reduce to Poisson brackets,[4] in a group contraction
.

In

coherent states. It is less clear, however, how the classical limit applies to chaotic systems, a field known as quantum chaos
.

Quantum mechanics and classical mechanics are usually treated with entirely different formalisms: quantum theory using

phase space formulation of quantum mechanics, which is statistical in nature, logical connections between quantum mechanics and classical statistical mechanics are made, enabling natural comparisons between them, including the violations of Liouville's theorem (Hamiltonian) upon quantization.[5][6]

In a crucial paper (1933),

extremal macroscopic actions S » ħ obliterate amplitude contributions in the path integral he introduced, leaving the extremal action Sclass, thus the classical action path as the dominant contribution, an observation further elaborated by Feynman in his 1942 PhD dissertation.[8] (Further see quantum decoherence
.)

Time-evolution of expectation values

One simple way to compare classical to quantum mechanics is to consider the time-evolution of the expected position and expected momentum, which can then be compared to the time-evolution of the ordinary position and momentum in classical mechanics. The quantum expectation values satisfy the Ehrenfest theorem. For a one-dimensional quantum particle moving in a potential , the Ehrenfest theorem says[9]

Although the first of these equations is consistent with the classical mechanics, the second is not: If the pair were to satisfy Newton's second law, the right-hand side of the second equation would have read

.

But in most cases,

.

If for example, the potential is cubic, then is quadratic, in which case, we are talking about the distinction between and , which differ by .

An exception occurs in case when the classical equations of motion are linear, that is, when is quadratic and is linear. In that special case, and do agree. In particular, for a free particle or a quantum harmonic oscillator, the expected position and expected momentum exactly follows solutions of Newton's equations.

For general systems, the best we can hope for is that the expected position and momentum will approximately follow the classical trajectories. If the wave function is highly concentrated around a point , then and will be almost the same, since both will be approximately equal to . In that case, the expected position and expected momentum will remain very close to the classical trajectories, at least for as long as the wave function remains highly localized in position.[10]

Now, if the initial state is very localized in position, it will be very spread out in momentum, and thus we expect that the wave function will rapidly spread out, and the connection with the classical trajectories will be lost. When the Planck constant is small, however, it is possible to have a state that is well localized in both position and momentum. The small uncertainty in momentum ensures that the particle remains well localized in position for a long time, so that expected position and momentum continue to closely track the classical trajectories for a long time.

Relativity and other deformations

Other familiar deformations in physics involve:

See also

References

  1. .
  2. .
  3. .
  4. .
  5. .
  6. ^ Conversely, in the lesser-known approach presented in 1932 by Koopman and von Neumann, the dynamics of classical mechanics have been formulated in terms of an operational formalism in Hilbert space, a formalism used conventionally for quantum mechanics.
  7. ^ Dirac, P.A.M. (1933). "The Lagrangian in quantum mechanics" (PDF). Physikalische Zeitschrift der Sowjetunion. 3: 64–72.
  8. ^ Feynman, R. P. (1942). The Principle of Least Action in Quantum Mechanics (Ph.D. Dissertation). Princeton University.
    Reproduced in Feynman, R. P. (2005). Brown, L. M. (ed.). Feynman's Thesis: a New Approach to Quantum Theory. .
  9. ^ Hall 2013 Section 3.7.5
  10. ^ Hall 2013 p. 78
  • Hall, Brian C. (2013), Quantum Theory for Mathematicians, Graduate Texts in Mathematics, vol. 267, Springer,