You are on page 1of 19

Quantization

(physics)

In physics, quantization is the process of


transition from a classical understanding
of physical phenomena to a newer
understanding known as quantum
mechanics. (It is a procedure for
constructing a quantum field theory
starting from a classical field theory.) This
is a generalization of the procedure for
building quantum mechanics from
classical mechanics. One also speaks of
field quantization, as in the "quantization
of the electromagnetic field", where one
refers to photons as field "quanta" (for
instance as light quanta). This procedure
is basic to theories of particle physics,
nuclear physics, condensed matter
physics, and quantum optics.

Quantization methods
Quantization converts classical fields into
operators acting on quantum states of the
field theory. The lowest energy state is
called the vacuum state. The reason for
quantizing a theory is to deduce properties
of materials, objects or particles through
the computation of quantum amplitudes,
which may be very complicated. Such
computations have to deal with certain
subtleties called renormalization, which, if
neglected, can often lead to nonsense
results, such as the appearance of
infinities in various amplitudes. The full
specification of a quantization procedure
requires methods of performing
renormalization.

The first method to be developed for


quantization of field theories was
canonical quantization. While this is
extremely easy to implement on
sufficiently simple theories, there are many
situations where other methods of
quantization yield more efficient
procedures for computing quantum
amplitudes. However, the use of canonical
quantization has left its mark on the
language and interpretation of quantum
field theory.

Canonical quantization

Canonical quantization of a field theory is


analogous to the construction of quantum
mechanics from classical mechanics. The
classical field is treated as a dynamical
variable called the canonical coordinate,
and its time-derivative is the canonical
momentum. One introduces a
commutation relation between these
which is exactly the same as the
commutation relation between a particle's
position and momentum in quantum
mechanics. Technically, one converts the
field to an operator, through combinations
of creation and annihilation operators. The
field operator acts on quantum states of
the theory. The lowest energy state is
called the vacuum state. The procedure is
also called second quantization.
This procedure can be applied to the
quantization of any field theory: whether of
fermions or bosons, and with any internal
symmetry. However, it leads to a fairly
simple picture of the vacuum state and is
not easily amenable to use in some
quantum field theories, such as quantum
chromodynamics which is known to have
a complicated vacuum characterized by
many different condensates.

Quantization schemes

Even within the setting of canonical


quantization, there is difficulty associated
to quantizing arbitrary observables on the
classical phase space. This is the ordering
ambiguity: Classically the position and
momentum variables x and p commute,
but their quantum mechanical
counterparts do not. Various quantization
schemes have been proposed to resolve
this ambiguity,[1] of which the most
popular is the Weyl quantization scheme.
Nevertheless, the Groenewold–van Hove
theorem says that no perfect quantization
scheme exists. Specifically, if the
quantizations of x and p are taken to be
the usual position and momentum
operators, then no quantization scheme
can perfectly reproduce the Poisson
bracket relations among the classical
observables.[2] See Groenewold's theorem
for one version of this result.

Covariant canonical
quantization

There is a way to perform a canonical


quantization without having to resort to
the non covariant approach of foliating
spacetime and choosing a Hamiltonian.
This method is based upon a classical
action, but is different from the functional
integral approach.

The method does not apply to all possible


actions (for instance, actions with a
noncausal structure or actions with gauge
"flows"). It starts with the classical algebra
of all (smooth) functionals over the
configuration space. This algebra is
quotiented over by the ideal generated by
the Euler–Lagrange equations. Then, this
quotient algebra is converted into a
Poisson algebra by introducing a Poisson
bracket derivable from the action, called
the Peierls bracket. This Poisson algebra
is then -deformed in the same way as in
canonical quantization.

There is also a way to quantize actions


with gauge "flows". It involves the Batalin–
Vilkovisky formalism, an extension of the
BRST formalism.
Deformation quantization

Geometric quantization

In mathematical physics, geometric


quantization is a mathematical approach
to defining a quantum theory
corresponding to a given classical theory.
It attempts to carry out quantization, for
which there is in general no exact recipe, in
such a way that certain analogies between
the classical theory and the quantum
theory remain manifest. For example, the
similarity between the Heisenberg
equation in the Heisenberg picture of
quantum mechanics and the Hamilton
equation in classical physics should be
built in.

One of the earliest attempts at a natural


quantization was Weyl quantization,
proposed by Hermann Weyl in 1927. Here,
an attempt is made to associate a
quantum-mechanical observable (a self-
adjoint operator on a Hilbert space) with a
real-valued function on classical phase
space. The position and momentum in this
phase space are mapped to the generators
of the Heisenberg group, and the Hilbert
space appears as a group representation
of the Heisenberg group. In 1946, H. J.
Groenewold [3] considered the product of a
pair of such observables and asked what
the corresponding function would be on
the classical phase space. This led him to
discover the phase-space star-product of a
pair of functions. More generally, this
technique leads to deformation
quantization, where the ★-product is taken
to be a deformation of the algebra of
functions on a symplectic manifold or
Poisson manifold. However, as a natural
quantization scheme (a functor), Weyl's
map is not satisfactory. For example, the
Weyl map of the classical angular-
momentum-squared is not just the
quantum angular momentum squared
operator, but it further contains a constant
term 3ħ2/2. (This extra term is actually
physically significant, since it accounts for
the nonvanishing angular momentum of
the ground-state Bohr orbit in the hydrogen
atom.[4] As a mere representation change,
however, Weyl's map underlies the
alternate Phase space formulation of
conventional quantum mechanics.

A more geometric approach to


quantization, in which the classical phase
space can be a general symplectic
manifold, was developed in the 1970s by
Bertram Kostant and Jean-Marie Souriau.
The method proceeds in two stages. First,
once constructs a "prequantum Hilbert
space" consisting of square-integrable
functions (or, more properly, sections of a
line bundle) over the phase space. Here
one can construct operators satisfying
commutation relations corresponding
exactly to the classical Poisson-bracket
relations. On the other hand, this
prequantum Hilbert space is too big to be
physically meaningful. One then restricts
to functions (or sections) depending on
half the variables on the phase space,
yielding the quantum Hilbert space.

Loop quantization

See Loop quantum gravity.


Path integral quantization

A classical mechanical theory is given by


an action with the permissible
configurations being the ones which are
extremal with respect to functional
variations of the action. A quantum-
mechanical description of the classical
system can also be constructed from the
action of the system by means of the path
integral formulation.

Quantum statistical mechanics


approach

See Uncertainty principle


Schwinger's variational
approach

See Schwinger's quantum action principle

See also
First quantization
Feynman path integral
Light front quantization
Photon polarization
Quantum Hall effect
Quantum number

References
Abraham, R. & Marsden (1985):
Foundations of Mechanics, ed. Addison–
Wesley, ISBN 0-8053-0102-X.
G. Giachetta, L. Mangiarotti, G.
Sardanashvily, Geometric and Algebraic
Topological Methods in Quantum
Mechanics (World Scientific, 2005)
ISBN 981-256-129-3.
Hall, Brian C. (2013), Quantum Theory for
Mathematicians, Graduate Texts in
Mathematics, 267, Springer
M. Peskin, D. Schroeder, An Introduction
to Quantum Field Theory (Westview
Press, 1995) ISBN 0-201-50397-2
Weinberg, Steven, The Quantum Theory
of Fields (3 volumes)
Ali, S. T., & Engliš, M. (2005).
"Quantization methods: a guide for
physicists and analysts". Reviews in
Mathematical Physics 17 (04), 391-490.
arXiv preprint
Todorov, Ivan (2012). "Quantization is a
mystery." arXiv preprint arXiv:1206.3116
(2012).

Notes
1. Hall 2013 Chapter 13
2. Hall 2013 Theorem 13.13
3. H.J. Groenewold, "On the Principles of
elementary quantum mechanics",
Physica,12 (1946) pp. 405–460
4. Dahl, J.; Schleich, W. (2002). "Concepts of
radial and angular kinetic energies".
Physical Review A 65 (2).
doi:10.1103/PhysRevA.65.022109.

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Quantization_(physics)&oldid=843274969"

Last edited 1 month ago by an anon…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like