Canonical commutation relation


In quantum mechanics, the canonical commutation relation is the fundamental relation between canonical conjugate quantities. For example,
between the position operator and momentum operator in the direction of a point particle in one dimension, where is the commutator of and , is the imaginary unit, and is the reduced Planck's constant . In general, position and momentum are vectors of operators and their commutation relation between different components of position and momentum can be expressed as
where is the Kronecker delta.
This relation is attributed to Max Born, who called it a "quantum condition" serving as a postulate of the theory; it was noted by E. Kennard to imply the Heisenberg uncertainty principle. The Stone–von Neumann theorem gives a uniqueness result for operators satisfying the canonical commutation relation.

Relation to classical mechanics

By contrast, in classical physics, all observables commute and the commutator would be zero. However, an analogous relation exists, which is obtained by replacing the commutator with the Poisson bracket multiplied by,
This observation led Dirac to propose that the quantum counterparts , of classical observables, satisfy
In 1946, Hip Groenewold demonstrated that a general systematic correspondence between quantum commutators and Poisson brackets could not hold consistently.
However, he further appreciated that such a systematic correspondence does, in fact, exist between the quantum commutator and a deformation of the Poisson bracket, today called the Moyal bracket, and, in general, quantum operators and classical observables and distributions in phase space. He thus finally elucidated the consistent correspondence mechanism, the Wigner–Weyl transform, that underlies an alternate equivalent mathematical representation of quantum mechanics known as deformation quantization.

The Weyl relations

The group generated by exponentiation of the 3-dimensional Lie algebra determined by the commutation relation is called the Heisenberg group. This group can be realized as the group of upper triangular matrices with ones on the diagonal.
According to the standard mathematical formulation of quantum mechanics, quantum observables such as and should be represented as self-adjoint operators on some Hilbert space. It is relatively easy to see that two operators satisfying the above canonical commutation relations cannot both be bounded. Certainly, if and were trace class operators, the relation gives a nonzero number on the right and zero on the left.
Alternately, if and were bounded operators, note that, hence the operator norms would satisfy
However, can be arbitrarily large, so at least one operator cannot be bounded, and the dimension of the underlying Hilbert space cannot be finite. If the operators satisfy the Weyl relations then as a consequence of the Stone–von Neumann theorem, both operators must be unbounded.
Still, these canonical commutation relations can be rendered somewhat "tamer" by writing them in terms of the unitary operators and. The resulting braiding relations for these operators are the so-called Weyl relations
These relations may be thought of as an exponentiated version of the canonical commutation relations; they reflect that translations in position and translations in momentum do not commute. One can easily reformulate the Weyl relations in terms of the representations of the Heisenberg group.
The uniqueness of the canonical commutation relations—in the form of the Weyl relations—is then guaranteed by the Stone–von Neumann theorem.
It is important to note that for technical reasons, the Weyl relations are not strictly equivalent to the canonical commutation relation. If and were bounded operators, then a special case of the Baker–Campbell–Hausdorff formula would allow one to "exponentiate" the canonical commutation relations to the Weyl relations. Since, as we have noted, any operators satisfying the canonical commutation relations must be unbounded, the Baker–Campbell–Hausdorff formula does not apply without additional domain assumptions. Indeed, counterexamples exist satisfying the canonical commutation relations but not the Weyl relations. These technical issues are the reason that the Stone–von Neumann theorem is formulated in terms of the Weyl relations.
A discrete version of the Weyl relations, in which the parameters s and t range over, can be realized on a finite-dimensional Hilbert space by means of the.

Generalizations

The simple formula
valid for the quantization of the simplest classical system, can be generalized to the case of an arbitrary Lagrangian. We identify canonical coordinates and canonical momenta :
This definition of the canonical momentum ensures that one of the Euler–Lagrange equations has the form
The canonical commutation relations then amount to
where is the Kronecker delta.
Further, it can be easily shown that
Using, it can be easily shown that by mathematical induction

Gauge invariance

Canonical quantization is applied, by definition, on canonical coordinates. However, in the presence of an electromagnetic field, the canonical momentum is not gauge invariant. The correct gauge-invariant momentum is
where is the particle's electric charge, is the vector potential, and is the speed of light. Although the quantity is the "physical momentum", in that it is the quantity to be identified with momentum in laboratory experiments, it does not satisfy the canonical commutation relations; only the canonical momentum does that. This can be seen as follows.
The non-relativistic Hamiltonian for a quantized charged particle of mass in a classical electromagnetic field is
where is the three-vector potential and is the scalar potential. This form of the Hamiltonian, as well as the Schrödinger equation , the Maxwell equations and the Lorentz force law are invariant under the gauge transformation
where
and Λ=Λ is the gauge function.
The angular momentum operator is
and obeys the canonical quantization relations
defining the Lie algebra for so, where is the Levi-Civita symbol. Under gauge transformations, the angular momentum transforms as
The gauge-invariant angular momentum is given by
which has the commutation relations
where
is the magnetic field. The inequivalence of these two formulations shows up in the Zeeman effect and the Aharonov–Bohm effect.

Uncertainty relation and commutators

All such nontrivial commutation relations for pairs of operators lead to corresponding uncertainty relations, involving positive semi-definite expectation contributions by their respective commutators and anticommutators. In general, for two Hermitian operators and, consider expectation values in a system in the state, the variances around the corresponding expectation values being, etc.
Then
where is the commutator of and, and is the anticommutator.
This follows through use of the Cauchy–Schwarz inequality, since
, and ; and similarly for the shifted operators and .
Substituting for and yield Heisenberg's familiar uncertainty relation for and, as usual.

Uncertainty relation for angular momentum operators

For the angular momentum operators, etc., one has that
where is the Levi-Civita symbol and simply reverses the sign of the answer under pairwise interchange of the indices. An analogous relation holds for the spin operators.
Here, for and, in angular momentum multiplets, one has, for the transverse components of the Casimir invariant , the -symmetric relations
as well as .
Consequently, the above inequality applied to this commutation relation specifies
hence
and therefore
so, then, it yields useful constraints such as a lower bound on the Casimir invariant:, and hence, among others.