Charge (physics)

Source: Wikipedia, the free encyclopedia.

In

commute with the Hamiltonian
. Charges are often denoted by , and so the invariance of the charge corresponds to the vanishing commutator , where is the Hamiltonian. Thus, charges are associated with conserved quantum numbers; these are the eigenvalues of the generator .

Abstract definition

Abstractly, a charge is any generator of a continuous symmetry of the physical system under study. When a physical system has a symmetry of some sort, Noether's theorem implies the existence of a conserved current. The thing that "flows" in the current is the "charge", the charge is the generator of the (local) symmetry group. This charge is sometimes called the Noether charge.

Thus, for example, the

U(1) symmetry of electromagnetism. The conserved current is the electric current
.

In the case of local, dynamical symmetries, associated with every charge is a

gauge field; when quantized, the gauge field becomes a gauge boson. The charges of the theory "radiate" the gauge field. Thus, for example, the gauge field of electromagnetism is the electromagnetic field; and the gauge boson is the photon
.

The word "charge" is often used as a synonym for both the generator of a symmetry, and the conserved quantum number (eigenvalue) of the generator. Thus, letting the upper-case letter Q refer to the generator, one has that the generator commutes with the Hamiltonian [Q, H] = 0. Commutation implies that the eigenvalues (lower-case) q are time-invariant: dq/dt = 0.

So, for example, when the symmetry group is a

discreteness of the root system accounting for the quantization of the charge. The simple roots are used, as all the other roots can be obtained as linear combinations of these. The general roots are often called raising and lowering operators, or ladder operators
.

The charge quantum numbers then correspond to the weights of the

highest-weight modules of a given representation of the Lie algebra. So, for example, when a particle in a quantum field theory
belongs to a symmetry, then it transforms according to a particular representation of that symmetry; the charge quantum number is then the weight of the representation.

Examples

Various charge quantum numbers have been introduced by theories of particle physics. These include the charges of the Standard Model:

Note that these charge quantum numbers show up in the Lagrangian via the Gauge covariant derivative#Standard_Model.

Charges of approximate symmetries:

Hypothetical charges of extensions to the Standard Model:

In supersymmetry:

  • The supercharge refers to the generator that rotates the fermions into bosons, and vice versa, in the supersymmetry.

In conformal field theory:

  • The
    energy–momentum tensor of the two-dimensional conformal field theory.[1]

In

gravitation
:

Charge conjugation

In the formalism of particle theories, charge-like quantum numbers can sometimes be inverted by means of a

adjoint representation
of the group.

Thus, a common example is that the

SL(2,C) (the spinors) forms the adjoint rep of the Lorentz group
SO(3,1); abstractly, one writes

That is, the product of two (Lorentz) spinors is a (Lorentz) vector and a (Lorentz) scalar. Note that the complex Lie algebra sl(2,C) has a

Clebsch–Gordan coefficients
.

A similar phenomenon occurs in the compact group

SU(3)
, where there are two charge-conjugate but inequivalent fundamental representations, dubbed and , the number 3 denoting the dimension of the representation, and with the quarks transforming under and the antiquarks transforming under . The Kronecker product of the two gives

That is, an eight-dimensional representation, the octet of the

. The decomposition of such products of representations into direct sums of irreducible representations can in general be written as

for representations . The dimensions of the representations obey the "dimension sum rule":

Here, is the dimension of the representation , and the integers being the

Littlewood–Richardson coefficients
. The decomposition of the representations is again given by the Clebsch–Gordan coefficients, this time in the general Lie-algebra setting.

See also

  • Casimir operator

References