Parity (physics)


In quantum mechanics, a parity transformation is the flip in the sign of one spatial coordinate. In three dimensions, it can also refer to the simultaneous flip in the sign of all three spatial coordinates :
It can also be thought of as a test for chirality of a physical phenomenon, in that a parity inversion transforms a phenomenon into its mirror image. All fundamental interactions of elementary particles, with the exception of the weak interaction, are symmetric under parity. The weak interaction is chiral and thus provides a means for probing chirality in physics. In interactions that are symmetric under parity, such as electromagnetism in atomic and molecular physics, parity serves as a powerful controlling principle underlying quantum transitions.
A matrix representation of P has determinant equal to −1, and hence is distinct from a rotation, which has a determinant equal to 1. In a two-dimensional plane, a simultaneous flip of all coordinates in sign is not a parity transformation; it is the same as a 180°-rotation.
In quantum mechanics, wave functions that are unchanged by a parity transformation are described as even functions, while those that change sign under a parity transformation are odd functions.

Simple symmetry relations

Under rotations, classical geometrical objects can be classified into scalars, vectors, and tensors of higher rank. In classical physics, physical configurations need to transform under representations of every symmetry group.
Quantum theory predicts that states in a Hilbert space do not need to transform under representations of the group of rotations, but only under projective representations. The word projective refers to the fact that if one projects out the phase of each state, where we recall that the overall phase of a quantum state is not an observable, then a projective representation reduces to an ordinary representation. All representations are also projective representations, but the converse is not true, therefore the projective representation condition on quantum states is weaker than the representation condition on classical states.
The projective representations of any group are isomorphic to the ordinary representations of a central extension of the group. For example, projective representations of the 3-dimensional rotation group, which is the special orthogonal group SO, are ordinary representations of the special unitary group SU. Projective representations of the rotation group that are not representations are called spinors, and so quantum states may transform not only as tensors but also as spinors.
If one adds to this a classification by parity, these can be extended, for example, into notions of
One can define reflections such as
which also have negative determinant and form a valid parity transformation. Then, combining them with rotations one can recover the particular parity transformation defined earlier. The first parity transformation given does not work in an even number of dimensions, though, because it results in a positive determinant. In even dimensions only the latter example of a parity transformation can be used.
Parity forms the abelian group due to the relation. All Abelian groups have only one-dimensional irreducible representations. For, there are two irreducible representations: one is even under parity,, the other is odd,. These are useful in quantum mechanics. However, as is elaborated below, in quantum mechanics states need not transform under actual representations of parity but only under projective representations and so in principle a parity transformation may rotate a state by any phase.

Classical mechanics

Newton's equation of motion equates two vectors, and hence is invariant under parity. The law of gravity also involves only vectors and is also, therefore, invariant under parity.
However, angular momentum is an axial vector,
In classical electrodynamics, the charge density is a scalar, the electric field,, and current are vectors, but the magnetic field, is an axial vector. However, Maxwell's equations are invariant under parity because the curl of an axial vector is a vector.

Effect of spatial inversion on some variables of classical physics

Even

Classical variables, predominantly scalar quantities, which do not change upon spatial inversion include:

Odd

Classical variables, predominantly vector quantities, which have their sign flipped by spatial inversion include:

Quantum mechanics

Possible eigenvalues

In quantum mechanics, spacetime transformations act on quantum states. The parity transformation,, is a unitary operator, in general acting on a state as follows:.
One must then have, since an overall phase is unobservable. The operator, which reverses the parity of a state twice, leaves the spacetime invariant, and so is an internal symmetry which rotates its eigenstates by phases. If is an element of a continuous U symmetry group of phase rotations, then is part of this U and so is also a symmetry. In particular, we can define, which is also a symmetry, and so we can choose to call our parity operator, instead of. Note that and so has eigenvalues. Wave functions with eigenvalue +1 under a parity transformation are even functions, while eigenvalue −1 corresponds to odd functions. However, when no such symmetry group exists, it may be that all parity transformations have some eigenvalues which are phases other than .
For electronic wavefunctions, even states are usually indicated by a subscript g for gerade and odd states by a subscript u for ungerade. For example, the lowest energy level of the hydrogen molecule ion is labelled and the next-closest energy level is labelled.
The wave functions of a particle moving into an external potential, which is centrosymmetric, either remain invariable or change signs: these two possible states are called the even state or odd state of the wave functions.
The law of conservation of parity of particle states that, if an isolated ensemble of particles has a definite parity, then the parity remains invariable in the process of ensemble evolution.
The parity of the states of a particle moving in a spherically symmetric external field is determined by the angular momentum, and the particle state is defined by three quantum numbers: total energy, angular momentum and the projection of angular momentum.

Consequences of parity symmetry

When parity generates the Abelian group2, one can always take linear combinations of quantum states such that they are either even or odd under parity. Thus the parity of such states is ±1. The parity of a multiparticle state is the product of the parities of each state; in other words parity is a multiplicative quantum number.
In quantum mechanics, Hamiltonians are invariant under a parity transformation if commutes with the Hamiltonian. In non-relativistic quantum mechanics, this happens for any scalar potential, i.e.,, hence the potential is spherically symmetric. The following facts can be easily proven:
Some of the non-degenerate eigenfunctions of are unaffected by parity and the others are merely reversed in sign when the Hamiltonian operator and the parity operator commute:
where is a constant, the eigenvalue of,

Many-particle systems: atoms, molecules, nuclei

The overall parity of a many-particle system is the product of the parities of the one-particle states. It is −1 if an odd number of particles are in odd-parity states, and +1 otherwise. Different notations are in use to denote the parity of nuclei, atoms, and molecules.

Atoms

s have parity , where the exponent ℓ is the azimuthal quantum number. The parity is odd for orbitals p, f, … with ℓ = 1, 3, …, and an atomic state has odd parity if an odd number of electrons occupy these orbitals. For example, the ground state of the nitrogen atom has the electron configuration 1s22s22p3, and is identified by the term symbol 4So, where the superscript o denotes odd parity. However the third excited term at about 83,300 cm−1 above the ground state has electron configuration 1s22s22p23s has even parity since there are only two 2p electrons, and its term symbol is 4P.

Molecules

The complete electromagnetic Hamiltonian
of any molecule commutes with the parity operation
P and its eigenvalues can be given the parity symmetry
label + or - as they are even or odd, respectively. The parity operation involves the inversion
of electronic and nuclear spatial coordinates at the molecular center of mass.
Centrosymmetric molecules at equilibrium have a centre of symmetry at their midpoint
. This includes all homonuclear diatomic molecules
as well as certain symmetric molecules such as ethylene, benzene,
xenon tetrafluoride and
sulphur hexafluoride. For centrosymmetric molecules, the point
group contains the operation i which is not to be confused
with the parity operation. The operation i involves the
inversion of the electronic and vibrational displacement coordinates
at the nuclear centre of mass. For centrosymmetric molecules the operation i
commutes with the rovibronic Hamiltonian and can be used to
label such states.
Electronic and vibrational states of centrosymmetric molecules are either
unchanged by the operation i, or they are changed in sign by i.
The former are denoted by the
subscript g and are called gerade, while the latter
are denoted by the subscript u and are called ungerade. The complete Hamiltonian of a centrosymmetric molecule
does not commute with the point group inversion operation i because of the effect of the nuclear hyperfine Hamiltonian. The nuclear hyperfine Hamiltonian can mix the rotational levels of g and u vibronic states and give
rise to ortho-para transitions

Nuclei

In atomic nuclei, the state of each nucleon has even or odd parity, and nucleon configurations can be predicted using the nuclear shell model. As for electrons in atoms, the nucleon state has odd overall parity if and only if the number of nucleons in odd-parity states is odd. The parity is usually written as a + or − following the nuclear spin value. For example, the isotopes of oxygen include 17O, meaning that the spin is 5/2 and the parity is even. The shell model explains this because the first 16 nucleons are paired so that each pair has spin zero and even parity, and the last nucleon is in the 1d5/2 shell, which has even parity since ℓ = 2 for a d orbital.

Quantum field theory

If we can show that the vacuum state is invariant under parity,, the Hamiltonian is parity invariant and the quantization conditions remain unchanged under parity, then it follows that every state has good parity, and this parity is conserved in any reaction.
To show that quantum electrodynamics is invariant under parity, we have to prove that the action is invariant and the quantization is also invariant. For simplicity we will assume that canonical quantization is used; the vacuum state is then invariant under parity by construction. The invariance of the action follows from the classical invariance of Maxwell's equations. The invariance of the canonical quantization procedure can be worked out, and turns out to depend on the transformation of the annihilation operator:
where p denotes the momentum of a photon and ± refers to its polarization state. This is equivalent to the statement that the photon has odd intrinsic parity. Similarly all vector bosons can be shown to have odd intrinsic parity, and all axial-vectors to have even intrinsic parity.
A straightforward extension of these arguments to scalar field theories shows that scalars have even parity, since
This is true even for a complex scalar field.
With fermions, there is a slight complication because there is more than one spin group.

Parity in the standard model

Fixing the global symmetries

In the Standard Model of fundamental interactions there are precisely three global internal U symmetry groups available, with charges equal to the baryon number B, the lepton number L and the electric charge Q. The product of the parity operator with any combination of these rotations is another parity operator. It is conventional to choose one specific combination of these rotations to define a standard parity operator, and other parity operators are related to the standard one by internal rotations. One way to fix a standard parity operator is to assign the parities of three particles with linearly independent charges B, L and Q. In general, one assigns the parity of the most common massive particles, the proton, the neutron and the electron, to be +1.
Steven Weinberg has shown that if, where F is the fermion number operator, then, since the fermion number is the sum of the lepton number plus the baryon number,, for all particles in the Standard Model and since lepton number and baryon number are charges Q of continuous symmetries eiQ, it is possible to redefine the parity operator so that. However, if there exist Majorana neutrinos, which experimentalists today believe is possible, their fermion number is equal to one because they are neutrinos while their baryon and lepton numbers are zero because they are Majorana, and so F would not be embedded in a continuous symmetry group. Thus Majorana neutrinos would have parity ±i.

Parity of the pion

In 1954, a paper by William Chinowsky and Jack Steinberger demonstrated that the pion has negative parity. They studied the decay of an "atom" made from a deuteron and a negatively charged pion in a state with zero orbital angular momentum into two neutrons.
Neutrons are fermions and so obey Fermi–Dirac statistics, which implies that the final state is antisymmetric. Using the fact that the deuteron has spin one and the pion spin zero together with the antisymmetry of the final state they concluded that the two neutrons must have orbital angular momentum. The total parity is the product of the intrinsic parities of the particles and the extrinsic parity of the spherical harmonic function. Since the orbital momentum changes from zero to one in this process, if the process is to conserve the total parity then the products of the intrinsic parities of the initial and final particles must have opposite sign. A deuteron nucleus is made from a proton and a neutron, and so using the aforementioned convention that protons and neutrons have intrinsic parities equal to they argued that the parity of the pion is equal to minus the product of the parities of the two neutrons divided by that of the proton and neutron in the deuteron, explicitly. Thus they concluded that the pion is a pseudoscalar particle.

Parity violation

Although parity is conserved in electromagnetism, strong interactions and gravity, it is violated in weak interactions. The Standard Model incorporates parity violation by expressing the weak interaction as a chiral gauge interaction. Only the left-handed components of particles and right-handed components of antiparticles participate in weak interactions in the Standard Model. This implies that parity is not a symmetry of our universe, unless a hidden mirror sector exists in which parity is violated in the opposite way.
By the mid-20th century, it had been suggested by several scientists that parity might not be conserved, but without solid evidence these suggestions were not considered important. Then, in 1956, a careful review and analysis by theoretical physicists Tsung-Dao Lee and Chen-Ning Yang went further, showing that while parity conservation had been verified in decays by the strong or electromagnetic interactions, it was untested in the weak interaction. They proposed several possible direct experimental tests. They were mostly ignored, but Lee was able to convince his Columbia colleague Chien-Shiung Wu to try it. She needed special cryogenic facilities and expertise, so the experiment was done at the National Bureau of Standards.
In 1957 Wu, E. Ambler, R. W. Hayward, D. D. Hoppes, and R. P. Hudson found a clear violation of parity conservation in the beta decay of cobalt-60. As the experiment was winding down, with double-checking in progress, Wu informed Lee and Yang of their positive results, and saying the results need further examination, she asked them not to publicize the results first. However, Lee revealed the results to his Columbia colleagues on 4 January 1957 at a "Friday Lunch" gathering of the Physics Department of Columbia. Three of them, R. L. Garwin, Leon Lederman, and R. Weinrich modified an existing cyclotron experiment, and they immediately verified the parity violation. They delayed publication of their results until after Wu's group was ready, and the two papers appeared back-to-back in the same physics journal.
After the fact, it was noted that an obscure 1928 experiment, done by R. T. Cox, G. C. McIlwraith, and B. Kurrelmeyer, had in effect reported parity violation in weak decays, but since the appropriate concepts had not yet been developed, those results had no impact. The discovery of parity violation immediately explained the outstanding τ–θ puzzle in the physics of kaons.
In 2010, it was reported that physicists working with the Relativistic Heavy Ion Collider had created a short-lived parity symmetry-breaking bubble in quark-gluon plasmas. An experiment conducted by several physicists including Yale's Jack Sandweiss as part of the STAR collaboration, suggested that parity may also be violated in the strong interaction. It is predicted that this local parity violation, which would be analogous to the effect that is induced by fluctuation of the axion field, manifest itself by chiral magnetic effect.

Intrinsic parity of hadrons

To every particle one can assign an intrinsic parity as long as nature preserves parity. Although weak interactions do not, one can still assign a parity to any hadron by examining the strong interaction reaction that produces it, or through decays not involving the weak interaction, such as rho meson decay to pions.