Nonlinear Dirac equation

Last updated
See Ricci calculus and Van der Waerden notation for the notation.

In quantum field theory, the nonlinear Dirac equation is a model of self-interacting Dirac fermions. This model is widely considered in quantum physics as a toy model of self-interacting electrons. [1] [2] [3] [4] [5]

Contents

The nonlinear Dirac equation appears in the Einstein–Cartan–Sciama–Kibble theory of gravity, which extends general relativity to matter with intrinsic angular momentum (spin). [6] [7] This theory removes a constraint of the symmetry of the affine connection and treats its antisymmetric part, the torsion tensor, as a variable in varying the action. In the resulting field equations, the torsion tensor is a homogeneous, linear function of the spin tensor. The minimal coupling between torsion and Dirac spinors thus generates an axial-axial, spin–spin interaction in fermionic matter, which becomes significant only at extremely high densities. Consequently, the Dirac equation becomes nonlinear (cubic) in the spinor field, [8] [9] which causes fermions to be spatially extended and may remove the ultraviolet divergence in quantum field theory. [10]

Models

Two common examples are the massive Thirring model and the Soler model.

Thirring model

The Thirring model [11] was originally formulated as a model in (1 + 1) space-time dimensions and is characterized by the Lagrangian density

where ψC2 is the spinor field, ψ = ψ*γ0 is the Dirac adjoint spinor,

(Feynman slash notation is used), g is the coupling constant, m is the mass, and γμ are the two-dimensional gamma matrices , finally μ = 0, 1 is an index.

Soler model

The Soler model [12] was originally formulated in (3 + 1) space-time dimensions. It is characterized by the Lagrangian density

using the same notations above, except

is now the four-gradient operator contracted with the four-dimensional Dirac gamma matrices γμ, so therein μ = 0, 1, 2, 3.

Other models

Besides the Soler model, extensive work has been done where nonlinear versions of Dirac’s equation are used to describe purely classical, nonlinear particle-like solutions (PLS) in (3 + 1) space-time dimensions. Rañada has given a review of the subject. [13] Although a more recent review specifically devoted to purely classical, nonlinear PLS has apparently not appeared, pertinent references are available in various more recent publications. [14] [15]

The models reviewed by Rañada are meant to be entirely classical in nature and should properly be regarded as having nothing to do with quantum mechanics, but the dependent variable in the Dirac equation is still typically taken as a spinor. When a purely classical model of this nature is to be considered, the use of a spinor as the dependent variable seems inappropriate.

If a minor modification of the underlying Dirac equation is used, the problem can be avoided in a relatively straightforward way. [16] Instead of using the usual column vector as the dependent variable in Dirac’s equation, one can use a 4 × 4 matrix. When there is no transformation of coordinates, the leftmost column of the matrix is used in Dirac’s equation in the usual manner, but when there is to be a transformation in space-time, the four components of the dependent variable are sometimes allowed to appear in various different positions in the 4 × 4 matrix.

The result can be understood in terms of a Clifford algebra since the dependent variable in Dirac’s equation can be represented as a 4 dimensional left ideal of a Clifford algebra. In this case one simply allows the dependent variable to lie in a different left ideal when there is a transformation in space-time.

Einstein–Cartan theory

In Einstein–Cartan theory the Lagrangian density for a Dirac spinor field is given by ()

where

is the Fock–Ivanenko covariant derivative of a spinor with respect to the affine connection, is the spin connection, is the determinant of the metric tensor , and the Dirac matrices satisfy

The Einstein–Cartan field equations for the spin connection yield an algebraic constraint between the spin connection and the spinor field rather than a partial differential equation, which allows the spin connection to be explicitly eliminated from the theory. The final result is a nonlinear Dirac equation containing an effective "spin-spin" self-interaction,

where is the general-relativistic covariant derivative of a spinor, and is the Einstein gravitational constant, . The cubic term in this equation becomes significant at densities on the order of .

See also

Related Research Articles

<span class="mw-page-title-main">Quantum electrodynamics</span> Quantum field theory of electromagnetism

In particle physics, quantum electrodynamics (QED) is the relativistic quantum field theory of electrodynamics. In essence, it describes how light and matter interact and is the first theory where full agreement between quantum mechanics and special relativity is achieved. QED mathematically describes all phenomena involving electrically charged particles interacting by means of exchange of photons and represents the quantum counterpart of classical electromagnetism giving a complete account of matter and light interaction.

In particle physics, the Dirac equation is a relativistic wave equation derived by British physicist Paul Dirac in 1928. In its free form, or including electromagnetic interactions, it describes all spin-12 massive particles, called "Dirac particles", such as electrons and quarks for which parity is a symmetry. It is consistent with both the principles of quantum mechanics and the theory of special relativity, and was the first theory to account fully for special relativity in the context of quantum mechanics. It was validated by accounting for the fine structure of the hydrogen spectrum in a completely rigorous way.

In physics, the Schwinger model, named after Julian Schwinger, is the model describing 1+1D Lorentzian quantum electrodynamics which includes electrons, coupled to photons.

In theoretical physics, a chiral anomaly is the anomalous nonconservation of a chiral current. In everyday terms, it is equivalent to a sealed box that contained equal numbers of left and right-handed bolts, but when opened was found to have more left than right, or vice versa.

<span class="mw-page-title-main">Relativistic wave equations</span> Wave equations respecting special and general relativity

In physics, specifically relativistic quantum mechanics (RQM) and its applications to particle physics, relativistic wave equations predict the behavior of particles at high energies and velocities comparable to the speed of light. In the context of quantum field theory (QFT), the equations determine the dynamics of quantum fields. The solutions to the equations, universally denoted as ψ or Ψ, are referred to as "wave functions" in the context of RQM, and "fields" in the context of QFT. The equations themselves are called "wave equations" or "field equations", because they have the mathematical form of a wave equation or are generated from a Lagrangian density and the field-theoretic Euler–Lagrange equations.

In theoretical physics, the Rarita–Schwinger equation is the relativistic field equation of spin-3/2 fermions. It is similar to the Dirac equation for spin-1/2 fermions. This equation was first introduced by William Rarita and Julian Schwinger in 1941.

In lattice field theory, fermion doubling occurs when naively putting fermionic fields on a lattice, resulting in more fermionic states than expected. For the naively discretized Dirac fermions in Euclidean dimensions, each fermionic field results in identical fermion species, referred to as different tastes of the fermion. The fermion doubling problem is intractably linked to chiral invariance by the Nielsen–Ninomiya theorem. Most strategies used to solve the problem require using modified fermions which reduce to the Dirac fermion only in the continuum limit.

In atomic physics, the electron magnetic moment, or more specifically the electron magnetic dipole moment, is the magnetic moment of an electron resulting from its intrinsic properties of spin and electric charge. The value of the electron magnetic moment is −9.2847647043(28)×10−24 J⋅T−1. In units of the Bohr magneton (μB), it is −1.00115965218059(13) μB, a value that was measured with a relative accuracy of 1.3×10−13.

<span class="mw-page-title-main">Mathematical formulation of the Standard Model</span> Mathematics of a particle physics model

This article describes the mathematics of the Standard Model of particle physics, a gauge quantum field theory containing the internal symmetries of the unitary product group SU(3) × SU(2) × U(1). The theory is commonly viewed as describing the fundamental set of particles – the leptons, quarks, gauge bosons and the Higgs boson.

In quantum field theory, a fermionic field is a quantum field whose quanta are fermions; that is, they obey Fermi–Dirac statistics. Fermionic fields obey canonical anticommutation relations rather than the canonical commutation relations of bosonic fields.

In differential geometry and mathematical physics, a spin connection is a connection on a spinor bundle. It is induced, in a canonical manner, from the affine connection. It can also be regarded as the gauge field generated by local Lorentz transformations. In some canonical formulations of general relativity, a spin connection is defined on spatial slices and can also be regarded as the gauge field generated by local rotations.

The Thirring model is an exactly solvable quantum field theory which describes the self-interactions of a Dirac field in (1+1) dimensions.

In quantum mechanics, the Pauli equation or Schrödinger–Pauli equation is the formulation of the Schrödinger equation for spin-½ particles, which takes into account the interaction of the particle's spin with an external electromagnetic field. It is the non-relativistic limit of the Dirac equation and can be used where particles are moving at speeds much less than the speed of light, so that relativistic effects can be neglected. It was formulated by Wolfgang Pauli in 1927.

In mathematical physics, spacetime algebra (STA) is a name for the Clifford algebra Cl1,3(R), or equivalently the geometric algebra G(M4). According to David Hestenes, spacetime algebra can be particularly closely associated with the geometry of special relativity and relativistic spacetime.

In physics, relativistic quantum mechanics (RQM) is any Poincaré covariant formulation of quantum mechanics (QM). This theory is applicable to massive particles propagating at all velocities up to those comparable to the speed of light c, and can accommodate massless particles. The theory has application in high energy physics, particle physics and accelerator physics, as well as atomic physics, chemistry and condensed matter physics. Non-relativistic quantum mechanics refers to the mathematical formulation of quantum mechanics applied in the context of Galilean relativity, more specifically quantizing the equations of classical mechanics by replacing dynamical variables by operators. Relativistic quantum mechanics (RQM) is quantum mechanics applied with special relativity. Although the earlier formulations, like the Schrödinger picture and Heisenberg picture were originally formulated in a non-relativistic background, a few of them also work with special relativity.

The Thirring–Wess model or Vector Meson model is an exactly solvable quantum field theory, describing the interaction of a Dirac field with a vector field in dimension two.

<span class="mw-page-title-main">Two-body Dirac equations</span> Quantum field theory equations

In quantum field theory, and in the significant subfields of quantum electrodynamics (QED) and quantum chromodynamics (QCD), the two-body Dirac equations (TBDE) of constraint dynamics provide a three-dimensional yet manifestly covariant reformulation of the Bethe–Salpeter equation for two spin-1/2 particles. Such a reformulation is necessary since without it, as shown by Nakanishi, the Bethe–Salpeter equation possesses negative-norm solutions arising from the presence of an essentially relativistic degree of freedom, the relative time. These "ghost" states have spoiled the naive interpretation of the Bethe–Salpeter equation as a quantum mechanical wave equation. The two-body Dirac equations of constraint dynamics rectify this flaw. The forms of these equations can not only be derived from quantum field theory they can also be derived purely in the context of Dirac's constraint dynamics and relativistic mechanics and quantum mechanics. Their structures, unlike the more familiar two-body Dirac equation of Breit, which is a single equation, are that of two simultaneous quantum relativistic wave equations. A single two-body Dirac equation similar to the Breit equation can be derived from the TBDE. Unlike the Breit equation, it is manifestly covariant and free from the types of singularities that prevent a strictly nonperturbative treatment of the Breit equation.

<span class="mw-page-title-main">Bargmann–Wigner equations</span> Wave equation for arbitrary spin particles

In relativistic quantum mechanics and quantum field theory, the Bargmann–Wigner equations describe free particles with non-zero mass and arbitrary spin j, an integer for bosons or half-integer for fermions. The solutions to the equations are wavefunctions, mathematically in the form of multi-component spinor fields.

The soler model is a quantum field theory model of Dirac fermions interacting via four fermion interactions in 3 spatial and 1 time dimension. It was introduced in 1938 by Dmitri Ivanenko and re-introduced and investigated in 1970 by Mario Soler as a toy model of self-interacting electron.

In mathematical physics, the Gordon decomposition of the Dirac current is a splitting of the charge or particle-number current into a part that arises from the motion of the center of mass of the particles and a part that arises from gradients of the spin density. It makes explicit use of the Dirac equation and so it applies only to "on-shell" solutions of the Dirac equation.

References

  1. Д.Д. Иваненко (1938). "Замечание к теории взаимодействия через частицы" [translated in: D.D. Ivanenko, Notes to the theory of interaction via particles, Sov. Phys. JETP 13 (1938), 141)](PDF). ЖЭТФ. 8: 260–266.
  2. R. Finkelstein; R. LeLevier & M. Ruderman (1951). "Nonlinear spinor fields". Phys. Rev. 83 (2): 326–332. Bibcode:1951PhRv...83..326F. doi:10.1103/PhysRev.83.326.
  3. R. Finkelstein; C. Fronsdal & P. Kaus (1956). "Nonlinear Spinor Field". Phys. Rev. 103 (5): 1571–1579. Bibcode:1956PhRv..103.1571F. doi:10.1103/PhysRev.103.1571.
  4. W. Heisenberg (1957). "Quantum Theory of Fields and Elementary Particles". Rev. Mod. Phys. 29 (3): 269–278. Bibcode:1957RvMP...29..269H. doi:10.1103/RevModPhys.29.269.
  5. Gross, David J. and Neveu, André (1974). "Dynamical symmetry breaking in asymptotically free field theories". Phys. Rev. D. 10 (10): 3235–3253. Bibcode:1974PhRvD..10.3235G. doi:10.1103/PhysRevD.10.3235.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  6. Dennis W. Sciama, "The physical structure of general relativity". Rev. Mod. Phys. 36, 463-469 (1964).
  7. Tom W. B. Kibble, "Lorentz invariance and the gravitational field". J. Math. Phys. 2, 212-221 (1961).
  8. F. W. Hehl & B. K. Datta (1971). "Nonlinear spinor equation and asymmetric connection in general relativity". J. Math. Phys. 12 (7): 1334–1339. Bibcode:1971JMP....12.1334H. doi:10.1063/1.1665738.
  9. Friedrich W. Hehl; Paul von der Heyde; G. David Kerlick & James M. Nester (1976). "General relativity with spin and torsion: Foundations and prospects". Rev. Mod. Phys. 48 (3): 393–416. Bibcode:1976RvMP...48..393H. doi:10.1103/RevModPhys.48.393.
  10. Nikodem J. Popławski (2010). "Nonsingular Dirac particles in spacetime with torsion". Phys. Lett. B. 690 (1): 73–77. arXiv: 0910.1181 . Bibcode:2010PhLB..690...73P. doi:10.1016/j.physletb.2010.04.073.
  11. Walter Thirring (1958). "A soluble relativistic field theory". Annals of Physics. 3 (1): 91–112. Bibcode:1958AnPhy...3...91T. doi:10.1016/0003-4916(58)90015-0.
  12. Mario Soler (1970). "Classical, Stable, Nonlinear Spinor Field with Positive Rest Energy". Phys. Rev. D. 1 (10): 2766–2769. Bibcode:1970PhRvD...1.2766S. doi:10.1103/PhysRevD.1.2766.
  13. Rañada, A.F. (1983). "Classical nonlinear Dirac field models of extended particles". Quantum Theory, Groups, Fields and Particles. Amsterdam: Reidel. pp. 271–291. ISBN   1-4020-0324-2.
  14. Esteban, M.J.; Lewin, M.; Séré, E. (2008). "Variational methods in relativistic quantum mechanics". Bull. Amer. Math. Soc. (N.S.). 45 (4): 535–593. doi: 10.1090/S0273-0979-08-01212-3 . S2CID   1686536.
  15. Dudnikova, T.V. (2020). "Virial identities and energy-momentum relation for solitary waves of nonlinear Dirac equations". Lobachevskii Journal of Mathematics. 41 (6): 956–981. doi:10.1134/S1995080220060074. S2CID   226199105.
  16. Luke, J.C. (May 2022). "A Note on the Use of the Nonlinear Dirac Equation for Soliton Particle Models". ScholarWorks.IUPUI.edu. Retrieved June 2, 2022.