Fermion doubling

Last updated

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.

Contents

Naive fermion discretization

For simplicity we will consider a four-dimensional theory of a free fermion, although the fermion doubling problem remains in arbitrary dimensions and even if interactions are included. Lattice field theory is usually carried out in Euclidean spacetime arrived at from Minkowski spacetime after a Wick rotation, where the continuum Dirac action takes the form

This is discretized by introducing a lattice with lattice spacing and points indexed by a vector of integers . The integral becomes a sum over all lattice points, while the fermionic fields are replaced by four-component Grassmann variables at each lattice site denoted by and . The derivative discretization used is the symmetric derivative discretization, with the vectors being unit vectors in the direction. These steps give the naive free fermion action [1]

This action reduces down to the continuum Dirac action in the continuum limit, so is expect to be a theory of a single fermion. However, it instead describes sixteen identical fermions, with each fermion said to have a different taste, analogously to how particles have different flavours in particle physics. The fifteen additional fermions are often referred to as doublers. This extended particle content can be seen by analyzing the symmetries or the correlation functions of the lattice theory.

Doubling symmetry

The naive fermion action possesses a new taste-exchange symmetry not found in the continuum theory acting on the fermion fields as [2]

where the vectors are the sixteen vectors with non-zero entries of specified by . For example, , , , and . The Dirac structure in the symmetry is similarly defined by the indices of as where and ; for example with .

The presence of these sixteen symmetry transformations implies the existence of sixteen identical fermion states rather than just one. Starting with a fermion field , the symmetry maps it to another field . Fourier transforming this shows that its momentum has been shifted as . Therefore, a fermion with momentum near the center of the Brillouin zone is mapped to one of its corners while one of the corner fermions comes in to replace the center fermion, showing that the transformation acts to exchange the tastes of the fermions. Since this is a symmetry of the action, the different tastes must be physically indistinguishable from each other. Here the Brillouin momentum for small is not the physical momentum of the particle, rather that is . Instead acts more as an additional quantum number specifying the taste of a fermion.

The term is responsible for changing the representation of the -matrices of the doublers to , which has the effect of changing the signs of the matrices as . Since any such sign change results in a set of matrices still satisfying the Dirac algebra, the resulting matrices form a valid representation. It is also the term that enters the wave function of the doublers given by and , where and are the usual Dirac equation solutions with momentum . [3]

Propagator and dispersion relation

In the continuum theory, the Dirac propagator has a single pole as the theory describes only a single particle. However, calculating the propagator from the naive action yields

for a fermion with momentum . [4] For low momenta this still has the expected pole at , but there are fifteen additional poles when . Each of these is a new fermion species with doubling arising because the function has two poles over the range . This is in contrast to what happens when particles of different spins are discretized. For example, scalars acquire propagators taking a similar form except with , which only has a single pole over the momentum range and so the theory does not suffer from a doubling problem. [5]

Comparison between the continuum and lattice dispersion relations of a free fermion, with the latter having a doubler at the Brillouin zone boundary. Fermion doubling dispersion relation.svg
Comparison between the continuum and lattice dispersion relations of a free fermion, with the latter having a doubler at the Brillouin zone boundary.

The necessity of fermion doubling can be deduced from the fact that the massless fermion propagator is odd around the origin. [6] That is, in the continuum limit it is proportional to , which must still be the case on the lattice in the small momentum limit. But since any local lattice theory that can be constructed must have a propagator that is continuous and periodic, it must cross the zero axis at least once more, which is exactly what occurs on the Brillouin zone corners where for the naive fermion propagator. This is in contrast to the bosonic propagator which is quadratic around the origin and so does not have such problem. Doubling can be avoided if a discontinuous propagator is used, but this results in a non-local theory.

The presence of doublers is also reflected in the fermion dispersion relation. Since this is a relation between the energy of the fermion and its momentum, it requires performing an inverse Wick transformation , with the dispersion relation arising from the pole of the propagator [7]

The zeros of this dispersion relation are local energy minima around which excitations correspond to different particle species. The above has eight different species arising due to doubling in the three spatial directions. The remaining eight doublers occur due to another doubling in the Euclidean temporal direction, which seems to have been lost. But this is due to a naive application of the inverse Wick transformation. The theory has an obstruction that does not allow for the simple replacement of and instead requires performing the full contour integration. Doing this for the position space propagator results in two separate terms, each of which has the same dispersion relation of eight fermion species, giving a total of sixteen. [8] The obstruction between the Minkowski and Euclidean naive fermion lattice theories occurs because doubling does not occur in the Minkowski temporal direction, so the two theories differ in their particle content.

Resolutions to fermion doubling

Fermion doubling is a consequence of a no-go theorem in lattice field theory known as the Nielsen–Ninomiya theorem. It states that any even dimensional local, hermitian, translationally invariant, bilinear fermionic theory always has the same number of left-handed and right-handed Weyl fermions, generating the additional fermions when they are lacking. [9] The theorem does not say how many doublers will arise, but without breaking the assumptions of the theorem, there will always be at least one doubler, with the naive discretization having fifteen. A consequence of the theorem is that the chiral anomaly cannot be simulated with chirally invariant theories as it trivially vanishes.

Simulating lattice field theories with fermion doubling leads to incorrect results due to the doublers, so many strategies to overcome this problem have been developed. While doublers can be ignored in a free theory as there the different tastes decouple, they cannot be ignored in an interacting theory where interactions mix different tastes, since momentum is conserved only up to modulo . For example, two taste fermions can scatter by the exchange of a highly virtual gauge boson to produce two taste fermions without violating momentum conservation. Therefore, to overcome the fermion doubling problem, one must violate one or more assumptions of the Nielsen–Ninomiya theorem, giving rise to a multitude of proposed resolutions:

These fermion formulations each have their own advantages and disadvantages. [26] They differ in the speed at which they can be simulated, the easy of their implementation, and the presence or absence of exceptional configurations. Some of them have a residual chiral symmetry allowing one to simulate axial anomalies. They can also differ in how many of the doublers they eliminate, with some consisting of a doublet, or a quartet of fermions. For this reason different fermion formulations are used for different problems.

Derivative discretization

Another possible although impractical solution to the doubling problem is to adopt a derivative discretization different from the symmetric difference

used in the naive fermion action. Instead it is possible to use the forward difference

or a backward difference discretizations. The effect of the derivative discretizations on doubling is seen by considering the one-dimensional toy problem of finding the eigensolutions of . [27] In the continuum this differential equation has a single solution. However, implementing the symmetric difference derivative leads to the presence of two distinct eigensolutions, while a forward or backward difference derivative has one eigensolution. This effect carries forward to the fermion action where fermion doubling is absent with forward or backward discretizations.

The reason for this particle content disparity is that the symmetric difference derivative maintains the hermiticity property of the continuum operator, while the forward and backward discretizations do not. These latter discretizations lead to non-hermitian actions, breaking the assumptions of the Nielsen–Ninomiya theorem, and so avoid the fermion doubling problem. Developing an interacting theory with a non-hermitian derivative discretization leads to a theory with non-covariant contributions to the fermion self-energy and vertex function, rendering the theory non-renormalizable and difficult to work with. [28] For this reason such a resolution to the fermion doubling problem is generally not implemented.

See also

Related Research Articles

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.

The strong CP problem is a question in particle physics, which brings up the following quandary: why does quantum chromodynamics (QCD) seem to preserve CP-symmetry?

<span class="mw-page-title-main">Wheeler–DeWitt equation</span> Field equation, part of a theory that attempts to combine quantum mechanics and general relativity

The Wheeler–DeWitt equation for theoretical physics and applied mathematics, is a field equation attributed to John Archibald Wheeler and Bryce DeWitt. The equation attempts to mathematically combine the ideas of quantum mechanics and general relativity, a step towards a theory of quantum gravity.

In particle physics, Yukawa's interaction or Yukawa coupling, named after Hideki Yukawa, is an interaction between particles according to the Yukawa potential. Specifically, it is a scalar field ϕ and a Dirac field ψ of the type

In theoretical physics, the Rarita–Schwinger equation is the relativistic field equation of spin-3/2 fermions in a four-dimensional flat spacetime. 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.

The QCD vacuum is the quantum vacuum state of quantum chromodynamics (QCD). It is an example of a non-perturbative vacuum state, characterized by non-vanishing condensates such as the gluon condensate and the quark condensate in the complete theory which includes quarks. The presence of these condensates characterizes the confined phase of quark matter.

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 lattice field theory, staggered fermions are a fermion discretization that reduces the number of fermion doublers from sixteen to four. They are one of the fastest lattice fermions when it comes to simulations and they also possess some nice features such as a remnant chiral symmetry, making them very popular in lattice QCD calculations. Staggered fermions were first formulated by John Kogut and Leonard Susskind in 1975 and were later found to be equivalent to the discretized version of the Dirac–Kähler fermion.

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 field theory, a Ward–Takahashi identity is an identity between correlation functions that follows from the global or gauge symmetries of the theory, and which remains valid after renormalization.

In lattice field theory, the Nielsen–Ninomiya theorem is a no-go theorem about placing chiral fermions on a lattice. In particular, under very general assumptions such as locality, hermiticity, and translational symmetry, any lattice formulation of chiral fermions necessarily leads to fermion doubling, where there are the same number of left-handed and right-handed fermions. It was first proved by Holger Bech Nielsen and Masao Ninomiya in 1981 using two methods, one that relied on homotopy theory and another that relied on differential topology. Another proof provided by Daniel Friedan uses differential geometry. The theorem was also generalized to any regularization scheme of chiral theories. One consequence of the theorem is that the Standard Model cannot be put on a lattice. Common methods for overcoming the fermion doubling problem is to use modified fermion formulations such as staggered fermions, Wilson fermions, or Ginsparg–Wilson fermions, among others.

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.

<span class="mw-page-title-main">Nonlinear Dirac equation</span> Dirac equation for self-interacting fermions

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.

<span class="mw-page-title-main">Light-front computational methods</span> Technique in computational quantum field theory

The light-front quantization of quantum field theories provides a useful alternative to ordinary equal-time quantization. In particular, it can lead to a relativistic description of bound systems in terms of quantum-mechanical wave functions. The quantization is based on the choice of light-front coordinates, where plays the role of time and the corresponding spatial coordinate is . Here, is the ordinary time, is one Cartesian coordinate, and is the speed of light. The other two Cartesian coordinates, and , are untouched and often called transverse or perpendicular, denoted by symbols of the type . The choice of the frame of reference where the time and -axis are defined can be left unspecified in an exactly soluble relativistic theory, but in practical calculations some choices may be more suitable than others.

In string theory, the Ramond–Neveu–Schwarz (RNS) formalism is an approach to formulating superstrings in which the worldsheet has explicit superconformal invariance but spacetime supersymmetry is hidden, in contrast to the Green–Schwarz formalism where the latter is explicit. It was originally developed by Pierre Ramond, André Neveu and John Schwarz in the RNS model in 1971, which gives rise to type II string theories and can also give type I string theory. Heterotic string theories can also be acquired through this formalism by using a different worldsheet action. There are various ways to quantize the string within this framework including light-cone quantization, old canonical quantization, and BRST quantization. A consistent string theory is only acquired if the spectrum of states is restricted through a procedure known as a GSO projection, with this projection being automatically present in the Green–Schwarz formalism.

In lattice field theory, Wilson fermions are a fermion discretization that allows to avoid the fermion doubling problem proposed by Kenneth Wilson in 1974. They are widely used, for instance in lattice QCD calculations.

In lattice field theory, the Ginsparg–Wilson equation generalizes chiral symmetry on the lattice in a way that approaches the continuum formulation in the continuum limit. The class of fermions whose Dirac operators satisfy this equation are known as Ginsparg–Wilson fermions, with notable examples being overlap, domain wall and fixed point fermions. They are a means to avoid the fermion doubling problem, widely used for instance in lattice QCD calculations. The equation was discovered by Paul Ginsparg and Kenneth Wilson in 1982, however it was quickly forgotten about since there were no known solutions. It was only in 1997 and 1998 that the first solutions were found in the form of the overlap and fixed point fermions, at which point the equation entered prominence.

The axial current, also denoted the pseudo-vector or chiral current, is the conserved current associated to the chiral symmetry or axial symmetry of a system.

In supersymmetry, 4D supergravity is the theory of supergravity in four dimensions with a single supercharge. It contains exactly one supergravity multiplet, consisting of a graviton and a gravitino, but can also have an arbitrary number of chiral and vector supermultiplets, with supersymmetry imposing stringent constraints on how these can interact. The theory is primarily determined by three functions, those being the Kähler potential, the superpotential, and the gauge kinetic matrix. Many of its properties are strongly linked to the geometry associated to the scalar fields in the chiral multiplets. After the simplest form of this supergravity was first discovered, a theory involving only the supergravity multiplet, the following years saw an effort to incorporate different matter multiplets, with the general action being derived in 1982 by Eugène Cremmer, Sergio Ferrara, Luciano Girardello, and Antonie Van Proeyen.

References

  1. Gupta, R. (1997). "Introduction to lattice QCD: Course". Les Houches Summer School in Theoretical Physics, Session 68: Probing the Standard Model of Particle Interactions: 83–219. arXiv: hep-lat/9807028 .
  2. Follana, E.; et al. (HPQCD,UKQCD) (2007). "Highly Improved Staggered Quarks on the Lattice, with Applications to Charm Physics". Phys. Rev. D. 75 (5): 054502. arXiv: hep-lat/0610092 . Bibcode:2007PhRvD..75e4502F. doi:10.1103/PhysRevD.75.054502. S2CID   119506250.
  3. DeGrand, T.; DeTar, C. (2006). "6". Lattice Methods for Quantum Chromodynamics. World Scientific Publishing. p. 103. Bibcode:2006lmqc.book.....D. doi:10.1142/6065. ISBN   978-9812567277.
  4. Gattringer, C.; Lang, C.B. (2009). "5". Quantum Chromodynamics on the Lattice: An Introductory Presentation. Lecture Notes in Physics 788. Springer. pp. 111–112. doi:10.1007/978-3-642-01850-3. ISBN   978-3642018497.
  5. Rothe, H.J. (2005). "3". Lattice Gauge Theories: An Introduction. World Scientific Lecture Notes in Physics: Volume 43. Vol. 82. World Scientific Publishing. pp. 39–40. doi:10.1142/8229. hdl:20.500.12657/50492. ISBN   978-9814365857.
  6. Makeenko, Y. (2002). "8". Methods of Contemporary Gauge Theory. Cambridge Monographs on Mathematical Physics. Cambridge: Cambridge University Press. pp. 149–150. doi:10.1017/CBO9780511535147. hdl:20.500.12657/64018. ISBN   978-0521809115.
  7. Montvay, I.; Munster, G. (1994). "4". Quantum Fields on a Lattice. Cambridge Monographs on Mathematical Physics. Cambridge: Cambridge University Press. pp. 116–117. doi:10.1017/CBO9780511470783. ISBN   9780511470783. S2CID   118339104.
  8. Smit, Jan (2002). "6". Introduction to Quantum Fields on a Lattice. Cambridge Lecture Notes in Physics. Cambridge: Cambridge University Press. pp. 153–156. doi:10.1017/CBO9780511583971. hdl:20.500.12657/64022. ISBN   9780511583971.
  9. Nielsen, H.B.; Ninomiya, M. (1981). "A no-go theorem for regularizing chiral fermions". Physics Letters B. 105 (2): 219–223. Bibcode:1981PhLB..105..219N. doi:10.1016/0370-2693(81)91026-1.
  10. Kaplan, D.B. (1992). "A Method for simulating chiral fermions on the lattice". Phys. Lett. B. 288 (3–4): 342–347. arXiv: hep-lat/9206013 . Bibcode:1992PhLB..288..342K. doi:10.1016/0370-2693(92)91112-M. S2CID   14161004.
  11. Shamir, Y. (1993). "Chiral fermions from lattice boundaries". Nucl. Phys. B. 406 (1–2): 90–106. arXiv: hep-lat/9303005 . Bibcode:1993NuPhB.406...90S. doi:10.1016/0550-3213(93)90162-I. S2CID   16187316.
  12. Wilson, K.G.; Ginsparg, P.H. (1982). "A remnant of chiral symmetry on the lattice". Phys. Rev. D. 25 (10): 2649–2657. Bibcode:1982PhRvD..25.2649G. doi:10.1103/PhysRevD.25.2649.
  13. Neuberger, H. (1998). "Exactly massless quarks on the lattice". Phys. Lett. B. 417 (1–2): 141–144. arXiv: hep-lat/9707022 . Bibcode:1998PhLB..417..141N. doi:10.1016/S0370-2693(97)01368-3. S2CID   119372020.
  14. Neuberger, H. (1998). "More about exactly massless quarks on the lattice". Phys. Lett. B. 427 (3–4): 353–355. arXiv: hep-lat/9801031 . Bibcode:1998PhLB..427..353N. doi:10.1016/S0370-2693(98)00355-4. S2CID   17397528.
  15. Wolfgang, B.; Wiese, U.J. (1996). "Perfect lattice actions for quarks and gluons". Nucl. Phys. B. 464 (1–2): 319–352. arXiv: hep-lat/9510026 . Bibcode:1996NuPhB.464..319B. doi:10.1016/0550-3213(95)00678-8. S2CID   119378690.
  16. Drell, S.D.; Weinstein, M.; Yankielowicz, S. (1976). "Strong-coupling field theories. II. Fermions and gauge fields on a lattice" (PDF). Phys. Rev. D. 14 (6): 1627–1647. Bibcode:1976PhRvD..14.1627D. doi:10.1103/PhysRevD.14.1627. OSTI   1446145.
  17. Stacey, Richard (1982). "Eliminating lattice fermion doubling". Phys. Rev. D. 26 (2). American Physical Society: 468–472. Bibcode:1982PhRvD..26..468S. doi:10.1103/PhysRevD.26.468.
  18. Kogut, J.; Susskind, L. (1975). "Hamiltonian formulation of Wilson's lattice gauge theories". Phys. Rev. D. 11 (2): 395–408. Bibcode:1975PhRvD..11..395K. doi:10.1103/PhysRevD.11.395.
  19. Tong, D. (2022). "Comments on symmetric mass generation in 2d and 4d". JHEP. 07 (7): 001. arXiv: 2104.03997 . Bibcode:2022JHEP...07..001T. doi:10.1007/JHEP07(2022)001.
  20. Wang, Juven; You, Yi-Zhuang (2022). "Symmetric Mass Generation". Symmetry. 14 (7): 1475. arXiv: 2204.14271 . Bibcode:2022Symm...14.1475W. doi: 10.3390/sym14071475 .
  21. Eichten, E.; Preskill, J. (1986). "Chiral gauge theories on the lattice". Nuclear Physics B. 268 (1): 179–208. Bibcode:1986NuPhB.268..179E. doi:10.1016/0550-3213(86)90207-5.
  22. Wang, J.; Wen, X. (2023). "Nonperturbative regularization of (1+1)-dimensional anomaly-free chiral fermions and bosons: On the equivalence of anomaly matching conditions and boundary gapping rules". Phys. Rev. B. 107 (1): 014311. arXiv: 1307.7480 . Bibcode:2023PhRvB.107a4311W. doi:10.1103/PhysRevB.107.014311.
  23. Zeng, M.; Zhu, Z.; Wang, J.; You, Y. (2022). "Symmetric Mass Generation in the 1+1 Dimensional Chiral Fermion 3-4-5-0 Model". Phys. Rev. Lett. 128 (18): 185301. arXiv: 2202.12355 . Bibcode:2022PhRvL.128r5301Z. doi:10.1103/PhysRevLett.128.185301.
  24. Frezzotti et al. (Alpha Collaboration), R. (2001). "Lattice QCD with a chirally twisted mass term". JHEP. 08 (8): 058. arXiv: hep-lat/0101001 . Bibcode:2001JHEP...08..058A. doi:10.1088/1126-6708/2001/08/058. S2CID   14746313.
  25. Wilson, K.G. (1974). "Confinement of quarks". Phys. Rev. D. 10 (8): 2445–2459. Bibcode:1974PhRvD..10.2445W. doi:10.1103/PhysRevD.10.2445.
  26. Tanabashi et al.(Particle Data Group), T. (2018). "17. Lattice Quantum Chromodynamics" (PDF). Phys. Rev. D. 98 (3). American Physical Society: 4–6. Bibcode:2018PhRvD..98c0001T. doi:10.1103/PhysRevD.98.030001.
  27. Rothe, H.J. (2005). "4". Lattice Gauge Theories: An Introduction. World Scientific Lecture Notes in Physics: Volume 43. Vol. 82. World Scientific Publishing. pp. 48–55. doi:10.1142/8229. hdl:20.500.12657/50492. ISBN   978-9814365857.
  28. Sadooghi, N.; Rothe, H.J. (1997). "Continuum behavior of lattice QED, discretized with one sided lattice differences, in one loop order". Phys. Rev. D. 55 (11): 6749–6759. arXiv: hep-lat/9610001 . Bibcode:1997PhRvD..55.6749S. doi:10.1103/PhysRevD.55.6749. S2CID   16624237.