Skip to main content
Log in

Null-plane causal perturbation theory

  • Regular Article
  • Published:
The European Physical Journal Plus Aims and scope Submit manuscript

Abstract

We reconstruct causal perturbation theory in a way compatible with light-front dynamics on the basis of Rohrlich’s invariant null-plane coordinatization. We show that, when the causality axiom is referred to the \(x^+\) time coordinate, the retarded and advanced distributions are possibly non-null on the entire \(x^-\)-axis, despite which Bogoliubov–Medvedev–Polivanov’s axioms still have a perturbative solution whose practical implementation is exposed in detail. Our results are then applied to the construction of the free fields Feynman’s propagators, for which instantaneous terms arise in the fermion and massless vector cases, and then to Yukawa’s model for the interaction of a neutral pseudo-scalar field with nucleons.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. This is to say, that every point of the surface can be mapped into any other point of it by some element of the stability group.

  2. There is also a difference in the method of quantization of fields: In Ref. [98] it is replaced by the analytic representation of the propagators, which in the case of the massless vector field uses the Lagrangian approach by the introduction of Lagrange’s multipliers. In our work we quantize the fields by directly constructing Fock’s space; a résumé of this quantization method is exposed in Appendix A. Particularly, the photons space is constructed with the gauge-fixed wave-functions, which will lead to a difference in the radiation field commutation distribution.

  3. So-called Cauchy–Kovalevskaya’s problem for initial data on non-characteristic surfaces, and Goursat’s problem on characteristic ones [101].

  4. Certainly, the initial data must be given in the two characteristic surfaces, but, as is explained in Ref. [6], one of them can be replaced by an asymptotic condition of vanishing at infinity.

  5. In the conception of Epstein and Glaser [58],in fact, the switching function “is” the coupling constant of the theory.

  6. Strictly speaking, this axiom concerns unitarity over Fock’s space of physical states; when it is expanded in order to contain non-physical states, then it must be replaced by a pseudo-unitarity axiom—this occurs in the construction of gauge theories, as will be explained in a future work.

  7. The unnecessarity of Lorentz’s covariance in instant dynamics was affirmed by Epstein and Glaser in Ref. [58], in spite of which they themselves and also Scharf in Ref. [64], used it to simplify the proofs.

  8. The notion of a Schwartz’s multi-index is defined, for example, in Ref. [106]: A multi-index \(k\in {\mathbb {R}}^N\) is a sequence of non-negative numbers, \(k=(k_1;\cdots ;k_N)\), \(k_j\ge 0\), for which the following notations are established:

    $$\begin{aligned} |k|\equiv \sum \limits _{j=1}^N k_j,\quad x^k\equiv \prod \limits _{j=1}^Nx_j^{k_j},\quad k!\equiv \prod \limits _{j=1}^Nk_j!,\quad D^kf(x)\equiv \prod \limits _{j=1}^N\partial _{x_j}^{k_j}f(x). \end{aligned}$$
  9. It must be phenomenologically constructed in the same way as a Lagrangian density: By imposing the symmetries of the theory, but it must be restricted to the first order in the coupling constant terms, only. For example, the construction based on quantum gauge invariance can be found in Ref. [75].

  10. Our convention for Fourier’s transform of the function \(f\in {\mathscr {S}}\left( {\mathbb {R}}^4\right) \) is the following:

    $$\begin{aligned} {{\hat{f}}}(p)=(2\pi )^{-2}\int d^4x f(x)e^{ipx},\quad px=p_+x^++p_\perp x^\perp +p_-x^-\,. \end{aligned}$$

    We conventionally use lower indices for momentum variables; according to the metrics given in Eq. (3) we have the equivalence: \(p_+=p^-\), \(p_\perp =-p^\perp \) and \(p_-=p^+\). Additionally, we will denote: \(d^3\varvec{p}:=d^2p_\perp dp_-\).

  11. The definition of the singular order at the origin is similar to that at the \(x^-\)-axis, except that all the four variables are scaled. In momentum space [64]: Let \({{\hat{d}}}\in {\mathscr {S}}'({\mathbb {R}}^m)\) be a distribution, and let \(\rho \) be a continuous positive function. If the limit

    $$\begin{aligned} \lim _{s\rightarrow 0^+}\rho (s)\left\langle {{\hat{d}}}\left( \frac{p}{s}\right) ;{\check{\varphi }}(p)\right\rangle =\left\langle {{\hat{d}}}_0;{\check{\varphi }}\right\rangle \end{aligned}$$

    exists for all test function \(\check{\varphi }\in {\mathscr {S}}({\mathbb {R}}^m)\), then \({{\hat{d}}}_0\) is the quasi-asymptotics of the distribution \({{\hat{d}}}\) at \(p\rightarrow +\infty \), with regard to the function \(\rho \). \(\square \)

  12. This means that \({{\widehat{D}}}^F_{ab}(p)\) is transverse both to \(p^a\) and \(\eta ^a\).

  13. This model was also studied by Chang and Yan in Ref. [20] in Schwinger’s functional-derivative method. However, in light-front dynamics literature it is customary to find the analysis of the neutral model with the \(\gamma ^5\) matrix replaced by the unit matrix 1, under the same name of “Yukawa’s model”; clearly, in it the meson field is not pseudo-scalar, but scalar. That model is studied, for example, in Refs. [7, 43, 44].

  14. Do not confuse with the positive and negative frequency parts, which are denoted by the sub-indices “\({\pm }\)” without parenthesis: \(\psi _{\pm }(x)\). For example, the negative frequency part of \(\Lambda _{(+)}\psi \) is denoted by \(\psi _{(+)-}\), and so on.

References

  1. P.A.M. Dirac, Forms of relativistic dynamics. Rev. Mod. Phys. 21(3), 392–399 (1949)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  2. G.E. Ramos, Teoria de Campos no Plano Nulo: Um estudo, Masters dissertation, IFT-UNESP (2005)

  3. G.E. Ramos, Formulação Canônica no Plano Nulo, PhD. thesis, IFT-UNESP (2009)

  4. H. Leutwyler, J. Stern, Relativistic dynamics on a null plane. Ann. Phys. 112, 94–164 (1978)

    Article  ADS  MathSciNet  Google Scholar 

  5. F. Rohrlich, Theory of photons and electrons on null planes. Acta Phys. Austriaca 32, 87–106 (1970)

    Google Scholar 

  6. F. Rohrlich, Null plane field theory. Acta Phys. Austriaca 8(Suppl), 277–322 (1971)

    Google Scholar 

  7. S.J. Brodsky, H.-C. Pauli, S.S. Pinsky, Quantum chromodynamics and other field theories on the light cone. Phys. Rep. 301, 299–486 (1998)

    Article  ADS  MathSciNet  Google Scholar 

  8. S. Fubini, G. Furlan, Renormalization effects for partially conserved currents. Phys. Phys. Fizika 1, 247–299 (1965)

    MathSciNet  Google Scholar 

  9. S. Weinberg, Dynamics at Infinite Momentum, Phys. Rev. 150, 1313-1318 (1966). Erratum: Phys. Rev. 158, 1638 (1967)

  10. H. Bebié, H. Leutwyler, Relativistically invariant solutions of current algebras at infinite momentum. Phys. Rev. Lett. 19, 618–621 (1967)

    Article  ADS  Google Scholar 

  11. R.A. Neville, Quantum Electrodynamics in a Laser Pulse, Ph.D. Thesis. Syracuse University (1968)

  12. R.A. Neville, F. Rohrlich, Quantum electrodynamics on null planes and applications to lasers. Phys. Rev. D 3(8), 1692–1707 (1971)

    Article  ADS  Google Scholar 

  13. S.-J. Chang, S. Ma, Feynman rules and quantum electrodynamics at infinite momentum. Phys. Rev. 180, 1506–1513 (1969)

    Article  ADS  Google Scholar 

  14. J.B. Kogut, D.E. Soper, Quantum electrodynamics in the infinite-momentum frame. Phys. Rev. D 1(10), 2901–2914 (1970)

    Article  ADS  Google Scholar 

  15. R.A. Neville, F. Rohrlich, Quantum field theory off null planes. Nuovo Cimento 1(4), 625–644 (1971)

    Article  ADS  MathSciNet  Google Scholar 

  16. H. Leutwyler, J.R. Klauder, L. Streit, Quantum Field Theory on Lightlike Slabs, Nuovo Cimento, Vol. LXVI A, N. 3, 536–554 (1970)

  17. J.H. Ten Eyck, Problems in Null-Plane Quantum-Electrodynamics, Ph.D. Thesis. Syracuse University (1973)

  18. J.H. Ten Eyck, F. Rohrlich, Equivalence of null-plane and conventional quantum electrodynamics. Phys. Rev. D 9(8), 2237–2245 (1974)

    Article  ADS  MathSciNet  Google Scholar 

  19. S.-J. Chang, R.G. Root, T.-M. Yan, Quantum field theories in the infinite-momentum frame. I. Quantization of scalar and dirac fields. Phys. Rev. D 7, 1133–1146 (1973)

    Article  ADS  Google Scholar 

  20. S.-J. Chang, T.-M. Yan, Quantum field theories in the infinite-momentum frame. II. Scattering matrices of scalar and dirac fields. Phys. Rev. D 7, 1147–1161 (1973)

    Article  ADS  Google Scholar 

  21. T.-M. Yan, Quantum field theories in the infinite-momentum frame. III. Quantization of coupled spin-one fields. Phys. Rev. D 7, 1760–1780 (1973)

    Article  ADS  Google Scholar 

  22. T.-M. Yan, Quantum field theories in the infinite-momentum frame. IV. Scattering matrix of vector and dirac fields and perturbation theory. Phys. Rev. D 7, 1780–1800 (1973)

    Article  ADS  Google Scholar 

  23. S.J. Brodsky, R. Roskies, R. Suaya, Quantum electrodynamics and renormalization theory in the infinite-momentum frame. Phys. Rev. D 8, 4574–4594 (1973)

    Article  ADS  Google Scholar 

  24. S.D. Drell, D. Levy, T.-M. Yan, Theory of deep-inelastic lepton-nucleon scattering and lepton-pair annihilation processes. I. Phys. Rev. 187, 2159–2171 (1969)

    Article  ADS  Google Scholar 

  25. S.D. Drell, D. Levy, T.-M. Yan, Theory of deep-inelastic Lepton–Nucleon scattering and lepton pair annihilation processes. II. Deep-inelastic electron scattering. Phys. Rev. D 1, 1035–1068 (1970)

    Article  ADS  Google Scholar 

  26. S.D. Drell, D. Levy, T.-M. Yan, Theory of deep-inelastic Lepton–Nucleon scattering and lepton-pair annihilation processes. III. Deep-inelastic electron-positron annihilation. Phys. Rev. D 1, 1617–1639 (1970)

    Article  ADS  Google Scholar 

  27. S.D. Drell, T.-M. Yan, Connection of elastic electromagnetic nucleon form factors at large \(Q^2\) and deep inelastic structure functions near threshold. Phys. Rev. Lett. 24, 181–186 (1970)

    Article  ADS  Google Scholar 

  28. T. Maskawa, K. Yamawaki, The Problem of \(P^+=0\) mode in the null-plane field theory and Dirac’s method of quantization. Progr. Theor. Phys. 56, 270–283 (1976)

  29. H.-C. Pauli, S.J. Brodsky, Solving field theory in one space and one time dimension. Phys. Rev. D 32, 1993–2000 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  30. H.-C. Pauli, S.J. Brodsky, Discretized light-cone quantization: solution to a field theory in one space and one time dimension. Phys. Rev. D 32, 2001–2013 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  31. J.P. Vary, X. Zhao, A. Ilderton, H. Honkanen, P. Maris, S.J. Brodsky, Applications of basis light-front quantization to QED, Nuc. Phys. B (Proc. Suppl.) 251-252, 10-15 (2014)

  32. X. Zhao, H. Honkanen, P. Maris, J.P. Vary, S.J. Brodsky, Electron g-2 in light-front quantization. Phys. Lett. B 737, 65–69 (2014)

    Article  ADS  MATH  Google Scholar 

  33. R.J. Perry, A. Harindranath, K.G. Wilson, Light-front Tamm–Dancoff field theory. Phys. Rev. Lett. 65(24), 2959–2962 (1990)

    Article  ADS  Google Scholar 

  34. B.L.G. Bakker, A. Bassetto, S.J. Brodsky, W. Broniowski, S. Dalley, T. Frederico et al., Light-front quantum chromodynamics. A framework for the analysis of hadron physics. Nucl. Phys. B (Proc. Suppl.) 251–252, 165–174 (2014)

    Article  ADS  Google Scholar 

  35. E. Tomboulis, Quantization of the Yang–Mills field in the null-plane frame. Phys. Rev. D 8, 2736–2740 (1973)

    Article  ADS  Google Scholar 

  36. A. Casher, Gauge fields on the null plane. Phys. Rev. D 14, 452–464 (1976)

    Article  ADS  Google Scholar 

  37. B.M. Pimentel, A.T. Suzuki, Light-cone gauge and the (causal) principal value prescription. Phys. Rev. D 42, 2115–2119 (1990)

    Article  ADS  Google Scholar 

  38. B.M. Pimentel, A.T. Suzuki, Causal prescription for the light-cone gauge. Modern Phys. Lett. A 6, 2649–2653 (1991)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. R. Casana, B.M. Pimentel, G.E.R. Zambrano, SQED\(_4\) and QED\(_4\) on the null-plane. Braz. J. Phys. 44, 398–409 (2014)

    Article  ADS  Google Scholar 

  40. B.M. Pimentel, A.T. Suzuki, G.E.R. Zambrano, Functional analysis for gauge fields on the front-form and the light-cone gauge. Few-Body Syst. 52, 437–442 (2012)

    Article  ADS  Google Scholar 

  41. D. Bhamre, A. Misra, V.K. Singh, Equivalence of one loop diagrams in covariant and light front QED revisited. Few Body Syst. 59, 107 (2018)

    Article  ADS  Google Scholar 

  42. A.T. Suzuki, T.H.O. Sales, The light-front gauge propagator: the status quo. arXiv:hep-th/0408135 (2004)

  43. V.A. Karmanov, J.-F. Mathiot, A.V. Smirnov, Regularization of the fermion self-energy and the electromagnetic vertex in the Yukawa model within light-front dynamics. Phys. Rev. D 75, 045012 (2007)

    Article  ADS  Google Scholar 

  44. B.L.G. Bakker, C.-R. Ji, Light-front singularities in the Yukawa model. Nucl. Phys. B (Proc. Suppl.) 161, 15–20 (2006)

    Article  ADS  Google Scholar 

  45. D. Mustaki, S. Pinsky, J. Shigemitsu, K. Wilson, Perturbative renormalization of null-plane QED. Phys. Rev. D 43(10), 3411–3427 (1991)

    Article  ADS  Google Scholar 

  46. W. Heisenberg, Die ’Beobachtbaren Grössen’ in der Theorie der Elementarteilchen, Z. Phys. 120 (1943) 513–538. This paper can be found in: W. Blum, H.-P. Dürr and H. Rechenberg (Eds.), Werner Heisenberg. Gesammelte Werke / Collected Works. Series A / Part II. Original Scientific Papers (Springer, 1989), in which excelent comments on Heisenberg’s motivation for the definition of the \(S\) matrix are given by R. Oehme

  47. R.F. Streater, A.S. Wightman, PCT, Spin and Statistics, and All That (The Benjamin/Cummings Publishing Company, New York, 1964)

    MATH  Google Scholar 

  48. J. Glimm, A. Jaffe, Quantum Physics (Springer, Berlin, 1981)

    Book  MATH  Google Scholar 

  49. H. Lehmann, K. Symanzik, W. Zimmermann, Zur Furmulierung quantisierter Feldtheorien, Nuovo Cim. 1 (1955) 205–225. Translation into English by C.F. Barger: On the formulation of quantized field theories (CBF Scientific Translations)

  50. H. Araki, Von Neumann algebras of local observables for free scalar field. J. Math. Phys. 5, 1–13 (1964)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  51. R. Haag, D. Kastler, An algebraic approach to quantum field theory. J. Math. Phys. 5, 848–861 (1964)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  52. J. Lacki, H. Ruegg, G. Wanders (Eds.), E.C.G. Stueckelberg, An Unconventional Figure of Twentieth Century Physics. Selected Scientific Papers with Commentaries (Birkhauser, New York, 2009)

  53. E. Stueckelberg, D. Rivier, Causalité et structure de la matrice S, Helv. Phys. Acta 23 (1950) 215–222. This paper was reprinted in Ref. [52]

  54. N.N. Bogoliubov, B.V. Medvedev, M.K. Polivanov, Voprossy Teorii Dispersionnykh Sootnoshenii (Moscow, Fizmatgiz, 1958). Translation into English: N.N. Bogoliubov, B.V. Medvedev and M.K. Polivanov, Problems in the Theory of Dispersion Relations (Institute for Advanced Study, Princeton, 1958)

  55. N.N. Bogoliubov, D. Shirkov, Introduction to the Theory of Quantized Fields, 3rd edn. (Wiley, New York, 1979)

    Google Scholar 

  56. N.N. Bogolubov, A.A. Logunov, A.I. Oksak, I.T. Todorov, General Principles of Quantum Field Theory (Kluwer Academic Publishers, London, 1990)

    Book  Google Scholar 

  57. B.M. Stepanov, On the construction of an \(S\)-matrix in accordance with the theory of perturbations (in russian), Izv. Akad. Nauk SSSR Ser. Mat. 29 (1965) 1037–1054. Translation into English in: Eighteen papers on analysis and quantum mechanics, American Mathematical Society Translations, Series 2, Volume 91 (American Mathematical Society, 1970)

  58. H. Epstein, V. Glaser, The role of locality in perturbation theory. Ann. Inst. H. Poincaré A 19, 211–295 (1973)

    MathSciNet  MATH  Google Scholar 

  59. N.N. Bogoliubow, O.S. Parasiuk, Über die multiplikation der kausalfunktionen in der quantentheorie der felder. Acta Math. 97, 227–266 (1957)

    Article  MathSciNet  MATH  Google Scholar 

  60. S. Łojasiewicz, Sur le problème de la division. Studia Mathematica T. XVII I, 87–136 (1959)

  61. B. Malgrange, Division des distributions. I: Distributions prolongeables, Séminaire Schwartz T. 4, exp. No. 21, 1–5 (1959–1960)

  62. B. Malgrange, Division des distributions. II: L’Inégalité de Łojasiewicz, Séminaire Schwartz T. 4, exp. No. 22, 1–8 (1959–1960)

  63. V.S. Vladimirov, Methods of the Theory of Functions of Many Complex Variables (Dover, New York, 2007)

    Google Scholar 

  64. G. Scharf, Finite Quantum Electrodynamics. The Causal Approach, 3rd edn. (Dover, New York, 2014)

    MATH  Google Scholar 

  65. S. Fassari, G. Scharf, Borel Summability in Finite QED I: Outline and Results, in Stochastic Processes, Physics and Geometry II. ed. by S. Albeverio, U. Cattaneo, D. Merlini (World Scientific, Singapore, 1991), pp. 278–288

    MATH  Google Scholar 

  66. V.S. Vladimirov, Y.N. Drozzinov, B.I. Zavialov, Tauberian Theorems for Generalized Functions (Kluwer Acadamic Publication, New York, 1988)

    Book  Google Scholar 

  67. M. Dütsch, F. Krahe, G. Scharf, Scalar QED Revisited, Il Nuovo Cimento 106 A, 3, 277–307 (1993)

  68. M. Dütsch, T. Hurth, F. Krahe, G. Scharf, Causal construction of Yang–Mills theories. I. Nuovo Cim. A 106, 1029–1041 (1993)

    Article  ADS  MathSciNet  Google Scholar 

  69. M. Dütsch, T. Hurth, F. Krahe, G. Scharf, Causal construction of Yang–Mills theories. II. Nuovo Cim. A 107, 375–406 (1994)

    Article  ADS  MathSciNet  Google Scholar 

  70. M. Dütsch, T. Hurth, G. Scharf, Causal construction of Yang–Mills theories. III. Nuovo Cim. A 108, 679–708 (1995)

    Article  ADS  MathSciNet  Google Scholar 

  71. M. Dütsch, T. Hurth, G. Scharf, Causal construction of Yang–Mills theories. IV. Unitarity. Nuovo Cim. A 108, 737–773 (1995)

    Article  ADS  MathSciNet  Google Scholar 

  72. M. Dütsch, G. Scharf, Electroweak theory without spontaneous symmetry breaking, arXiv:hep-th/9612091 (1996)

  73. A.W. Aste, G. Scharf, NonAbelian gauge theories as a consequence of perturbative quantum gauge invariance. Int. J. Mod. Phys. A 14, 3421–3434 (1998)

    Article  ADS  MATH  Google Scholar 

  74. A.W. Aste, G. Scharf, M. Dütsch, Perturbative gauge invariance: electroweak theory II. Ann. Phys. 8, 389–404 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  75. G. Scharf, Gauge Field Theories: Spin One and Spin Two. 100 Years After General Relativity (Dover, New York, 2016)

    Google Scholar 

  76. M. Dütsch, Slavnov–Taylor identities from the causal point of view. Int. J. Mod. Phys. A 12, 3205–3248 (1997)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  77. M. Dütsch, Non-uniqueness of quantized Yang–Mills theories. J. Phys. A: Math. Gen. 29, 7597–7617 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  78. D.R. Grigore, Trivial Lagrangians in the causal approach. Rom. J. Phys. 61, 320–332 (2016)

    Google Scholar 

  79. D.R. Grigore, Cohomological aspects of gauge invariance in the causal approach. Rom. J. Phys. 55, 386–438 (2010)

    MathSciNet  MATH  Google Scholar 

  80. D.R. Grigore, The structure of the anomalies of gauge theories in the causal approach. J. Phys. A 35, 1665–1689 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  81. D.R. Grigore, G. Scharf, No-go result for supersymmetric gauge theories in the causal approach. Ann. Phys. 17, 864–880 (1997)

    MathSciNet  MATH  Google Scholar 

  82. G. Scharf, W.F. Wreszinski, B.M. Pimentel, J.L. Tomazelli, Causal approach to (2+1)-dimensional QED. Ann. Phys. 231, 185–208 (1994)

    Article  ADS  Google Scholar 

  83. L.A. Manzoni, B.M. Pimentel, J.L. Tomazelli, Causal theory for the gauged Thirring model. Eur. Phys. J. C 8, 353–361 (1999)

    Article  ADS  MathSciNet  Google Scholar 

  84. L.A. Manzoni, B.M. Pimentel, J.L. Tomazelli, Radiative corrections to the causal Thirring model. Eur. Phys. J. C 12, 701–705 (2000)

    Article  ADS  Google Scholar 

  85. J.T. Lunardi, B.M. Pimentel, J.S. Valverde, L.A. Manzoni, Duffin–Kemmer–Petiau theory in the causal approach. Int. J. Mod. Phys. A 17(2), 205–227 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  86. J. Beltrán, B.M. Pimentel, D.E. Soto, The causal approach proof for the equivalence of SDKP\(_4\) and SQED\(_4\) at tree-level. Int. J. Mod. Phys. A 35(9), 2050042 (2020)

    Article  ADS  MathSciNet  Google Scholar 

  87. R. Bufalo, B.M. Pimentel, D.E. Soto, Causal approach for the electron-positron scattering in generalized quantum electrodynamics. Phys. Rev. D 90, 085012 (2014)

    Article  ADS  Google Scholar 

  88. R. Bufalo, B.M. Pimentel, D.E. Soto, Normalizability analysis of the generalized quantum electrodynamics from the causal point of view. Int. J. Mod. Phys. A 32, 1750165 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  89. O.A. Acevedo, J. Beltrán, B.M. Pimentel, D.E. Soto, Causal perturbation theory approach to Yukawa’s model. Eur. Phys. J. Plus 136, 895 (2021)

  90. M. Dütsch, K. Fredenhagen, A local (perturbative) construction of observables in gauge theories: the example of QED. Commun. Math. Phys. 203, 71–105 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  91. R. Brunetti, K. Fredenhagen, Microlocal analysis and interacting quantum field theories: renormalization on physical backgrounds. Commun. Math. Phys. 208, 623–661 (2000)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  92. M. Dütsch, K. Fredenhagen, Causal perturbation theory in terms of retarded products, and a proof of the Action Ward Identity. Rev. Math. Phys. 16, 1291–1348 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  93. R. Brunetti, M. Dütsch, K. Fredenhagen, Perturbative algebraic quantum field theory and the renormalization groups. Adv. Theor. Math. Phys. 13, 1541–1599 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  94. K. Fredenhagen, K. Rejzner, Batalin–Vilkoviski formalism in perturbative algebraic quantum field theory. Commun. Math. Phys. 317, 697–725 (2013)

    Article  ADS  MATH  Google Scholar 

  95. M. Dütsch, From Classical Field Theory to Perturbative Quantum Field Theory (Birkkhäuser, London, 2019)

    Book  MATH  Google Scholar 

  96. S. Hollands, R. Wald, Local wick polynomials and time ordered products of quantum fields in curved spacetime. Commun. Math. Phys. 223, 289–326 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  97. S. Hollands, R.M. Wald, Existence of local covariant time-ordered products of quantum fields in curved spacetime. Commun. Math. Phys. 231, 309–345 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  98. R. Bufalo, B.M. Pimentel, D.E. Soto, The Epstein–Glaser causal approach to the light-front QED\(_4\). I: free theory. Ann. Phys. 351, 1034–1061 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  99. R. Bufalo, B.M. Pimentel, D.E. Soto, The Epstein–Glaser causal approach to the light-front QED\(_4.\) II: vacuum polarization tensor. Ann. Phys. 351, 1062–1084 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  100. O.A. Acevedo, B.M. Pimentel, D.E. Soto, Epstein–Glaser’s causal light-front field theory, Proc. Sci. LC2019, 021 (2020)

  101. A.N. Tikhonov, A.A. Samarskii, Equations of Mathematical Physics, Reprint ed. (Dover, 2011)

  102. J. Hadamard, Lectures on Cauchy’s Problem in Linear Partial Differential Equations (Yale University Press, New York, 1923)

  103. E. Goursat, A Course in Mathematical Analysis, vol. I (Ginn and Company, London, 1904)

    MATH  Google Scholar 

  104. S. Chandrasekhar, The Mathematical Theory of Black Holes (Oxford Classic Texts in the Physical Sciences, Clarendon Press, Oxford, 2006)

    MATH  Google Scholar 

  105. P.P. Srivastava, Perspectives of Light-Front Quantized Field Theory: Some New Results, arXiv:hep-ph/9908492v1 (1999)

  106. L. Schwartz, Méthodes mathématiques pour les sciences physiques, 2 edn. (Hermann, Pairs, 1979)

  107. F.G. Friedlander, M. Joshi, Introduction to the Theory of Distributions, 2nd edn. (Cambridge University Press, Cambridge, 1998)

    Google Scholar 

  108. J. Hilgevoord, Dispersion Relations and Causal Description (North Holland, Amsterdam, 1960)

    MATH  Google Scholar 

  109. D. Prange, Lorentz covariance in Epstein-Glaser renormalization. arXiv preprint: hep-th/9904136 (1999)

  110. T. Hurth, K. Skenderis, Quantum Noether method. Nucl. Phys. B 541, 566–614 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  111. A.G. Sveshnikov, A.N. Tikhonov, The Theory of Functions of a Complex Variable (Mir Publishers, New York, 1971)

    MATH  Google Scholar 

  112. P.P. Srivastava, S.J. Brodsky, Light-front quantized QCD in the light-cone gauge: the doubly transverse gauge propagator. Phys. Rev. D 64, 045006 (2001)

    Article  ADS  Google Scholar 

  113. K. Nishijima, Fundamental Particles (Benjamin, W.A, 1963)

    MATH  Google Scholar 

  114. V. Mukherji, A history of the meson theory of nuclear forces from 1935 to 1952. Arch. Hist. Exact Sci. 13(1), 27–102 (1974)

    Article  MathSciNet  Google Scholar 

  115. L.M. Brown, H. Rechenberg, The Origin of the Concept of Nuclear Forces (CRC Press, London, 1996)

    Google Scholar 

  116. L.M. Brown, H. Rechenberg, Quantum field theories, nuclear forces, and the cosmic rays (1934–1938). Am. J. Phys. 59(7), 595–605 (1991)

    Article  ADS  Google Scholar 

  117. W. Heisenberg, Über den Bau der Atomkerne. I, Z. Phys. 77 (1932) 1–11. Translation into English in: D.M. Brink, Nuclear Forces (Oxford, 1965)

  118. W. Heisenberg, Über den Bau der Atomkerne. II, Z. Phys. 78, 156–164 (1932)

  119. W. Heisenberg, Über den Bau der Atomkerne. III, Z. Phys. 80 (1933) 587–596. Partial translation into English in: D.M. Brink, Nuclear Forces (Oxford, 1965)

  120. A.I. Miller, Werner Heisenberg and the beginning of nuclear physics. Phys. Today 38(11), 60–68 (1985)

    Article  Google Scholar 

  121. H. Yukawa, On the interaction of elementary particles I. Proc. Physico-Math. Soc. Jpn. 17, 48–57 (1935)

    MATH  Google Scholar 

  122. N. Kemmer, The charge-dependence of nuclear forces. Proc. Camb. Philos. Soc. 34, 354–364 (1938)

    Article  ADS  Google Scholar 

  123. G.C. Wick, Range of nuclear forces in Yukawa’s theory. Nature 142, 993–994 (1938)

  124. N. Kemmer, Quantum theory of Einstein-Bose particles and nuclear interaction. Proc. Roy. Soc. (Lond.) A 166, 127–153 (1938)

    Article  ADS  MATH  Google Scholar 

  125. S. Glazek, A. Harindranath, S. Pinsky, J. Shigemitsu, K. Wilson, Relativistic bound-state problem in the light-front Yukawa model. Phys. Rev. D 47, 1599–1619 (1993)

    Article  ADS  Google Scholar 

  126. O.A. Acevedo, B.M. Pimentel, Radiative corrections in the Yukawa model within the null-plane causal perturbation theory framework. Phys. Rev. D 103, 076022 (2021)

    Article  ADS  MathSciNet  Google Scholar 

  127. R. Casana, B.M. Pimentel, G.E.R. Zambrano, SQED\(_2\) on the null-plane. Revista Colombiana de Física 41(1), 220–222 (2009)

    Google Scholar 

Download references

Acknowledgements

O.A.A. thanks CAPES-Brazil and CNPq-Brazil for total financial support, and B.M.P. thanks CNPq-Brazil for partial financial support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to O. A. Acevedo.

Additional information

Dedicated to Prof. Günter Robert Scharf, in memoriam.

Quantization of fields on the null-plane

Quantization of fields on the null-plane

In quantum mechanics a particle is described by a wave-function. If the description is relativistic, such wave-functions are the positive frequency part of the so-called relativistic classical fields, this is to say, solutions of some relativistic equation of motion. The restriction to consider only the positive frequency part of the solution is done because in the real world the energy of the particle must be positive; the frequency and the energy are related by Schrödinger’s equation, which is an axiom in the quantum theory. In instant dynamics this could be confusing: We know that every component of every field satisfies Klein–Gordon–Fock’s equation, but it is impossible to put that equation in Schrödinger’s form. In light-front dynamics this issue is clear: Klein–Gordon–Fock’s equation can be put in Schrödinger’s form,

$$\begin{aligned} i\partial _+\varphi _1(x)=P_+\varphi _1(x), \end{aligned}$$
(244)

with the Hamiltonian:

$$\begin{aligned} P_+=\frac{i}{2}\partial _-^{-1}(\partial _\perp ^2-m^2)=\frac{i}{4}\int dv\text {sgn}(x^--v)(\partial _\perp ^2-m^2). \end{aligned}$$
(245)

All the solutions \(\varphi _1\) of this equation constitute the one-particle Hilbert space \({\mathcal {H}}_1\). This is sufficient if we want to treat free particles only, but when the particles are subjected to interaction we need to consider the possibility that they change their identity and their number. Of course, such changes cannot be described in detail: the very concept of particle is then lost; but we can describe the asymptotically free incoming and outgoing particles of a scattering process. And that could be done only if one works on a space containing states describing different number of particles.

Definition

Fock’s space \({\mathcal {F}}^{\pm }\) of bosons or fermions, respectively, is the direct sum of Hilbert’s spaces of symmetrized or anti-symmetrized wave-functions of n particles, \({\mathcal {H}}_n^{\pm }\), extended to all non-negative values of n:

$$\begin{aligned} {\mathcal {F}}^{\pm }:=\bigoplus \limits _{n=0}^{+\infty }{\mathcal {H}}_n^{\pm }. \end{aligned}$$
(246)

Their elements are:

$$\begin{aligned} \Phi \equiv \left( \varphi _0;\varphi _1;\cdots ;\varphi _n;\cdots \right) ;\quad \varphi _n\in {\mathcal {H}}_n^{\pm }. \end{aligned}$$
(247)

Particularly, the vacuum state of zero particles is:

$$\begin{aligned} \Omega :=\left( 1;0;\cdots \right) . \end{aligned}$$
(248)

Fock’s space \({\mathcal {F}}^{\pm }\) is equipped with the interior product \(\left( \bullet ;\bullet \right) \) defined as:

$$\begin{aligned} (\Phi ;\Psi ):=\sum \limits _{n=0}^{+\infty }(\varphi _n;\psi _n)_n, \end{aligned}$$
(249)

and norm induced by it:

$$\begin{aligned} \Vert \Phi \Vert ^2:=\sum \limits _{n=0}^{+\infty }\Vert \varphi _n\Vert _n^2=\sum \limits _{n=0}^{+\infty }(\varphi _n;\varphi _n)_n, \end{aligned}$$
(250)

with the restriction—over the states—that its value to be finite.

The passage from states of a given number of particles to another is done by the emission and absorption operators.

Definition

The emission operator of a particle with wave-function f, with \(f\in {\mathcal {H}}_1\), is the operator \(a^*(f):\text {Dom}(a^*(f))\subseteq {\mathcal {F}}^{\pm }\rightarrow \mathcal F^{\pm }\), defined as:

$$\begin{aligned} (a^*(f)\Phi )_0:=0;\quad (a^*(f)\Phi )_n:=\sqrt{n}S_n^{\pm }(f\otimes \varphi _{n-1})\quad (n\in {\mathbb {N}}). \end{aligned}$$
(251)

The absorption operator of a particle with wave-function f, with \(f\in {\mathcal {H}}_1\), is the operator \(a(f):{\mathcal {F}}^{\pm }\rightarrow {\mathcal {F}}^{\pm }\), defined such that:

$$\begin{aligned} a(f)\Omega :=0;\quad (a(f)\Phi )_n(\varvec{x}_1;\cdots ;\varvec{x}_n):=\sqrt{n+1}\left( f(\varvec{x});\varphi _{n+1}(\varvec{x};\varvec{x}_1;\cdots ;\varvec{x}_n)\right) _{1,\varvec{x}}. \end{aligned}$$
(252)

The Hermiticity of the (anti-)symmetrization operators \(S^{\pm }_n\) implies the following relation between the emission and absorption operators:

$$\begin{aligned} a^*(f)=a(f)^\dagger \equiv a^\dagger (f), \end{aligned}$$
(253)

The (anti-)commutation relations of the emission and absorption operators are found by using their definitions given in Eqs. (251) and (252). Let \(f,g\in {\mathcal {H}}_1\), \(\Phi \in \text {Dom}\left( [a(f);a^\dagger (g)]_{{\mp }}\right) \subset {\mathcal {F}}^{\pm }\). By direct calculus:

$$\begin{aligned}&\displaystyle \left[ a(f);a^\dagger (g)\right] _{\mp }\Phi =(f;g)_1\Phi , \end{aligned}$$
(254)
$$\begin{aligned}&\displaystyle \left[ a(f);a(g)\right] _{\mp }\Phi =0;\quad \left[ a^\dagger (f);a^\dagger (g)\right] _{\mp }\Phi =0\, . \end{aligned}$$
(255)

As we see, these relations consider the commutator in \({\mathcal {F}}^+\) (bosons space) and the anti-commutator in \({\mathcal {F}}^-\) (fermions space). Under the assumption that \({\mathcal {H}}_1\) is separable, it is possible to introduce a countable and orthonormal basis of functions \(f_j\):

$$\begin{aligned} (f_j;f_k)_1=\delta _{jk}, \end{aligned}$$
(256)

in function of which every \(f\in {\mathcal {H}}_1\) is expanded as:

$$\begin{aligned} f(\varvec{x})=\sum _j(f_j;f)_1f_j(\varvec{x}). \end{aligned}$$
(257)

Therefore, defining the emission and absorption operators on those basis functions:

$$\begin{aligned} a^\dagger _j:=a^\dagger (f_j),\quad a_j:=a(f_j), \end{aligned}$$
(258)

we obtain that \(a_j\) and \(a^\dagger _k\) satisfy the following relations [see Eqs. (254), (255) and (256)]:

$$\begin{aligned} {[}a_j;a^\dagger _k]_{\mp }=\delta _{jk},\quad [a_j;a_k]_{{\mp }}=0=\left[ a^\dagger _j;a^\dagger _k\right] _{\mp }. \end{aligned}$$
(259)

Now we must specify the interior product in Eq. (254) in such a way that it is relativistic invariant.

Definition

Let \({\mathcal {H}}_1\) be the one-particle Hilbert’s space. Consider \(f,g\in {\mathcal {H}}_1\), and let \({{\hat{f}}}\) and \({{\hat{g}}}\) be their Fourier’s transforms. The interior product of them is defined by:

$$\begin{aligned} (f;g)_1:=\int {\mathrm{d}}\mu _m(p){{\hat{f}}}(p)^*{{\hat{g}}}(p), \end{aligned}$$
(260)

with \({\mathrm{d}}\mu _m(p)\) the relativistic invariant measure:

$$\begin{aligned} {\mathrm{d}}\mu _m(p):=\delta (p^2-m^2)\Theta (p_-)d^4p=\left. \Theta (p_-)\frac{d^3\varvec{p}}{2p_-}\right| _{p_+=E};\quad E:=\frac{\omega _p^2}{|2p_-|}=\frac{p_\perp ^2+m^2}{|2p_-|}. \end{aligned}$$
(261)

This measure \({\mathrm{d}}\mu _m(p)\) is defined on the “upper mass-shell”

$$\begin{aligned} {\mathcal {M}}^+:=\left\{ p\in {\mathbb {R}}^4\ \big |\ p^2=m^2\ \wedge \ p_->0\right\} , \end{aligned}$$
(262)

so we identify the space \({\mathcal {H}}_1\) with the space of square-integrable functions on \({\mathcal {M}}^+\) with respect to the measure \({\mathrm{d}}\mu _m\):

$$\begin{aligned} {\mathcal {H}}_1:=L^2({\mathcal {M}}^+;{\mathrm{d}}\mu _m). \end{aligned}$$
(263)

We present the following results without proof—the proof in instant dynamics is given in Ref. [64]; in light-front dynamics the proof follows with simple modifications.

Theorem

Every irreducible representation of the (anti-)commutation relations of Eqs. (254) and (255), with vacuum state \(\Omega \), are equivalent to Fock’s representation.

Corollary

Every bounded operator acting on Fock’s space \(\mathcal F^{\pm }\) can be expressed in function of the emission and absorption operators \(a^\dagger (f)\) and a(f).

Also, Eq. (261) implies that, when one restricts the wave-functions to the null-plane, the measure of the interior product becomes independent of the mass. This leads to the following theorem [16]:

Theorem

Let \({\mathcal {F}}_\nu ^{\pm }\) (\(\nu =1,2\)) be two Fock’s spaces corresponding to particles with different masses \(m_\nu \), with \(m_1\ne m_2\). The Fock’s spaces \({\mathcal {F}}_1^{\pm }\) and \(\mathcal F_2^{\pm }\) are unitarily equivalents.

The next step in the construction of the quantum theory is the adequate election of the basis functions in \({\mathcal {H}}_1\). If the basis is chosen in a such a way to satisfy the equations and constraints of the classical field theory, then we will have no problem in recognizing that their evolution according to the Hamiltonian \(P_+\) is relativistic. Our strategy will be the following: We will write a basis for the physical solutions of Klein–Gordon–Fock’s equation; with them we will construct the complete multi-component fields by using the constraint equations and the adequate statistics. In order to do that, we remember that the initial value problem in the null-plane has the solution for the positive frequency part [5, 6]:

$$\begin{aligned} f(x)=\left\langle iD_+(x-y)^*;f(y)\right\rangle _y, \end{aligned}$$
(264)

with \(\left\langle \bullet ;\bullet \right\rangle \) meaning that the distribution is applied to the function according to the formal rule given by the interior product \((\bullet ;\bullet )_1\), differentiating the notation in order to clarify that \(\left\langle \bullet ;\bullet \right\rangle \) is not an interior product, but the application of a distribution to a function. Also, \(D_+\) is the positive frequency part of Jordan-Pauli’s distribution. The expansion of f in the basis \(\left\{ f_j\right\} \) implies the following completeness relation:

$$\begin{aligned} \sum _j f_j(x)f_j(y)^*=-iD_+(x-y). \end{aligned}$$
(265)

The mapping from the function f to the operators a(f) and \(a^\dagger (f)\) is making by:

Definition

Let \(f\in {\mathcal {H}}_1\) be a wave-function. The emission and absorption field operators in real space, \(a^\dagger (\varvec{x})\) and \(a(\varvec{x})\), respectively, are the operator-valued distributions which lead f to the operators \(a^\dagger (f)\) and a(f), respectively, by means of the rules:

$$\begin{aligned} a^\dagger (f):=\left\langle f^*;a^\dagger (\varvec{x})\right\rangle ;\quad a(f):=\left\langle f;a(\varvec{x})\right\rangle , \end{aligned}$$
(266)

Additionally, if \(f\in {\mathscr {S}}({\mathbb {R}}^3)\), we define the emission and absorption field operators in momentum space, \(a^\dagger (\varvec{p})\) and \(a(\varvec{p})\), respectively, as the distributional Fourier’s transforms of the corresponding field operators in real space.

Eqs. (257) and (266) imply that we can write the emission and absorption field operators as:

$$\begin{aligned} a^\dagger (\varvec{x})=\sum _j a^\dagger _j f_j(\varvec{x})^*,\quad a(\varvec{x})=\sum _j a_jf_j(\varvec{x}), \end{aligned}$$
(267)

and in momentum space:

$$\begin{aligned} a^\dagger (\varvec{p})=\sum _j {{\hat{a}}}^\dagger _j {{\hat{f}}}_j(\varvec{p})^*,\quad a(\varvec{p})=\sum _j {{\hat{a}}}_j{{\hat{f}}}_j(\varvec{p}). \end{aligned}$$
(268)

From these relations, jointly with the expansions in Eq. (259) and the completeness relation given in Eq. (265):

$$\begin{aligned} \left[ a(\varvec{x});a^\dagger (\varvec{y})\right] _{\mp }=-iD_+(0;\varvec{x}-\varvec{y}),\quad [a(\varvec{x});a(\varvec{y})]_{\mp }=0=\left[ a^\dagger (\varvec{x});a^\dagger (\varvec{y})\right] _{\mp }. \end{aligned}$$
(269)

Definition

The quantized field operator is the operator-valued distribution u(x) defined as:

$$\begin{aligned} u(x):=a(x)+a^\dagger (x). \end{aligned}$$
(270)

Its part corresponding to the absorption field operator is called “negative frequency part”, denoted \(u_-\); the corresponding to the emission field operator, “positive frequency part”, denoted \(u_+\).

To that definition many forms can be given. Using the expansion showed in Eq. (267):

$$\begin{aligned} u(x)=\sum _j\left( a_jf_j(x)+a_j^\dagger f_j(x)^*\right) . \end{aligned}$$
(271)

Here we see that the dependence with the coordinates is contained in the functions \(f_j(x)\). Since \(\left\{ f_k\right\} \) is basis of \({\mathcal {H}}_1\), Eq. (271) implies that the quantized field operator satisfies the equation of motion of the classical field. Also, writing the function \(f_j(x)\) in Eq. (271) in function of their Fourier’s transform,

$$\begin{aligned} f_j(x)=(2\pi )^{-3/2}\int {\mathrm{d}}\mu (\varvec{p}){{\hat{f}}}_j(\varvec{p})e^{-ipx}, \end{aligned}$$
(272)

we obtain the following wave-packet expansion for the quantized field operator:

$$\begin{aligned} u(x)=(2\pi )^{-3/2}\sum _j\int {\mathrm{d}}\mu (\varvec{p})\left( a_j{{\hat{f}}}_j(\varvec{p})e^{-ipx}+a^\dagger _j{{\hat{f}}}(\varvec{p})^*e^{ipx}\right) ;\quad p_+=\frac{\omega _p}{|2p_-|}. \end{aligned}$$
(273)

Finally, by applying Fourier’s transformation directly to the emission and absorption field operators a(x) and \(a^\dagger (x)\) in Eq. (270):

$$\begin{aligned} u(x)=(2\pi )^{-3/2}\int {\mathrm{d}}\mu (\varvec{p})\left( a(\varvec{p})e^{-ipx}+a^\dagger (\varvec{p})e^{ipx}\right) . \end{aligned}$$
(274)

Once again the field operators \(a^\dagger (\varvec{p})\) and \(a(\varvec{p})\) can be expanded as in Eq. (268), with the same result as in Eq. (273); this shows that, in general: \({{\hat{a}}}_j=a_j\) and \({{\hat{a}}}_j^\dagger =a_j^\dagger \); this is not surprising, because \(a_j\) and \(a_j^\dagger \) are independent of the coordinates and momenta.

The most important property of the quantized field operators is the (anti-)commutation relation between their negative and positive frequency parts. Using the expression in Eq. (271) and Eqs. (259) and (265):

$$\begin{aligned} \left[ u_-(x);u_+(y)\right] _{\mp }=-iD_+(x-y). \end{aligned}$$
(275)

This commutator appears in a central rôle in CPT and was called “Wick’s contraction”: It comes from the solution of the classical field theory Goursat’s problem. If the same were calculated from Eq. (274), then we would obtain:

$$\begin{aligned} \left[ u_-(x);u_+(y)\right] _{\mp }=(2\pi )^{-3}\int {\mathrm{d}}\mu (\varvec{p}){\mathrm{d}}\mu (\varvec{q})\left[ a(\varvec{p});a^\dagger (\varvec{q})\right] _{\mp } \left. e^{-i(px-qy)}\right| _{p_+=E_p,q_+=E_q}. \end{aligned}$$
(276)

The comparison between Eqs. (275) and (276) allows the obtention of \(\left[ a(\varvec{p});a^\dagger (\varvec{q})\right] _{\mp }\).

Now we turn to the study of the scalar, fermion and massless vector fields.

1.1 Neutral scalar field

In this case we need to introduce one set of emission and absorption operators, only, \(a^\dagger (f)\) and a(f). The quantized field operator is:

$$\begin{aligned} \varphi (x)=(2\pi )^{-3/2}\int {\mathrm{d}}\mu (\varvec{p})\left( a(\varvec{p})e^{-ipx}+a^\dagger (\varvec{p})e^{ipx}\right) . \end{aligned}$$
(277)

Their commutation distribution will be Jordan-Pauli’s distribution, because it solves Goursat’s problem in the classical theory. Then:

$$\begin{aligned} \left[ \varphi _-(x);\varphi _+(y)\right] =-iD_+(x-y), \end{aligned}$$
(278)

and:

$$\begin{aligned} {[}\varphi (x);\varphi (y)]=-iD(x-y). \end{aligned}$$
(279)

Jordan-Pauli’s distribution is:

$$\begin{aligned} D(x)=i(2\pi )^{-3}\int d^4p\delta (p^2-m^2)\text {sgn}(p_-)e^{-ipx}; \end{aligned}$$
(280)

their positive and negative frequency parts are, respectively, given by:

$$\begin{aligned} D_+(x)=i(2\pi )^{-3}\int {\mathrm{d}}\mu _m(p)e^{-ipx},\quad D_-(x)=-D_+(-x). \end{aligned}$$
(281)

From Eqs. (278), (281) and (276) we obtain:

$$\begin{aligned} \int {\mathrm{d}}\mu (\varvec{p})\left. e^{-ip(x-y)}\right| _{p_+=E_p}=\int {\mathrm{d}}\mu (\varvec{p}){\mathrm{d}}\mu (\varvec{q})\left[ a(\varvec{p});a^\dagger (\varvec{q})\right] \left. e^{-i(px-qy)}\right| _{p_+=E_p,q_+=E_q}, \end{aligned}$$
(282)

which implies:

$$\begin{aligned} \left[ a(\varvec{p});a^\dagger (\varvec{q})\right] =2p_-\delta (\varvec{p}-\varvec{q}). \end{aligned}$$
(283)

Particularly, from Eq. (279) we can obtain the commutation relation at equal times. From Eq. (280), by integrating the variables \(p^+\) and \(p^\perp \), we find:

$$\begin{aligned} D(0;\varvec{x})=\frac{1}{2\pi }\delta (x^\perp )\int \limits _0^{+\infty }dp_-\frac{\sin (p_-x^-)}{p_-}=\frac{1}{4}\text {sgn}(x^-)\delta (x^\perp ), \end{aligned}$$
(284)

hence:

$$\begin{aligned} \left[ \varphi (x^+;\varvec{x});\varphi (x^+;\varvec{y})\right] =-\frac{i}{4}\text {sgn}(x^--y^-)\delta (x^\perp -y^\perp ). \end{aligned}$$
(285)

Here we observe the non-commutativity of the field in the \(x^-\)-axis, which does not violate causality. Additionally, taking the derivative with respect to \(y^-\) in Eq. (285):

$$\begin{aligned} \left[ \varphi (x^+;\varvec{x});\partial _-\varphi (x^+;\varvec{y})\right] =-\frac{i}{2}\delta (\varvec{x}-\varvec{y}). \end{aligned}$$
(286)

In a Hamiltonian analysis, the conjugate momenta to the field \(\varphi \) is \(\pi =\partial _-\varphi \), so Eq. (286) is the canonical commutation relation at equal times. Equation (286), however, is unusual in the sense that it contains a factor of 1/2, which is due to the fact that \(\varphi \) and \(\pi \) are not dynamically independent in light-front dynamics [6]. It is usual that systems which are not constrained in instant dynamics turn to be constrained in the light-front one; we remember to the reader, in this sense, that the stability group of the null-plane is bigger than that of the spatial plane and the number of dynamical generators is less [2, 3]. Equation (286) is the same relation found by the Dirac–Bergmann algorithm [127]; we have found it without using this algorithm because we have constructed Fock’s space with an appropriate basis of one-particle wave-functions, which already satisfy all the constraints of the theory. Finally, we see that the commutation relations at equal times require only the knowledge of the field on the null-plane.

1.2 Fermion field

Consider now Dirac’s fermion field, so we will use anti-commutation relations. Since it describes two particles, each one with two polarization states, as it has spin 1/2, we will use four sets of emission and absorption operators: each set for each polarization of each particle. We start by remembering the reader that the classical Dirac’s field is decomposed into two parts by the projection operators \(\Lambda _{({\pm })}\):Footnote 14

$$\begin{aligned} \psi (x)=\psi _{(+)}(x)+\psi _{(-)}(x);\quad \psi _{({\pm })}:=\Lambda _{({\pm })}\psi , \quad \Lambda _{({\pm })}:=\frac{1}{\sqrt{2}}\gamma ^0\gamma ^{\pm }. \end{aligned}$$
(287)

From them, only \(\psi _{(+)}\) is dynamical and follows Klein–Gordon–Fock’s equation. The projection \(\psi _{(-)}\), on the other hand, is obtained from \(\psi _{(+)}\) by means of the constraint equation:

$$\begin{aligned} \psi _{(-)}=\frac{1}{i\sqrt{2}\partial _-}\gamma ^0(m-i\gamma ^\perp \partial _\perp )\psi _{(+)}. \end{aligned}$$
(288)

If we decompose \(\psi _{(+)}(x)\) in the basis of the space projected by \(\Lambda _{(+)}\):

$$\begin{aligned} \psi _{(+)}(x)=\left( \begin{matrix} 0\\ 1\\ 0\\ 0 \end{matrix}\right) \alpha (x)+\left( \begin{matrix} 0\\ 0\\ 1\\ 0 \end{matrix}\right) \beta (x), \end{aligned}$$
(289)

then we can consider \(\alpha (x)\) and \(\beta (x)\) as fermion charged scalar fields, so their corresponding quantized field operators are:

$$\begin{aligned}&\alpha (x)=(2\pi )^{-3/2}\int {\mathrm{d}}\mu (\varvec{p})\left( b_1(\varvec{p})e^{-ipx}+d_1^\dagger (\varvec{p})e^{ipx}\right) , \end{aligned}$$
(290)
$$\begin{aligned}&\beta (x)=(2\pi )^{-3/2}\int {\mathrm{d}}\mu (\varvec{p})\left( b_{-1}(\varvec{p})e^{-ipx}+d_{-1}^\dagger (\varvec{p})e^{ipx}\right) . \end{aligned}$$
(291)

They are subjected to the following anti-commutation relations:

$$\begin{aligned} \left\{ \alpha (x);{{\widetilde{\alpha }}}(y)\right\} =-iD(x-y),\quad \left\{ \beta (x);{{\widetilde{\beta }}}(y)\right\} =-iD(x-y), \end{aligned}$$
(292)

and for the emission and absorption field operators (\(s={\pm }1\)):

$$\begin{aligned} \left\{ b_s(\varvec{p});b_s^\dagger (\varvec{q})\right\} =2p_-\delta (\varvec{p}-\varvec{q}),\quad \left\{ d_s(\varvec{p});d_s^\dagger (\varvec{q})\right\} =2p_-\delta (\varvec{p}-\varvec{q}). \end{aligned}$$
(293)

Although the quantization is complete, it is convenient to use Eq. (288) to reconstruct the entire \(\psi (x)\) field and write the anti-commutation relation in function of it. Substituting Eqs. (290) and (291) into Eq. (289), we obtain \(\psi _{(+)}\). Then, using Eq. (288) we find \(\psi _{(-)}\); their sum—as in Eq. (287)—is the quantized field operator \(\psi (x)\). In this procedure, no degree of freedom is increased, and the complete field is implicitly written in function of the dynamical degrees of freedom of the theory, \(\alpha (x)\) and \(\beta (x)\). We obtain:

$$\begin{aligned} \psi (x)=(2\pi )^{-3/2}\sum _s\int {\mathrm{d}}\mu (\varvec{p})\sqrt{2p_-}\left( u_s(\varvec{p})b_s(\varvec{p})e^{-ipx}+v_s(\varvec{p})d_s^\dagger (\varvec{p})e^{ipx}\right) . \end{aligned}$$
(294)

And for the adjoint field:

$$\begin{aligned} {{\overline{\psi }}}(x)=(2\pi )^{-3/2}\sum _s\int {\mathrm{d}}\mu (\varvec{p})\sqrt{2p_-}\left( {{\overline{v}}}_s(\varvec{p})d_s(\varvec{p})e^{-ipx}+\overline{u}_s(\varvec{p})b_s^\dagger (\varvec{p})e^{ipx}\right) , \end{aligned}$$
(295)

with the four-component functions \(u, {{\overline{u}}}\) and \(v,\overline{v}\) normalized so as to satisfy the sum rules:

$$\begin{aligned} \sum _s u_s(\varvec{p}){\overline{u}}_s(\varvec{p})= & {} \frac{E\gamma ^++|p_-|\gamma ^-+p_\perp \gamma ^\perp +m}{|2p_-|}, \end{aligned}$$
(296)
$$\begin{aligned} \sum _s v_s(\varvec{p}){{\overline{v}}}_s(\varvec{p})= & {} \frac{E\gamma ^++|p_-|\gamma ^-+p_\perp \gamma ^\perp -m}{|2p_-|}. \end{aligned}$$
(297)

With the expressions given in Eqs. (294) and (295) and using the anti-commutation rules of Eq. (293) we find:

(298)

and:

(299)

Therefore, since \(\Theta (p_-)-\Theta (-p_-)=\text {sgn}(p_-)\), the anti-commutator of Dirac’s field with its adjoint is:

(300)

We identify here the distribution S(x) which in the classical case solves Goursat’s problem for Dirac’s equation:

$$\begin{aligned} \left\{ \psi (x);{{\overline{\psi }}}(y)\right\} =-iS(x-y), \end{aligned}$$
(301)

with:

(302)

1.3 Neutral massless vector field

As it is established at the classical level, the neutral massless vector field \(A^a\) has only two degrees of freedom, corresponding to the transversal components \(A^\alpha \) (\(\alpha =1,2\)), obeying Klein–Gordon–Fock’s equation:

$$\begin{aligned} \square A^\alpha =0\, . \end{aligned}$$
(303)

The component \(A^+\) is zero—this is the null-plane gauge condition, imposed in order to have a well-defined initial value problem [6], and the component \(A^-\) can be obtained from Lorenz’s gauge condition, implied by the null-plane one in the free case:

$$\begin{aligned} A^+=0,\quad A^-=-\partial _-^{-1}\partial _\alpha A^\alpha \, . \end{aligned}$$
(304)

By construction, only physical states are in Fock’s space, so that only the potentials \(A^a\) satisfying Eqs. (303) and (304) can be one-particle wave-functions. The covariance of Fock’s space is assured by the fact that the null-plane gauge condition completely determines the gauge.

Accordingly with our quantization program, the field operators associated to the components \(A_1\) and \(A_2\) are:

$$\begin{aligned} A_1(x)= & {} (2\pi )^{-3/2}\int {\mathrm{d}}\mu (\varvec{p})\left( a_1(\varvec{p})e^{-ipx}+a_1^\dagger (\varvec{p})e^{ipx}\right) , \end{aligned}$$
(305)
$$\begin{aligned} A_2(x)= & {} (2\pi )^{-3/2}\int {\mathrm{d}}\mu (\varvec{p})\left( a_2(\varvec{p})e^{-ipx}+a_2^\dagger (\varvec{p})e^{ipx}\right) . \end{aligned}$$
(306)

They satisfy the commutation relations:

$$\begin{aligned} \left[ A_{1-}(x);A_{1+}(y)\right] =-iD_+(x-y),\quad \left[ A_{2-}(x);A_{2+}(y)\right] =-iD_+(x-y). \end{aligned}$$
(307)

From Eq. (307) we obtain:

$$\begin{aligned} \left[ A_1(x);A_1(y)\right] =-iD(x-y),\quad \left[ A_2(x);A_2(y)\right] =-iD(x-y), \end{aligned}$$
(308)

and:

$$\begin{aligned} \left[ a_1(\varvec{p});a_1^\dagger (\varvec{q})\right] =2p_-\delta (\varvec{p}-\varvec{q}),\quad \left[ a_2(\varvec{p});a_2^\dagger (\varvec{q})\right] =2p_-\delta (\varvec{p}-\varvec{q}). \end{aligned}$$
(309)

As in the fermion field case, it will be useful to write an expression for the complete vector field, for which we use the second of Eqs. (304), from that the component \(A^-(x)\) can be found:

$$\begin{aligned} A^-(x)=-(2\pi )^{-3/2}\int \frac{{\mathrm{d}}\mu (\varvec{p})}{p_-}\left[ \left( p_1a_1(\varvec{p})+p_2a_2(\varvec{p})\right) e^{-ipx}+\left( p_1a_1^\dagger (\varvec{p})+p_2a_2^\dagger (\varvec{p})\right) e^{ipx}\right] , \end{aligned}$$
(310)

where we have identified the “−” component of the polarization vectors \(\varepsilon _{1,2}(\varvec{p})^-\). For completeness we give the expression of the polarization vectors, which can be found in a classical analysis:

$$\begin{aligned} \varepsilon _1(\varvec{p})^a=\left( 0;1;0;-\frac{p_1}{p_-}\right) ,\quad \varepsilon _2(\varvec{p})^a=\left( 0;0;1;-\frac{p_2}{p_-}\right) , \end{aligned}$$
$$\begin{aligned} \varepsilon _+(\varvec{p})^a=\left( 1;-\frac{p_1}{p_-};-\frac{p_2}{p_-};\frac{p_\perp ^2}{2p_-^2}\right) ,\quad \varepsilon _-(\varvec{p})^a=\left( 0;0;0;1\right) . \end{aligned}$$
(311)

Being that way, we can write the quantized field operator:

$$\begin{aligned} A^a(x)=(2\pi )^{-3/2}\sum _{\lambda =1,2}\int {\mathrm{d}}\mu (\varvec{p})\varepsilon _\lambda (\varvec{p})^a\left( a_\lambda (\varvec{p})e^{-ipx}+a^\dagger _\lambda (\varvec{p})e^{ipx}\right) . \end{aligned}$$
(312)

With the aid of the commutation relations of Eq. (309) we can already calculate the commutation distribution of the quantized field operators in Eq. (312):

$$\begin{aligned} \left[ A^a(x);A^b(y)\right]&=(2\pi )^{-3}\int {\mathrm{d}}\mu (\varvec{p})\sum _{\lambda =1,2}\varepsilon _\lambda (\varvec{p})^a\varepsilon _\lambda (\varvec{p})^b\left. \left( e^{-ip(x-y)}-e^{ip(x-y)}\right) \right| _{p_+=E}\nonumber \\&=(2\pi )^{-3}\int d^4p\delta (p^2)\Theta \left( p_-\right) \left( -g^{ab}+\frac{p^a\eta ^b+\eta ^a p^b}{p_-}\right) \left( e^{-ip(x-y)}-e^{ip(x-y)}\right) \nonumber \\&=-(2\pi )^{-3}\int d^4p\text {sgn}\left( p_-\right) \delta (p^2)\left( g^{ab}-\frac{p^a\eta ^b+\eta ^ap^b}{p_-}\right) e^{-ip(x-y)}. \end{aligned}$$
(313)

In these expressions, the light-like vector \(\eta ^a\) is defined as:

$$\begin{aligned} (\eta ^a):=(0;0^\perp ;1),\quad (\eta _a)=(1;0_\perp ;0). \end{aligned}$$
(314)

Therefore, defining the distribution \(D^{ab}(x)\) by:

$$\begin{aligned} \left[ A^a(x);A^b(y)\right] =:iD^{ab}(x-y), \end{aligned}$$
(315)

we found that it is:

$$\begin{aligned} D^{ab}(x)=i(2\pi )^{-3}\int d^4p\text {sgn}\left( p_-\right) \delta (p^2)\left( g^{ab}-\frac{p^a\eta ^b+\eta ^ap^b}{p_-}\right) e^{-ipx}. \end{aligned}$$
(316)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Acevedo, O.A., Pimentel, B.M. Null-plane causal perturbation theory. Eur. Phys. J. Plus 137, 287 (2022). https://doi.org/10.1140/epjp/s13360-022-02459-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-022-02459-3

Navigation