Skip to main content
Log in

Nakanishi–Kugo–Ojima quantization of general relativity in Heisenberg picture

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

Abstract

The Chern–Weil topological theory is applied to a classical formulation of general relativity in four-dimensional spacetime. Einstein–Hilbert gravitational action is shown to be invariant with respect to a novel translation (co-translation) operator up to the total derivative; thus, a topological invariant of a second Chern class exists owing to Chern–Weil theory. Using topological insight, fundamental forms can be introduced as a principal bundle of the spacetime manifold. Canonical quantization of general relativity is performed in a Heisenberg picture using the Nakanishi–Kugo–Ojima formalism in which a complete set of quantum Lagrangian and BRST transformations including auxiliary and ghost fields is provided in a self-consistent manner. An appropriate Hilbert space and physical states are introduced into the theory, and the positivity of these physical states and the unitarity of the transition matrix are ensured according to the Kugo–Ojima theorem. The nonrenormalizability of quantum gravity is reconsidered under the formulation proposed herein.

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.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. Capital Roman letters indicate indices of the structural group, and the Einstein convention is also applied on them.

  2. See Section 5 in Ref. [31].

References

  1. B.P. Abbott et al., GW151226: Observation of gravitational waves from a 22-solar-mass binary black hole coalescence. Phys. Rev. Lett. 116, 241103 (2016). https://doi.org/10.1103/PhysRevLett.116.241103

    Article  ADS  Google Scholar 

  2. G. Aad et al., Observation of a new particle in the search for the Standard Model Higgs boson with the ATLAS detector at the LHC. Phys. Lett. B 716, 1–29 (2012). https://doi.org/10.1016/j.physletb.2012.08.020arXiv:1207.7214

    Article  ADS  Google Scholar 

  3. S. Chatrchyan et al., Observation of a new boson at a mass of 125 GeV with the CMS experiment at the LHC. Phys. Lett. B 716, 30–61 (2012). https://doi.org/10.1016/j.physletb.2012.08.021arXiv:1207.7235

    Article  ADS  Google Scholar 

  4. C. Rovelli, Notes for a brief history of quantum gravity, in: Recent developments in theoretical and experimental general relativity, gravitation and relativistic field theories. in Proceedings, 9th Marcel Grossmann Meeting, MG’9, Rome, Italy, July 2-8, 2000. Pts. A-C, 2000, pp. 742–768. arXiv:gr-qc/0006061 (2000)

  5. S. Carlip, D.-W. Chiou, W.-T. Ni, R. Woodard, Quantum gravity: A brief history of ideas and some prospects. Int. J. Modern Phys. D 24(11), 1530028 (2015). https://doi.org/10.1142/S0218271815300281

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. M. Fierz, W. Pauli, On relativistic wave equations for particles of arbitrary spin in an electromagnetic field. Proc. Royal Soc. London A: Math. Phys. Eng. Sci. 173(953), 211–232 (1939). https://doi.org/10.1098/rspa.1939.0140

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. G. t Hooft, M.J.G. Veltman, One loop divergencies in the theory of gravitation. Ann. Inst. H. Poincare Phys. Theor. A20, 69–94 (1974)

  8. B.S. DeWitt, Quantum theory of gravity. 1. The canonical theory,. Phys. Rev. 160, 1113–1148 (1967). https://doi.org/10.1103/PhysRev.160.1113

    Article  ADS  MATH  Google Scholar 

  9. B.S. DeWitt, Quantum theory of gravity. 2. The manifestly covariant theory. Phys. Rev. 162, 1195–1239 (1967). https://doi.org/10.1103/PhysRev.162.1195

    Article  ADS  MATH  Google Scholar 

  10. B.S. DeWitt, Quantum theory of gravity. 3. Applications of the covariant theory,. Phys. Rev. 162, 1239–1256 (1967). https://doi.org/10.1103/PhysRev.162.1239

    Article  ADS  MATH  Google Scholar 

  11. C. Rovelli, Quantum Gravity (Cambridge University Press, Cambridge, 2004). https://doi.org/10.1017/CBO9780511755804

    Book  MATH  Google Scholar 

  12. C. Rovelli, L. Smolin, Spin networks and quantum gravity. Phys. Rev. D 52, 5743–5759 (1995). https://doi.org/10.1103/PhysRevD.52.5743arXiv:gr-qc/9505006

    Article  ADS  MathSciNet  Google Scholar 

  13. A. Perez, The spin foam approach to quantum gravity. Living Rev. Rel. 16, 3 (2013). https://doi.org/10.12942/lrr-2013-3arXiv:1205.2019

    Article  MATH  Google Scholar 

  14. M. Barenz, General Covariance and Background Independence in Quantum Gravity arXiv:1207.0340

  15. C. Krishnan, K.V.P. Kumar, A. Raju, An alternative path integral for quantum gravity. JHEP 10, 043 (2016). https://doi.org/10.1007/JHEP10(2016)043arXiv:1609.04719

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. H. Hamber, Quantum Gravitation: The Feynman Path Integral Approach (Springer, Berlin Heidelberg, 2008)

    Book  Google Scholar 

  17. Jan Ambjorn, Quantum gravity represented as dynamical triangulations. Class. Quant. Grav. 12, 2079–2134 (1995). https://doi.org/10.1088/0264-9381/12/9/002

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. Jan Ambjorn, J. Jurkiewicz, Y. Watabiki, Dynamical triangulations, a gateway to quantum gravity? J. Math. Phys 36, 6299–6339 (1995). https://doi.org/10.1063/1.531246arXiv:hep-th/9503108

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. J Ambjorn, J. Andrzej, L. Jerzy, G. Renate, Quantum Gravity via Causal Dynamical Triangulations, (2013). arXiv:1302.2173, https://doi.org/10.1007/978-3-642-41992-8_34

  20. Marko Vojinović, Causal dynamical triangulations in the spincube model of quantum gravity. Phys. Rev. D 94, 024058 (2016). https://doi.org/10.1103/PhysRevD.94.024058

    Article  ADS  MathSciNet  Google Scholar 

  21. R. Loll, Quantum gravity from causal dynamical triangulations: a review. Classical Quantum Gravity 37(1), 013002 (2019). https://doi.org/10.1088/1361-6382/ab57c7

    Article  ADS  MathSciNet  Google Scholar 

  22. J. Ambjorn, M. Carfora, A. Marzuoli, The Geometry of Dynamical Triangulations, in Lecture Notes in Physics Monographs, Springer, Berlin Heidelberg, (2009)

  23. M. Carfora, A. Marzuoli, Quantum Triangulations: Moduli Spaces, Strings, and Quantum Computing, in Lecture Notes in Physics, Springer, Berlin Heidelberg, (2012)

  24. E. Witten, Topological quantum field theory. Comm. Math. Phys. 117(3), 353–386 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  25. E. Witten, 2 + 1 dimensional gravity as an exactly soluble system. Nuclear Phys. B 311(1), 46–78 (1988). https://doi.org/10.1016/0550-3213(88)90143-5

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. J. Zanelli, Chern-Simons forms in gravitation theories. Classical and Quantum Gravity 29(13), 133001 (2012). https://doi.org/10.1088/0264-9381/29/13/133001

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Y. Kurihara, Characteristic classes in general relativity on a modified poincaré curvature bundle. J. Math. Phys. 58(9), 092502 (2017). https://doi.org/10.1063/1.4990708

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. F.W. Hehl, P. Von Der Heyde, G.D. Kerlick, J.M. Nester, General relativity with spin and torsion: foundations and prospects. Rev. Mod. Phys. 48, 393–416 (1976). https://doi.org/10.1103/RevModPhys.48.393

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Y. Ne’eman, Gravity is the gauge theory of the parallel transport modification of the poincare group, in 2nd Conference on Differential Geometrical Methods in Mathematical Physics., (1978)

  30. F.W. Hehl, J.D. McCrea, E.W. Mielke, Y. Ne’eman, Metric affine gauge theory of gravity: Field equations, Noether identities, world spinors, and breaking of dilation invariance. Phys. Rept. 258, 1–171 (1995). https://doi.org/10.1016/0370-1573(94)00111-FarXiv:gr-qc/9402012

  31. N. Nakanishi, I. Ojima, Covariant Operator Formalism of Gauge Theories and Quantum Gravity, in Lecture Notes in Physics Series, World Scientific Publishing Company, Incorporated, 1990, see also references there in (1990)

  32. N. Nakanishi, Quantum gravity and general relativity. Soryusiron Kenkyu 1, 1–8 (2009). (In Japanese)

    Google Scholar 

  33. Y. Kurihara, Stochastic metric space and quantum mechanics. J. Phys. Commun. 2(3), 035025 (2018). https://doi.org/10.1088/2399-6528/aaa851

    Article  Google Scholar 

  34. C. Becchi, A. Rouet, R. Stora, Renormalization of the Abelian Higgs–Kibble Model. Commun. Math. Phys. 42, 127–162 (1975). https://doi.org/10.1007/BF01614158

    Article  ADS  MathSciNet  Google Scholar 

  35. I. Tyutin, Gauge Invariance in Field Theory and Statistical Physics in Operator Formalism arXiv:0812.0580

  36. T. Kugo, I. Ojima, Local covariant operator formalism of non-abelian gauge theories and quark confinement problem. Prog. Theor. Phys. Suppl. 66, 1–130 (1979)

    Article  ADS  Google Scholar 

  37. T. Kugo, I. Ojima, Manifestly covariant canonical formulation of yang-mills theories physical state subsidiary conditions and physical s-matrix unitarity. Phys. Lett. B 73(4), 459–462 (1978)

    Article  ADS  Google Scholar 

  38. Y. Kurihara, Symplectic structure for general relativity and Einstein–Brillouin–Keller quantization. Classical and Quantum Gravity 37(23), 235003 (2020). https://doi.org/10.1088/1361-6382/abbc44

    Article  ADS  MathSciNet  Google Scholar 

  39. P. Frè, Gravity, a Geometrical Course: Volume 1: Development of the Theory and Basic Physical Applications, Gravity, a Geometrical Course, Springer Netherlands, (2012)

  40. Y. Kurihara, Geometrothermodynamics for black holes and de Sitter space. General Relativ. Gravit. 50(2), 20 (2018). https://doi.org/10.1007/s10714-018-2341-0

    Article  ADS  MathSciNet  MATH  Google Scholar 

  41. A. Palatini, Deduzione invariantiva delle equazioni gravitazionali dal principio di hamilton. Rendiconti del Circolo Matematico di Palermo (1884-1940) 43(1), 203–212 (2008). https://doi.org/10.1007/BF03014670

  42. M. Ferraris, M. Francaviglia, C. Reina, Variational formulation of general relativity from 1915 to 1925 “palatini’s method” discovered by einstein in 1925. General Relativ. Gravit.14(3), 243–254 (1982). https://doi.org/10.1007/BF00756060

  43. S. Bates, A. Weinstein, B.C. for Pure, A. Mathematics, A. M. Society, Lectures on the Geometry of Quantization, Berkeley mathematics lecture notes, American Mathematical Society, (1997). https://books.google.co.jp/books?id=wRWoELu0uWkC

  44. V. Nair, Quantum Field Theory: A Modern Perspective, Graduate Texts in Contemporary Physics (Springer, Berlin, 2005)

    Google Scholar 

  45. R. Utiyama, Invariant theoretical interpretation of interaction. Phys. Rev. 101, 1597–1607 (1956). https://doi.org/10.1103/PhysRev.101.1597

    Article  ADS  MathSciNet  MATH  Google Scholar 

  46. N. Nakanishi, Indefinite-metric quantum field theory of general relativity. Prog. Theor. Phys. 59, 972 (1978). https://doi.org/10.1143/PTP.59.972

    Article  ADS  MathSciNet  MATH  Google Scholar 

  47. N. Nakanishi, Indefinite metric quantum field theory of general relativity. 2. Commutation Relations. Prog. Theor. Phys. 60, 1190 (1978). https://doi.org/10.1143/PTP.60.1190

    Article  ADS  MathSciNet  MATH  Google Scholar 

  48. N. Nakanishi, Indefinite metric quantum field theory of general relativity. 3. Poincare Generators. Prog. Theor. Phys. 60, 1890 (1978). https://doi.org/10.1143/PTP.60.1890

    Article  ADS  MathSciNet  MATH  Google Scholar 

  49. N. Nakanishi, Indefinite metric quantum field theory of general relativity. 4. Background Curved Space-time. Prog. Theor. Phys. 61, 1536 (1979). https://doi.org/10.1143/PTP.61.1536

    Article  ADS  MathSciNet  MATH  Google Scholar 

  50. N. Nakanishi, Indefinite metric quantum field theory of general relativity. 5. Vierbein Formalism. Prog. Theor. Phys. 62, 779 (1979). https://doi.org/10.1143/PTP.62.779

    Article  ADS  MathSciNet  MATH  Google Scholar 

  51. N. Nakanishi, Indefinite metric quantum field theory of general relativity. 6. Commutation Relations in the Vierbein Formalism. Prog. Theor. Phys. 62, 1101 (1979). https://doi.org/10.1143/PTP.62.1101

    Article  ADS  MathSciNet  MATH  Google Scholar 

  52. N. Nakanishi, Indefinite metric quantum field theory of general relativity. 7. Supplementary Remarks. Prog. Theor. Phys. 62, 1385 (1979). https://doi.org/10.1143/PTP.62.1385

    Article  ADS  MathSciNet  MATH  Google Scholar 

  53. N. Nakanishi, Indefinite metric quantum field theory of general relativity. 8. Commutators Involving \(b_\rho \). Prog. Theor. Phys. 63, 656 (1980). https://doi.org/10.1143/PTP.63.656

    Article  ADS  MATH  Google Scholar 

  54. N. Nakanishi, Indefinite Metric Quantum Field Theory of General Relativity. 9. ‘Choral’ of Symmetries. Prog. Theor. Phys. 63, 2078 (1980). https://doi.org/10.1143/PTP.63.2078

  55. N. Nakanishi, Indefinite metric quantum field theory of general relativity. 10. Sixteen-dimensional Superspace. Prog. Theor. Phys. 64, 639 (1980). https://doi.org/10.1143/PTP.64.639

    Article  ADS  MATH  Google Scholar 

  56. N. Nakanishi, I. Ojima, Indefinite metric quantum field theory of general relativity. 11. Structure of Spontaneous Breakdown of the Superalgebra. Prog. Theor. Phys. 65, 728 (1981). https://doi.org/10.1143/PTP.65.728

    Article  ADS  MATH  Google Scholar 

  57. N. Nakanishi, I. Ojima, Indefinite metric quantum field theory of general relativity. 12. Extended Superalgebra and Its Spontaneous Breakdown. Prog. Theor. Phys. 65, 1041 (1981). https://doi.org/10.1143/PTP.65.1041

    Article  ADS  MATH  Google Scholar 

  58. N. Nakanishi, K. Yamagishi, Indefinite Metric Quantum Field Theory of General Relativity. 13. Perturbation Theoretical Approach. Prog. Theor. Phys. 65, 1719 (1981). https://doi.org/10.1143/PTP.65.1719

    Article  ADS  MATH  Google Scholar 

  59. N. Nakanishi, Indefinite Metric Quantum Field Theory of General Relativity. 14. Sixteen-dimensional Noether Supercurrents and General Linear Invariance. Prog. Theor. Phys. 66, 1843 (1981). https://doi.org/10.1143/PTP.66.1843

    Article  ADS  MATH  Google Scholar 

  60. N. Nakanishi, Manifestly covariant canonical formalism of quantum gravity-systematic presentation of the theory. Publ. Res. Inst. Math. Sci. 19(3), 1095–1137 (1983)

    Article  MathSciNet  Google Scholar 

  61. N. Nakanishi, Covariant quantization of the electromagnetic field in the landau gauge. Progress Theoret. Phys. 35(6), 1111–1116 (1966). https://doi.org/10.1143/PTP.35.1111

    Article  ADS  Google Scholar 

  62. N. Nakanishi, Remarks on the indefinite-metric quantum field theory of general relativity. Progress Theor. Phys. 59(6), 2175–2177 (1978). https://doi.org/10.1143/PTP.59.2175

    Article  ADS  Google Scholar 

  63. N. Nakanishi, A new way of describing the lie algebras encountered in quantum field theory. Progress Theor. Phys. 60(1), 284–294 (1978). https://doi.org/10.1143/PTP.60.284

    Article  ADS  MathSciNet  MATH  Google Scholar 

  64. R. Arnowitt, S. Deser, C.W. Misner, Dynamical structure and definition of energy in general relativity. Phys. Rev. 116, 1322–1330 (1959). https://doi.org/10.1103/PhysRev.116.1322

    Article  ADS  MathSciNet  MATH  Google Scholar 

  65. F. Berends, R. Gastmans, On the high-energy behaviour of Born cross sections in quantum gravity. Nuclear Phys. B 88(1), 99–108 (1975). https://doi.org/10.1016/0550-3213(75)90528-3

    Article  ADS  Google Scholar 

  66. M.H. Goroff, A. Sagnotti, A. Sagnotti, Quantum gravity at two loops. Phys. Lett. B 160, 81–86 (1985). https://doi.org/10.1016/0370-2693(85)91470-4

    Article  ADS  Google Scholar 

  67. M.H. Goroff, A. Sagnotti, The ultraviolet behavior of Einstein gravity. Nuclear Phys. B 266(3), 709–736 (1986). https://doi.org/10.1016/0550-3213(86)90193-8

    Article  ADS  Google Scholar 

  68. C. Llewellyn-Smith, High energy behaviour and gauge symmetry. Phys. Lett. B 46(2), 233–236 (1973). https://doi.org/10.1016/0370-2693(73)90692-8

    Article  ADS  Google Scholar 

  69. N. Nakanishi, Method for solving quantum field theory in the heisenberg picture. Prog. Theor. Phys. 111(3), 301 (2004). https://doi.org/10.1143/PTP.111.301

    Article  ADS  MathSciNet  MATH  Google Scholar 

  70. B. Delamotte, An Introduction to the nonperturbative renormalization group. Lect. Notes Phys. 852, 49–132 (2012). https://doi.org/10.1007/978-3-642-27320-9_2arXiv:cond-mat/0702365

    Article  ADS  MathSciNet  MATH  Google Scholar 

  71. N. Dupuis, L. Canet, A. Eichhorn, W. Metzner, J.M. Pawlowski, M. Tissier, N. Wschebor, The nonperturbative functional renormalization group and its applications arXiv:2006.04853, https://doi.org/10.1016/j.physrep.2021.01.001

  72. J.F. Plebański, On the separation of einsteinian substructures. J. Math. Phys. 18(12), 2511–2520 (1977). https://doi.org/10.1063/1.523215

    Article  ADS  MathSciNet  MATH  Google Scholar 

  73. G.T. Horowitz, Exactly soluble diffeomorphism invariant theories. Comm. Math. Phys. 125(3), 417–437 (1989). https://doi.org/10.1007/BF01218410

    Article  ADS  MathSciNet  MATH  Google Scholar 

  74. D. Birmingham, M. Blau, M. Rakowski, G. Thompson, Topological field theory. Phys. Report 209, 129–340 (1991). https://doi.org/10.1016/0370-1573(91)90117-5

    Article  ADS  MathSciNet  MATH  Google Scholar 

  75. K. Krasnov, Plebański formulation of general relativity: a practical introduction. General Relat. Gravit. 43(1), 1–15 (2011). https://doi.org/10.1007/s10714-010-1061-x

    Article  ADS  MATH  Google Scholar 

  76. Y. Kurihara, Gravitational theories with topological invariant. Phys. Astron. J. 2(3), 361–363 (2018). https://doi.org/10.15406/paij.2018.02.00110

    Article  Google Scholar 

  77. F. Girelli, H. Pfeiffer, Higher gauge theory-differential versus integral formulation. J. Math. Phys. 45(10), 3949–3971 (2004). https://doi.org/10.1063/1.1790048

    Article  ADS  MathSciNet  MATH  Google Scholar 

  78. S. Gielen, D. Oriti, Classical general relativity as BF-Plebanski theory with linear constraints. Class. Quant. Grav. 27, 185017 (2010). https://doi.org/10.1088/0264-9381/27/18/185017arXiv:1004.5371

    Article  ADS  MathSciNet  MATH  Google Scholar 

  79. M. Celada, D. González, M. Montesinos, BF gravity. Classical and Quantum Gravity 33(21), 213001 (2016). https://doi.org/10.1088/0264-9381/33/21/213001

    Article  ADS  MathSciNet  MATH  Google Scholar 

  80. R.D. Pietri, L. Freidel, SO(4) Plebański action and relativistic spin-foam model. Classical and Quantum Gravity 16(7), 2187 (1999). https://doi.org/10.1088/0264-9381/16/7/303

    Article  ADS  MathSciNet  MATH  Google Scholar 

  81. I.V. Kanatchikov, De Donder-Weyl Hamiltonian formulation and precanonical quantization of vielbein gravity. Journal of Physics: Conference Series 442, 012041 (2013) https://doi.org/10.1088/1742-6596/442/1/012041. 10.1088%2F1742-6596%2F442%2F1%2F012041

Download references

Acknowledgements

I appreciate the kind hospitality of all members of the theory group of Nikhef, particularly Prof. J. Vermaseren and Prof. E. Laenen. A major part of this study was conducted during my stay at Nikhef in 2017. I would also like to thank Dr. Y. Sugiyama for his continuous encouragement and fruitful discussions. The author would like to thank Enago for the English language review.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yoshimasa Kurihara.

Appendix

Appendix

1.1 A Co-translation operator and co-Poincaré symmetry

In this Appendix, two remarks related to co-translation symmetry are given based on previous studies [27, 38] by the author. They are essentially important to construct the Chern–Weil gravitational theory introduced in this study.

Suppose \(P_a\) be a generator of translation group \(T^4\) on four-dimensional local Lorentz manifold \(\mathcal{M}\). Translation operator \(\delta _T\) acts on vierbein and spin forms, respectively, as [26]

$$\begin{aligned} \delta _T{{\mathfrak {e}}}^a=\mathrm{d}\xi ^a+{{\mathfrak {w}}}^a_{~{\circ }}\xi ^{\circ }=\mathrm{d}_{{\mathfrak {w}}}\xi ^a&\mathrm{and}&\delta _T{{\mathfrak {w}}}^{ab}=0, \end{aligned}$$

where \(\xi ^a\) is a local vector to characterize the translation. Zanelli showed the Einstein–Hilbert Lagrangian in a three-dimensional spacetime had a translation invariant owing to an accidental opportunity in the three dimensional space [26]. The co-translation is introduced to realize an extension of a translation symmetry in the Einstein–Hilbert Lagrangian in a four-dimensional spacetime.

A generator of the co-translation is defined using the translation and contraction operators as follows:

Definition A.1

(co-translation): Co-translation generator \(P_{ab}\) is defined as

$$\begin{aligned} P_{ab}= & {} P_a\iota _b, \end{aligned}$$

where \(\iota _\bullet \) is a contraction operator with respect to local vector \(\xi ^\bullet \) in \({T\mathcal{M}}\), whose representation is provided using the trivial basis as

$$\begin{aligned} \iota _{a}= & {} \iota _{\xi ^a},~~\xi ^a=\eta ^{ab}{{\mathcal {E}}}^{\mu }_b\partial _\mu . \end{aligned}$$

The co-translation operator is represented as \(\delta _{CT}=\xi ^a\times \delta _T\;\iota _a\), where “\(\times \)” is an antisymmetric tensor product of the two vectors such that

$$\begin{aligned} \xi ^a\times \xi ^b=\xi ^a\otimes \xi ^b-\xi ^b\otimes \xi ^a=-\xi ^b\times \xi ^a \end{aligned}$$

A contraction is a map \(\iota :\Omega ^p({T^*\mathcal{M}}) \rightarrow \Omega ^{p-1}({T^*\mathcal{M}})\). One can obtain a relation,

$$\begin{aligned} \iota _{a}{{\mathfrak {e}}}^b={{\mathcal {E}}}^{\mu }_a{{\mathcal {E}}}_\nu ^b\delta ^\nu _\mu =\delta ^b_a,&\mathrm{and}&\delta _{CT}=\xi ^{\circ }\times \delta _T\left( \iota _{\circ }\;{{\mathfrak {e}}}^a\right) = \xi ^{\circ }\times \left( \delta _T\delta ^a_{\circ }\right) =\xi ^a, \end{aligned}$$

where \(\delta _T\delta ^a_b=\delta ^a_b\) is used owing to its translation invariance in \({T\mathcal{M}}\). Above-mentioned relation is independent of the choice of bases in \({T\mathcal{M}}\). A projection manifold and a projection bundle can be introduced using a contraction operator as follows: let \(\mathcal{M}_{\perp a}\subset \mathcal{M}\) be a three-dimensional submanifold of \(\mathcal{M}\), such that \(\iota _a{{\mathfrak {a}}}=0\) for \({{\mathfrak {a}}}\) with \(\Omega ^1(\mathcal{M}_{\perp a})\ni {{\mathfrak {a}}}\ne 0\). The trivial frame bundles in \(T\mathcal{M}_{\perp a}\) and \(T^*\mathcal{M}_{\perp a}\) are regarded as sub-bundles of \(T\mathcal{M}\) and \(T^*\mathcal{M}\), respectively. If \(\mathcal{M}\) has Poincaré symmetry ISO(1, 3), the submanifold \(\mathcal{M}_{\perp a}\) has the ISO(1, 2) or the ISO(3) symmetry. Therefore, connection \({{\mathfrak {A}}}_a\) and curvature \({\mathfrak {F}}_a=\mathrm{d}{{\mathfrak {A}}}_a+{{\mathfrak {A}}}_a\wedge {{\mathfrak {A}}}_a\) are provided in \(\mathcal{M}_{\perp a}\) as sub-bundles of \({{\mathfrak {A}}}\) and \({\mathfrak {F}}\) on \(\mathcal{M}\), respectively. Using an equivalence relation \(\sim _a\) such as \({{\mathfrak {a}}}\sim _a{{\mathfrak {b}}}\Leftrightarrow \iota _a{{\mathfrak {a}}}=\iota _a{{\mathfrak {b}}}\), quotient bundles \({\widetilde{{{\mathfrak {A}}}}}_a={{\mathfrak {A}}}/{\sim _a}\) and \({\widetilde{{\mathfrak {F}}}}_a={\mathfrak {F}}/{\sim _a}\) can be introduced, where \({{\mathfrak {a}}}\) and \({{\mathfrak {b}}}\) are any p-forms on \(T^*\mathcal{M}\). It is clear that \({{\mathfrak {A}}}_a\simeq {\widetilde{{{\mathfrak {A}}}}}_a\subset {{\mathfrak {A}}}\) and \({\mathfrak {F}}_a\simeq {\widetilde{{\mathfrak {F}}}}_a\subset {\mathfrak {F}}\).

The co-translation acts on the fundamental forms as follows:

$$\begin{aligned} \delta _{CT}\left( {{\mathfrak {S}}}_{ab}\right)= & {} \frac{1}{2}\epsilon _{abcd}\left( \xi ^c\times \mathrm{d}_{{{\mathfrak {w}}}}\xi ^d-\xi ^\mathrm{d}\times \mathrm{d}_{{{\mathfrak {w}}}}\xi ^c,\right) \end{aligned}$$
(52)
$$\begin{aligned} \delta _{CT} {{\mathfrak {w}}}^{ab}= & {} 0. \end{aligned}$$
(53)

Remark A.2

(Remark 3.2 in Ref. [27]) The Einstein–Hilbert Lagrangian given in (11) with \(\Lambda _c=0\) is invariant under the co-Poincaré transformation up to a total derivative.

Proof

The gravitational Lagrangian with \(\Lambda _c=0\) is considered. The invariance under the local SO(1, 3) symmetry of the Einstein–Hilbert Lagrangian is trivial from its definition. A co-translation operator \(\delta _{CT}\) induces a transformation on the Einstein–Hilbert Lagrangian \({{\mathfrak {L}}}_G\) as

$$\begin{aligned} (\hbar \kappa )\delta _{CT}{{\mathfrak {L}}}_G= & {} \delta _{CT}\left( \frac{1}{2} \epsilon _{{\circ \circ }{\circ \circ }}{\mathfrak {R}}^{\circ \circ }\wedge {{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\right) =\epsilon _{{\circ \circ }{\circ \circ }}{\mathfrak {R}}^{\circ \circ }\wedge \left( \xi ^{\circ }\times \mathrm{d}_{{{\mathfrak {w}}}}\xi ^{\circ }\right) \nonumber \\= & {} \mathrm{d}_{{{\mathfrak {w}}}}\left\{ \epsilon _{{\circ \circ }{\circ \circ }}{\mathfrak {R}}^{\circ \circ }(\xi ^{\circ }\times \xi ^{\circ }) \right\} , \end{aligned}$$
(54)

where (52), (53) and the Bianchi identity \(\mathrm{d}_{{{\mathfrak {w}}}}{\mathfrak {R}}=0\) are used. Owing to the definition of the covariant derivative (2), \(\mathrm{d}_{{{\mathfrak {w}}}}\{\cdots \}=\mathrm{d}\{\cdots \}\) is maintained in (54); thus, the co-translation operator only induces a total derivative term on the Einstein–Hilbert Lagrangian without the cosmological constant. \(\square \)

Remark A.3

(Theorem 4.2 in Ref. [27] and Remark 2.1 in Ref. [38]) The Einstein–Hilbert Lagrangian without the cosmological constant is defined using the curvature of the co-Poincaré bundle as (24). It has a topological invariant as the second Chern class \(c_2({\mathfrak {F}}_{c\mathrm {P}}):\)

$$\begin{aligned} {{\mathfrak {L}}}= \frac{8\pi ^2}{\kappa }c_2({\mathfrak {F}}_{c\mathrm {P}})\in {\mathbb {R}}\otimes H^4(\mathcal{M},{\mathbb {Z}}). \end{aligned}$$

Proof

The right-hand side of (24) is expressed using the structure constants as

$$\begin{aligned} \mathrm {Tr}\left[ {\mathfrak {F}}_{c\mathrm {P}}\wedge {\mathfrak {F}}_{c\mathrm {P}}\right]:= & {} \frac{1}{2} {{\mathcal {F}}}^K_{~IJ}\;{\mathfrak {F}}_{c\mathrm {P}}^I\wedge {\mathfrak {F}}_{c\mathrm {P}}^J\;T_K. \end{aligned}$$

Owing to the definition of the co-Poincaré curvature (23) and the Lie algebra of the co-Poincaré symmetry, a direct calculation with a trivial basis yields

$$\begin{aligned} \mathrm {Tr}\left[ {\mathfrak {F}}_{c\mathrm {P}}\wedge {\mathfrak {F}}_{c\mathrm {P}}\right]= & {} \mathrm {Tr}\left[ {\mathfrak {R}}^{ac}\wedge \mathrm{d}_{{\mathfrak {w}}}\left( P_{a}^{b}\;{{\mathfrak {S}}}_{cb}\right) \right] = \mathrm{d}_{{\mathfrak {w}}}\left( \mathrm {Tr}\left[ P_{a}^{b}\right] {\mathfrak {R}}^{ac}\wedge {{\mathfrak {S}}}_{cb}\right) = -\mathrm{d}_{{\mathfrak {w}}} \left( P_{{\circ }}^{{\circ }} \;{\mathfrak {R}}\wedge {{\mathfrak {S}}}\right) , \end{aligned}$$

where Bianchi identity \(\mathrm{d}_{{\mathfrak {w}}}{\mathfrak {R}}=0\) and co-translation invariance \(P\;({\mathfrak {R}})=0\) are used. For the last equality, we note that \(\mathrm {Tr}\left[ {\mathfrak {R}}^{ac}\wedge {{\mathfrak {S}}}_{cb}\right] =-\delta ^a_b({\mathfrak {R}}\wedge {{\mathfrak {S}}})\). The Einstein–Hilbert gravitational Lagrangian is co-translation-invariant up to total derivative as given as Remark A.2; thus, P is removed from the last expression and it is embedded in a five-dimensional line bundle of \({\mathscr {M}}_5:={\mathscr {M}}\otimes {\mathbb {R}}\), whose boundary is given as \(\partial {\mathscr {M}}_5=\Sigma \). Consequently,

$$\begin{aligned} -\int _{\Sigma }\mathrm {Tr}\left[ {\mathfrak {F}}_{c\mathrm {P}}\wedge {\mathfrak {F}}_{c\mathrm {P}}\right]= & {} \int _{{\mathscr {M}}_5}\mathrm{d}_{{\mathfrak {w}}}\left( {\mathfrak {R}}\wedge {{\mathfrak {S}}}\right) = \int _{{\mathscr {M}}_5}\mathrm{d}\left( {\mathfrak {R}}\wedge {{\mathfrak {S}}}\right) = \int _{\partial {\mathscr {M}}_5=\Sigma }{\mathfrak {R}}\wedge {{\mathfrak {S}}}. \end{aligned}$$

A surface term is eliminated owing to a boundary condition. We note that

$$\begin{aligned} \mathrm{d}_{{\mathfrak {w}}}\left( {\mathfrak {R}}\wedge {{\mathfrak {S}}}\right) = \mathrm{d}\left( {\mathfrak {R}}\wedge {{\mathfrak {S}}}\right) +{{\mathfrak {w}}}\wedge {\mathfrak {R}}\wedge {{\mathfrak {S}}}-{\mathfrak {R}}\wedge {{\mathfrak {S}}}\wedge {{\mathfrak {w}}}=\mathrm{d}\left( {\mathfrak {R}}\wedge {{\mathfrak {S}}}\right) , \end{aligned}$$

owing to the definition of a covariant derivative.

The second Chern class with respect to the co-Poincaré curvature is obtained as

$$\begin{aligned} c_2({\mathfrak {F}}_{c\mathrm {P}}):= & {} \frac{1}{8\pi ^2}\left( \mathrm {Tr}\left[ {\mathfrak {F}}_{c\mathrm {P}}\right] ^2 -\mathrm {Tr}\left[ {\mathfrak {F}}_{c\mathrm {P}}\wedge {\mathfrak {F}}_{c\mathrm {P}}\right] \right) , \end{aligned}$$

and it has a homology class as \(c_2({\mathfrak {F}}_{c\mathrm {P}})\in H^4(\mathcal{M},{\mathbb {R}})\subset \mathrm {Im}\;H^4(\mathcal{M},{\mathbb {Z}})\) owing to the Chern–Weil theory. The co-Poincaré curvature \({\mathfrak {F}}_{c\mathrm {P}}\) is traceless; thus, the remark is maintained. \(\square \)

This result suggests that appropriate fundamental forms of the symplectic geometry for general relativity can be identified as \(({{\mathfrak {w}}},{{\mathfrak {S}}})\). This choice of the canonical pair of general relativity is equivalent to that proposed by Kanatchikov [81] on the basis of the de Donder–Weyl Hamiltonian theory.

1.2 B Proof of nilpotency

figure a

The coordinate vectors are fundamental vectors on \(T{\mathscr {M}}\). Nilpotent can be confirmed as

$$\begin{aligned}&\delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ x^{\mu }\right] \right] =\delta _{\mathrm {BRST}}\left[ \chi ^\mu \right] =\delta _{\mathrm {BRST}}\left[ g^{\mu \nu }\chi _\nu \right] ,\\&\quad = g^{\mu \rho }\left( \partial _\rho \chi ^\nu \right) \chi _\nu + g^{\rho \nu }\left( \partial _\rho \chi ^\mu \right) \chi _\nu -g^{\mu \nu }g_{\nu \rho }\left( \partial ^\rho \chi ^\sigma + \partial ^\sigma \chi ^\rho \right) \chi _\sigma ,\\&\quad = \left( \partial ^\mu \chi ^\nu \right) \chi _\nu + \left( \partial ^\nu \chi ^\mu \right) \chi _\nu -\left( \partial ^\mu \chi ^\sigma \right) \chi _\sigma - \left( \partial ^\sigma \chi ^\mu \right) \chi _\sigma =0. \end{aligned}$$
figure b

Starting from the BRST transformation of the metric tensor (29), nilpotent is provided as

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ g_{\mu \nu }\right] \right]= & {} \delta _{\mathrm {BRST}}\left[ -g_{\mu \rho }\partial _\nu \chi ^\rho \right] +\delta _{\mathrm {BRST}}\left[ \mu \leftrightarrow \nu \right] ,\\= & {} \left\{ -\delta _{\mathrm {BRST}}\left[ g_{\mu \rho }\right] \partial _\nu \chi ^\rho +g_{\mu \rho }\left( \partial _\nu \delta _{\mathrm {BRST}}\left[ x^\sigma \right] \right) \partial _\sigma \chi ^\rho \right\} +\left\{ \mu \leftrightarrow \nu \right\} ,\\= & {} \left\{ g_{\mu \sigma }\left( \partial _\rho \chi ^\sigma \right) \partial _\nu \chi ^\rho + g_{\mu \rho }\left( \partial _\nu \chi ^\sigma \right) \partial _\sigma \chi ^\rho \right\} +\left\{ \mu \leftrightarrow \nu \right\} =0, \end{aligned}$$

where anticommutativity of the ghost filed is used.

figure c

Since the ghost field has two parts, nilpotent is checked separately. First, nilpotent of the \(\chi _\mu \) is shown as

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ \chi _\mu \right] \right]= & {} \delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ g_{\mu \nu }\chi ^\nu \right] \right] =\delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ g_{\mu \nu }\right] \right] \chi ^\nu =0, \end{aligned}$$

where \(\delta _{\mathrm {BRST}}[\chi ^\mu ]=0\) and nilpotent of the metric tensor are used. Direct calculation from (27) gives the same result, too. The second part becomes

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ \chi ^a_{~b}\right] \right]= & {} \delta _{\mathrm {BRST}}\left[ \chi ^a_{~c}\chi ^{c}_{~b}\right] = \chi ^a_{~c_2}\chi ^{c_2}_{~~c_1}\chi ^{c_1}_{~~b}-\chi ^a_{~c_1}\chi ^{c_1}_{~~c_2}\chi ^{c_2}_{~~b}=0, \end{aligned}$$

due to anticommutativity of the ghost field.

Next, a tensor \(\partial _{\mu }\chi ^{\nu }\) is also nilpotent as

$$\begin{aligned}&\delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ \partial _{\mu }\chi ^{\nu }\right] \right] = -\delta _{\mathrm {BRST}}\left[ \partial _{\mu }\chi ^{\rho }\partial _{\rho }\chi ^{\nu }\right] \\&\quad =-\partial _{\mu }\chi ^{\rho _1}\partial _{\rho _1}\chi ^{\rho _2}\partial _{\rho _2}\chi ^{\nu } +\partial _{\mu }\chi ^{\rho _1}\partial _{\rho _1}\chi ^{\rho _2}\partial _{\rho _2}\chi ^{\nu }=~0. \end{aligned}$$
figure d

Nilpotent of the vierbein form is provided as

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ {{\mathfrak {e}}}^a\right] \right]= & {} \delta _{\mathrm {BRST}}\left[ {{\mathfrak {e}}}^b\chi ^a_{~b}\right] =~ {{\mathfrak {e}}}^{b_1}\chi ^{b_2}_{~~b_1}\chi ^a_{~b_2}+{{\mathfrak {e}}}^{b_2}\chi ^a_{~b_1}\chi ^{b_1}_{~~b_2}=0, \end{aligned}$$

due to anticommutativity of the ghost field.

figure e

One can trace the same calculation as a case of the vierbein form due to \(\delta _{\mathrm {BRST}}\left[ \mathrm{d}\chi ^{ab}\right] =0\). Detailed calculations are omitted here.

figure f

The volume form is global scalar, and their BRST transformation is expected to vanish, which can be confirmed as

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ {\mathfrak {v}}\right]= & {} \frac{1}{4!}\epsilon _{{\circ \circ \circ \circ }} \delta _{\mathrm {BRST}}\left[ {{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\right] \\= & {} \frac{1}{3!}\epsilon _{a_1{\circ }{\circ \circ }} \delta _{\mathrm {BRST}}\left[ \chi ^{a_1}_{~~a_2}{{\mathfrak {e}}}^{a_2}\wedge {{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\right] ~=~0, \end{aligned}$$

due to \({{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\propto \epsilon ^{{\circ \circ \circ \circ }}\) and \(\chi ^{a_1}_{~~a_2}=0\) when \(a_1=a_2\).

figure g

The BRST transformation of the surface form is provided as

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ {{\mathfrak {S}}}_{ab}\right]= & {} \frac{1}{2}\epsilon _{abc_1c_2}\delta _{\mathrm {BRST}}\left[ {{\mathfrak {e}}}^{c_1}\wedge {{\mathfrak {e}}}^{c_2}\right] =\epsilon _{abc_1c_2}\chi ^{c_1}_{~c_3}{{\mathfrak {e}}}^{c_3}\wedge {{\mathfrak {e}}}^{c_2}~ \left( =~{{\mathfrak {c}}}^{\circ }\wedge {\overline{{{\mathfrak {e}}}}}_{ab{\circ }}\right) . \end{aligned}$$

Applying the BRST-transformation on it again, one can get

$$\begin{aligned}&\delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ {{\mathfrak {S}}}_{ab}\right] \right] ~=~ \epsilon _{abc_1c_2}\delta _{\mathrm {BRST}}\left[ \chi ^{c_1}_{~c_3}{{\mathfrak {e}}}^{c_3}\wedge {{\mathfrak {e}}}^{c_2}\right] ,\\&\quad =\epsilon _{abc_1c_2}\Bigl \{ \chi ^{c_1}_{~c_4}\chi ^{c_4}_{~c_3}{{\mathfrak {e}}}^{c_3}\wedge {{\mathfrak {e}}}^{c_2}- \chi ^{c_1}_{~c_3}\chi ^{c_3}_{~c_4}{{\mathfrak {e}}}^{c_4}\wedge {{\mathfrak {e}}}^{c_2}- \chi ^{c_1}_{~c_3}\chi ^{c_2}_{~c_4}{{\mathfrak {e}}}^{c_3}\wedge {{\mathfrak {e}}}^{c_4} \Bigr \}=0, \end{aligned}$$

because first term is the same as the second term and the third term is symmetric with \(c_1\) and \(c_2\) exchange.

figure h

The BRST transformation of \({{\mathfrak {c}}}^a\) is given by

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ {{\mathfrak {c}}}^a\right]= & {} \delta _{\mathrm {BRST}}\left[ \chi ^a_{~b}~{{\mathcal {E}}}^b_\mu ~\mathrm{d}x^\mu \right] ,\\= & {} \chi ^a_{~b_1}\chi ^{b_1}_{~b_2}{{\mathcal {E}}}^{b_2}_\mu \mathrm{d}x^\mu - \chi ^a_{~b_1}~{{\mathcal {E}}}^{b_2}_\mu \chi ^{b_2}_{~b_1}\mathrm{d}x^\mu +\chi ^a_{~b}~\left( \partial _\mu \chi ^\nu \right) {{\mathcal {E}}}^b_\nu ~\mathrm{d}x^\mu -\chi ^a_{~b}~{{\mathcal {E}}}^b_\mu \mathrm{d}\chi ^\mu =0, \end{aligned}$$
figure i

Nilpotent of other forms is trivial, and the proof is omitted here.

figure j

The quantum Lagrangian must be the BRST-null. Gauge-fixing and Fadeef–Popov Lagrangians are constructed to satisfy the BRST-null condition in Sect. 6-2. Nilpotent of only the gravitational Lagrangian is given here. The BRST transformation of the gravitational Lagrangian is provided as

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ {{\mathfrak {L}}}_G\right]= & {} \frac{1}{2}\delta _{\mathrm {BRST}}\left[ \left( \mathrm{d}{{\mathfrak {w}}}^{\circ \circ }+{c_{gr}}\;{{\mathfrak {w}}}^{\circ }_{~\star }\wedge {{\mathfrak {w}}}^{\star {\circ }}\right) \wedge {{{\mathfrak {S}}}}_{\circ \circ }-\frac{\Lambda }{3!}{\mathfrak {v}}\right] . \end{aligned}$$

The BRST transformation for the volume form is vanished by itself. For the derivative term,

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ \mathrm{d}{{\mathfrak {w}}}^{\circ \circ }\wedge {{{\mathfrak {S}}}}_{\circ \circ }\right]= & {} \epsilon _{abc_2c_3}\chi _{~c_1}^{b}\mathrm{d}{{\mathfrak {w}}}^{ac_1}\wedge {{\mathfrak {e}}}^{c_2}\wedge {{\mathfrak {e}}}^{c_3}+ \epsilon _{abc_2c_3}{{\mathfrak {w}}}^{ac_1}\wedge \mathrm{d}\chi _{~c_1}^{b}\wedge {{\mathfrak {e}}}^{c_2}\wedge {{\mathfrak {e}}}^{c_3}\\&+\epsilon _{abc_1c_2}\chi ^{c_1}_{~c_3}~\mathrm{d}{{\mathfrak {w}}}^{ab}\wedge {{\mathfrak {e}}}^{c_3}\wedge {{\mathfrak {e}}}^{c_2}\\= & {} 2~{{\mathfrak {w}}}^{ac_1}\wedge \mathrm{d}\chi ^{b}_{~c_1}\wedge {{\mathfrak {S}}}_{ab}, \end{aligned}$$

where first- and third-terms are cancelled each other. Remnant term is transformed as

$$\begin{aligned}&\delta _{\mathrm {BRST}}\left[ {{\mathfrak {w}}}^{\circ }_{~\star }\wedge {{\mathfrak {w}}}^{\star {\circ }}\wedge {{{\mathfrak {S}}}}_{\circ \circ }\right] \\&\quad =\epsilon _{abc_2c_3}\chi ^{c_2}_{~c_4}{{\mathfrak {w}}}^{ac_1}\wedge {{\mathfrak {w}}}_{c_1}^{~~b}\wedge {{\mathfrak {e}}}^{c_4}\wedge {{\mathfrak {e}}}^{c_3}+ \epsilon _{abc_3c_4}\chi ^{c_2}_{~c_1}{{\mathfrak {w}}}^{ac_1}\wedge {{\mathfrak {w}}}_{c_2}^{~~b}\wedge {{\mathfrak {e}}}^{c_3}\wedge {{\mathfrak {e}}}^{c_4}\\&\qquad +\epsilon _{abc_3c_4}\chi ^{b}_{~c_2}{{\mathfrak {w}}}^{ac_1}\wedge {{\mathfrak {w}}}_{c_1}^{~~c2}\wedge {{\mathfrak {e}}}^{c_3} \wedge {{\mathfrak {e}}}^{c_4}-2{{c_{gr}}}^{-1}{{\mathfrak {w}}}^{ac_1}\wedge \mathrm{d}\chi ^b_{~c_1}\wedge {{\mathfrak {S}}}_{ab},\\&\quad =-2{{c_{gr}}}^{-1}{{\mathfrak {w}}}^{ac_1}\wedge \mathrm{d}\chi ^b_{~c_1}\wedge {{\mathfrak {S}}}_{ab}. \end{aligned}$$

In the r.h.s of the first line, the second term is zero as itself, and first- and third-terms are cancelled each other. Therefore, one can confirm \(\delta _{\mathrm {BRST}}\left[ {{\mathfrak {L}}}_G\right] =0\) after summing up all terms.

If we use a following remake, we can give simpler proofs for above forms.

figure k

If both of two fields, \(\alpha \) and \(\beta \), are nilpotent, \(\alpha \beta \) is also nilpotent.

Proof

If a field X is nilpotent, signatures of the Leibniz rule satisfy \(\epsilon _{X}=-\epsilon _{\delta X}\) due to \(\delta _{\mathrm {BRST}}[\delta _{\mathrm {BRST}}[X]]=0\) and (28), where \(\epsilon _{X}\) (\(\epsilon _{\delta X}\)) is a signature of X (\(\delta _{\mathrm {BRST}}[X]\)), respectively. Therefore,

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ \delta _{\mathrm {BRST}}\left[ \alpha \beta \right] \right]= & {} \epsilon _{\alpha }\delta _{\mathrm {BRST}}\left[ \alpha \right] \delta _{\mathrm {BRST}}\left[ \beta \right] + \epsilon _{\delta \alpha }\delta _{\mathrm {BRST}}\left[ a\right] \delta _{\mathrm {BRST}}\left[ \beta \right] =0. \end{aligned}$$

\(\square \)

1.3 C Equations of motion

From the classical Lagrangian form, the torsionless condition and the Einstein equation are obtained as equations of motion by requiring a stationary condition for variation of action. The same procedure can be used to obtain Euler–Lagrange equations from the quantum Lagrangian. Here, the quantum Lagrangian is summarized as

$$\begin{aligned}&{{{\mathfrak {L}}}_{QG}}={{\mathfrak {L}}}_G+{{{\mathfrak {L}}}_{GF}}+{{{\mathfrak {L}}}_{FP}},~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~(31)\\&\quad \left\{ \begin{array}{llc} {{\mathfrak {L}}}_G&{}=&{}~~~~~~\frac{1}{2}\left( {\mathfrak {R}}^{\circ \circ }-\frac{\Lambda }{3!}{\overline{{{\mathfrak {S}}}}}^{\circ \circ }\right) \wedge {{{\mathfrak {S}}}}_{\circ \circ },~~~~~~~~~~~~~~~~~~~~~~(11)\\ {{{\mathfrak {L}}}_{GF}}&{}=&{} -\frac{1}{2}\left( \mathrm{d}{{\mathfrak {b}}}^{\circ \circ }+\alpha {{\mathfrak {b}}}^{\circ }_{~\star }\wedge {{\mathfrak {b}}}^{\star {\circ }} \right) \wedge {{\mathfrak {S}}}_{\circ \circ },~~~~~~~~~~~~~~~~~~~~(32)\\ {{{\mathfrak {L}}}_{FP}}&{}=&{} -\frac{i}{2}\left( d{\tilde{{{\mathfrak {c}}}}}^{\circ \circ }+ \alpha {\tilde{{{\mathfrak {c}}}}}^{{\circ }}_{~\star }\wedge {{\mathfrak {b}}}^{\star {\circ }}\right) \wedge {{\mathfrak {c}}}^\star \wedge {\overline{{{\mathfrak {e}}}}}_{\star {\circ \circ }},~~~~~~~~~~~~~~~(33) \end{array} \right. \end{aligned}$$

Euler–Lagrange equations are obtained as follows:

figure l
$$\begin{aligned} {{\mathfrak {T}}}^a= & {} \mathrm{d}{{\mathfrak {e}}}^a+{c_{gr}}\;{{\mathfrak {w}}}^a_{~{\circ }}\wedge {{\mathfrak {e}}}^{\circ }=0. \end{aligned}$$

This is the torsionless condition in the same form as the classical Lagrangian.

figure m
$$\begin{aligned}&\frac{1}{2}\overline{\left( {\mathfrak {R}}^{\circ \circ }\wedge {{\mathfrak {e}}}^{\circ }\right) }_a-\Lambda {\mathfrak {V}}_a -\frac{1}{2}\left( \mathrm{d}{{\mathfrak {b}}}^{\circ \circ }+\alpha {{\mathfrak {b}}}^{\circ }_{~\star }\wedge {{\mathfrak {b}}}^{\star {\circ }} \right) \wedge {\overline{{{\mathfrak {e}}}}}_{a{\circ \circ }}\nonumber \\&\quad -\frac{i}{2}\left( d{\tilde{{{\mathfrak {c}}}}}^{\circ \circ }+\alpha {\tilde{{{\mathfrak {c}}}}}^{\circ }_{~\star }\wedge {{\mathfrak {b}}}^{\star {\circ }}\right) \wedge {\overline{{{\mathfrak {c}}}}}_{a{\circ \circ }} =0, \end{aligned}$$
(55)

where \({\mathfrak {V}}_a=\epsilon _{ab_1b_2b_3}{{\mathfrak {e}}}^{b_1}\wedge {{\mathfrak {e}}}^{b_2}\wedge {{\mathfrak {e}}}^{b_3}/3!\). The first two terms give the Einstein equation without matter fields. Third- and fourth-terms newly appeared from the gauge-fixing and Faddeev–Popov Lagrangian forms.

figure n
$$\begin{aligned} \mathrm{d}{{\mathfrak {S}}}_{ab}-\alpha \left( {{\mathfrak {b}}}^{\circ }_a\wedge {{\mathfrak {S}}}_{{\circ }b}-i{\tilde{{{\mathfrak {c}}}}}^{\circ }_a\wedge {{\mathfrak {c}}}^{\circ }\wedge {\overline{{{\mathfrak {e}}}}}_{{\circ \circ }b} \right)= & {} 0. \end{aligned}$$

We note that \({{\mathfrak {b}}}\) and \(\delta _{{\mathfrak {b}}}\) are anticommute each other, and the variation operator is applied from the left. When the Landau gauge is used, the de Donder gauge-fixing condition \(\mathrm{d}{{\mathfrak {S}}}_{ab}=0\) is obtained.

figure o
$$\begin{aligned} \left( d{\tilde{{{\mathfrak {c}}}}}^{\circ \circ }+ \alpha {\tilde{{{\mathfrak {c}}}}}^{{\circ }}_{~\star }\wedge {{\mathfrak {b}}}^{\star {\circ }}\right) \wedge {\overline{{{\mathfrak {e}}}}}_{a{\circ }{\circ }}= & {} 0, \end{aligned}$$
(56)

where the anticommutation among \({{\mathfrak {c}}}\), \(\delta {{\mathfrak {c}}}\) and \({{\mathfrak {b}}}\) is used.

figure p
$$\begin{aligned} \epsilon _{ab{\circ \circ }}\left( \mathrm{d}\left( {{\mathfrak {c}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\right) -\alpha {{\mathfrak {b}}}^{\circ }_{~\star }\wedge {{\mathfrak {c}}}^\star \wedge {{\mathfrak {e}}}^{\circ }\right) = \epsilon _{ab{\circ \circ }}\left( \mathrm{d}{{\mathfrak {c}}}^{\circ }-\alpha {{\mathfrak {b}}}^{\circ }_{~\star }\wedge {{\mathfrak {c}}}^\star \right) \wedge {{\mathfrak {e}}}^{\circ }~=~0, \end{aligned}$$

where the de Donder condition is used.

The BRST transformation may give another set of equations, which must be consistent with above equations:

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ {{\mathfrak {T}}}^a\right]= & {} \chi ^a_{~{\circ }}~\mathrm{d}{{\mathfrak {e}}}^{\circ }+\mathrm{d}\chi ^a_{~{\circ }}\wedge {{\mathfrak {e}}}^{\circ }+{c_{gr}}\;\chi ^a_{~{\circ }}~{{\mathfrak {w}}}^{~{\circ }}_{\star }\wedge {{\mathfrak {e}}}^\star \\&+{c_{gr}}\;\chi _{\star }^{~{\circ }}~{{\mathfrak {w}}}^{a}_{~{\circ }}\wedge {{\mathfrak {e}}}^\star -\mathrm{d}\chi ^a_{~{\circ }}\wedge {{\mathfrak {e}}}^{\circ }+{c_{gr}}\;\chi ^{\circ }_{~\star }{{\mathfrak {w}}}^a_{~{\circ }}\wedge {{\mathfrak {e}}}^\star \\= & {} \chi ^a_{~{\circ }}~{{\mathfrak {T}}}^{\circ }~=~0. \end{aligned}$$

This is consistent with the torsion-less condition. The BRST transformation for the volume form is vanished, and last two terms are cancelled each other such as

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ i\left( d{\tilde{{{\mathfrak {c}}}}}^{\circ \circ }+\alpha {\tilde{{{\mathfrak {c}}}}}^{\circ }_{~\star } \wedge {{\mathfrak {b}}}^{\star {\circ }}\right) \wedge {\overline{{{\mathfrak {c}}}}}_{a{\circ \circ }}\right]= & {} -\left( \mathrm{d}{{\mathfrak {b}}}^{\circ \circ }+\alpha {{\mathfrak {b}}}^{\circ }_{~\star }\wedge {{\mathfrak {b}}}^{\star {\circ }}\right) \wedge {\overline{{{\mathfrak {e}}}}}_{a{\circ \circ }} \end{aligned}$$

Therefore, the BRST transformation of (55) is given by

$$\begin{aligned} 0= & {} \epsilon _{abc\bullet }\delta _{\mathrm {BRST}}\left[ \left( \mathrm{d}{{\mathfrak {w}}}^{ab} +{c_{gr}}\;{{\mathfrak {w}}}^a_{~{\circ }}\wedge {{\mathfrak {w}}}^{{\circ }b}\right) \wedge {{\mathfrak {e}}}^c\right] \\= & {} \epsilon _{abc\bullet }\Bigl \{ \chi ^b_{~{\circ }}\left( \mathrm{d}{{\mathfrak {w}}}^{a{\circ }} +{c_{gr}}\;{{\mathfrak {w}}}^a_{~\star }\wedge {{\mathfrak {w}}}^{\star {\circ }}\right) \wedge {{\mathfrak {e}}}^c +\left( \mathrm{d}{{\mathfrak {w}}}^{ab}+{c_{gr}}\;{{\mathfrak {w}}}^a_{~{\circ }}\wedge {{\mathfrak {w}}}^{{\circ }b}\right) \wedge {{\mathfrak {c}}}^c \Bigr \}. \end{aligned}$$

This is consistent with the Einstein equation.

$$\begin{aligned} \delta _{\mathrm {BRST}}\left[ \mathrm{d}{{\mathfrak {e}}}^a-\alpha \left( {{\mathfrak {b}}}^a_{~{\circ }}\wedge {{\mathfrak {e}}}^{\circ }-i{\tilde{{{\mathfrak {c}}}}}_{\circ }^a\wedge {{\mathfrak {c}}}^{\circ }\right) \right]= & {} \mathrm{d}{{\mathfrak {c}}}^a=0, \end{aligned}$$

where \(\alpha \)-terms are cancelled each other. The BRST transformation of (56) gives an equation of motion for \({{\mathfrak {b}}}\) in (55).

$$\begin{aligned} \epsilon _{ab{\circ \circ }}\delta _{\mathrm {BRST}}\left[ \mathrm{d}\left( {{\mathfrak {c}}}^{\circ }\wedge {{\mathfrak {e}}}^{\circ }\right) -\alpha {{\mathfrak {b}}}^{\circ }_{~\star }\wedge {{\mathfrak {c}}}^\star \wedge {{\mathfrak {e}}}^{\circ }\right]= & {} 0, \end{aligned}$$

where \(\epsilon _{ab{\circ \circ }}{{\mathfrak {c}}}^{\circ }\wedge {{\mathfrak {c}}}^{\circ }=0\) is used. This is not an equation, but an identity.

1.4 D proof of (44)

The Proof of (44) can be given as follows: Shorthand notations are introduced to omit indices in this appendix for simplicity as follows:

$$\begin{aligned} {\widehat{{\mathfrak {Q}}}}_{{\mathfrak {b}}}=\frac{1}{2}\widehat{{\varvec{\beta }}}{\circ }\delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] ,&~~&{\widehat{{\mathfrak {Q}}}}_{{{\tilde{{{\mathfrak {c}}}}}}}=\frac{i}{4}\widehat{{\varvec{\chi }}}{\circ }\delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] -\frac{1}{4}\widehat{{\varvec{\beta }}}{\circ }\widehat{{\varvec{{\mathfrak {Q}}}}}, \end{aligned}$$

where \(\widehat{{\varvec{{\mathfrak {Q}}}}}\) is defined in Eq. (41). By using these notions, the commutation relation can be represented as

$$\begin{aligned} -8i\left[ {\widehat{{\mathfrak {Q}}}}_{{{\tilde{{{\mathfrak {c}}}}}}},{\widehat{{\mathfrak {Q}}}}_{{\mathfrak {b}}}\right]= & {} \left[ \widehat{{\varvec{\beta }}}{\circ }\delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] , \widehat{{\varvec{\chi }}}{\circ }\delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] \right] +i \left[ \widehat{{\varvec{\beta }}}{\circ }\delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] , \widehat{{\varvec{\beta }}}{\circ }\widehat{{\varvec{{\mathfrak {Q}}}}} \right] ,\nonumber \\= & {} \widehat{{\varvec{\beta }}}\left[ \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] , \widehat{{\varvec{\chi }}}\right] \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] +\widehat{{\varvec{\chi }}} \left[ \widehat{{\varvec{\beta }}},\delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] \right] \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] \nonumber \\&+i\; \widehat{{\varvec{\beta }}}\left[ \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] ,\widehat{{\varvec{\beta }}}{\circ }\widehat{{\varvec{{\mathfrak {Q}}}}}\right] +i \left[ \widehat{{\varvec{\beta }}}, \widehat{{\varvec{\beta }}}{\circ }\widehat{{\varvec{{\mathfrak {Q}}}}} \right] \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] . \end{aligned}$$
(57)

where \([\widehat{{\varvec{\beta }}},\widehat{{\varvec{\chi }}}]=0\) is used. The first term of (57) becomes

$$\begin{aligned} \widehat{{\varvec{\beta }}}\left[ \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] , \widehat{{\varvec{\chi }}}\right] \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right]= & {} \widehat{{\varvec{\beta }}} \left( \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] \widehat{{\varvec{\chi }}}- \widehat{{\varvec{\chi }}}~\delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] \right) \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] , \\= & {} \widehat{{\varvec{\beta }}} \left( \delta _{\mathrm {BRST}}\left[ \left\{ \widehat{{\varvec{{\mathfrak {Q}}}}},\widehat{{\varvec{\chi }}}\right\} \right] +i\left[ \widehat{{\varvec{\beta }}},\widehat{{\varvec{{\mathfrak {Q}}}}}\right] \right) \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] ,\\= & {} i\widehat{{\varvec{\beta }}}~\left[ \widehat{{\varvec{\beta }}},\widehat{{\varvec{{\mathfrak {Q}}}}}\right] \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] =~4{\widehat{{\mathfrak {Q}}}}_{{\mathfrak {b}}}, \end{aligned}$$

where \(\delta _{\mathrm {BRST}}\left[ \left\{ \widehat{{\varvec{{\mathfrak {Q}}}}},\widehat{{\varvec{\chi }}}\right\} \right] =0\) due to (40), and (38) are used. Note that \(\widehat{{\varvec{\chi }}}\) and \(\widehat{{\varvec{{\mathfrak {Q}}}}}\) have \(\epsilon _X=-1\) in (28). The second term of (57) is zero, because

$$\begin{aligned} \left[ \widehat{{\varvec{\beta }}},\delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] \right]= & {} \delta _{\mathrm {BRST}}\left[ \left[ \widehat{{\varvec{\beta }}},\widehat{{\varvec{{\mathfrak {Q}}}}} \right] \right] =~0, \end{aligned}$$

due to (38). The third term is also zero as

$$\begin{aligned} i\widehat{{\varvec{\beta }}}\left[ \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] ,\widehat{{\varvec{\beta }}}{\circ }\widehat{{\varvec{{\mathfrak {Q}}}}}\right]= & {} i\widehat{{\varvec{\beta }}}\widehat{{\varvec{\beta }}} \left[ \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] ,\widehat{{\varvec{{\mathfrak {Q}}}}}\right] + i\widehat{{\varvec{\beta }}}~ \delta _{\mathrm {BRST}}\left[ \left[ \widehat{{\varvec{{\mathfrak {Q}}}}},\widehat{{\varvec{\beta }}}\right] \right] \widehat{{\varvec{{\mathfrak {Q}}}}}=~0, \end{aligned}$$

due to (38). A relation \(\left[ \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] ,\widehat{{\varvec{{\mathfrak {Q}}}}}\right] =0\) can be confirmed by direct calculations. The last term of (57) becomes

$$\begin{aligned} i\left[ \widehat{{\varvec{\beta }}}, \widehat{{\varvec{\beta }}}{\circ }\widehat{{\varvec{{\mathfrak {Q}}}}} \right] \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right]= & {} i\widehat{{\varvec{\beta }}}~ \left[ \widehat{{\varvec{\beta }}},\widehat{{\varvec{{\mathfrak {Q}}}}}\right] \delta _{\mathrm {BRST}}\left[ \widehat{{\varvec{{\mathfrak {Q}}}}}\right] =~4{\widehat{{\mathfrak {Q}}}}_{{\mathfrak {b}}}. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kurihara, Y. Nakanishi–Kugo–Ojima quantization of general relativity in Heisenberg picture. Eur. Phys. J. Plus 136, 462 (2021). https://doi.org/10.1140/epjp/s13360-021-01463-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-021-01463-3

Navigation