Skip to main content
Log in

The Quasi-Maxwellian Equations of General Relativity: Applications to Perturbation Theory

  • Particles and Fields
  • Published:
Brazilian Journal of Physics Aims and scope Submit manuscript

Abstract

A comprehensive review of the equations of general relativity in the quasi-Maxwellian (QM) formalism introduced by Jordan, Ehlers and Kundt is presented. Our main interest concerns its applications to the analysis of the perturbation of standard cosmology in the Friedman-Lemaître-Robertson-Walker framework. The major achievement of the QM scheme is its use of completely gauge-independent quantities. We shall see that in the QM-scheme, we deal directly with observable quantities. This reveals its advantage over the old method introduced by Lifshitz that deals with perturbation in the standard framework. For completeness, we compare the QM-scheme to the gauge-independent method of Bardeen, a procedure consisting of particular choices of the perturbed variables as a combination of gauge-dependent quantities.

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.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. We introduce the subscript indexes E and QM to distinguish the solution of the Einstein equations valid only on the Cauchy surface and the solution propagated by the quasi-Maxwellian equations, respectively.

  2. If we consider that the origin of these corrections comes from quantum fluctuations, then the value of the constants α and β are fixed—see Heisenberg and Euler [66].

  3. We shall refer to the gauge-invariant (gauge-dependent) variables as “good” (“bad”) ones, a terminology inspired in the Stewart’s Lemma [148].

  4. This procedure was adopted in several instances in the literature. See, for example, Sagnotti [145].

  5. This choice is necessary to avoid that complex terms or explicit dependences on spatial coordinates that occur when the simple derivatives of the vector basis are calculated.

  6. Equation (154) calls for a choice of a specific direction for the basis, which has been made a number of times in the literature. See, for example, [145].

  7. This choice was made to avoid spatially-dependent terms when calculating the derivative \((\hat {U}^{\mu }{}{~}_{\nu })\dot {}\)

  8. Since it has no dynamical equation of its own, the acceleration, i.e., the variable α, must be eliminated to make the dynamical system closed. This will be achieved by fixing a value for the function (δ a α ).

  9. The case λ < 0 yields (n 3)2 < 0 for all x and will therefore not be considered here.

  10. The same observation regarding the variable a α that was stated for the tensorial case holds here as well.

  11. In nonequilibrium, thermodynamics, both τ and ξ, are functions of the system equilibrium variables, such as the density ρ and the temperature T.

  12. Let us point out that some of the gauge-dependent terms are particularly relevant, δ ρ here included.

  13. However, as we shall see soon, we can construct associated “good” quantities in terms of these scalars.

  14. One can write these invariants in a pure geometrical way without using Einstein’s equations. This does not modify our argument.

  15. This is a consequence of the vanishing of the perturbation of the magnetic part of Weyl tensor(cf.above).

  16. The vorticity is of course zero, since we are limiting ourselves to the irrotational case.

  17. We will set Y = Z = μ = 0, since these homogeneous terms are just a matter of choosing the coordinate system. Nevertheless, we have are not interested in examining such pure gauge quantities as Y, Z and μ.

  18. In the general case, ξ and τ are functions of the equilibrium variables (for instance, ρ and the temperature T) and, since both variations δπ i j and δ σ i j are expanded in terms of the traceless tensor \(\hat {Q}_{ij}\), it follows that the above relation does not restrain the kind of fluid we are examining. However, if we consider ξ as time-dependent, the quantity δπ i j must be included in the fundamental set \(\mathcal {M}_{[A]}\).

  19. Recall that our discussion is restricted to irrotational perturbations. Our results are therefore simpler. Our method, however, is free from this restriction and generic cases can equally well be obtained along the same lines.

  20. The quantity Q in this subsection should not be confused with the previous scalar basis.

  21. The above-mentioned gauge problem has been widely discussed in the literature (see [124] and references therein).

  22. This tensor was introduced in the 1930s to provide, much as the symmetric tensor φ μ ν does—in a more often used approach—an alternative description of spin-2 field in the Minkowski background. In the 1960s, Lanczos rediscovered it, without recognizing he was dealing with the same object, as a Lagrange multiplier used to obtain the Bianchi identities in the context of Einstein’s General Relativity. However, only recently (cf. Novello [114, 115]) a complete analysis of Fierz-Lanczos object was undertaken and it was discovered that its generic (Fierz) version describes not only one, but two spin-2 fields. The restriction to a single spin-2 field is usually called the Lanczos tensor. We will limit all our considerations to this restricted quantity.

  23. Terminology due to Bergmann relative to Dirac’s work [41] on constrained systems.

  24. Linearity required to preserve the coherence with our basic assumption of linear-perturbations approximation. For the understanding of the physical meaning of this relation, see the examples in Section 3.5.1.

  25. Astronomical observations show that the Hubble constant, here translated to θ, is positive, even if there is no universal agreement on its magnitude. Thermodynamical reasoning ensures the nonnegativeness of the parameter ξ.

  26. Here, we are considering only the cases in which the minimal closed set of observables contains only two variables. In the more general vectorial case, ℳ should be a 4 × 4 matrix, as stated above (see Section 3.7.3 for more details).

  27. In the specific case of a Stokesian fluid, the observables for vectorial perturbations yield a reduced dynamical system (for more information on that issue, see Novello [123]).

  28. To be compatible with most of the cosmological models, in this section, we assume that the quantum phase of the Universe is radiation dominated. Note that the calculations can be easily extended for the case in which a previous inflationary regime is present.

  29. This quantity is defined as the mean-square uncertainty in the annihilation operator \(\hat {a}\). The total noise of a Gaussian pure state is conserved even if the total number of photons is not and, it is therefore more useful to describe the quantum wave functions obtained from the Schrödinger equation. See Novello et al. [124] for more details on this.

  30. Since we are dealing with a linear process, each mode can be analyzed separately.

  31. This resemblance is far from being accepted by all physics community. Indeed, in the final session of the 1972 Copenhagen International Conference on Gravitation and Relativity, A. Trautman argued that perhaps many of the difficulties of gravitational theory may be due to the extension of this similarity to all aspects of both fields.

  32. This hypothesis is made here only for simplicity. It is an oversimplification under certain drastic situations, such as very strong gravitational fields.

  33. We are, perhaps, in a situation similar to that experienced by Maxwell, a century ago. His theory described the electromagnetic fields in the interior of substances by means of the same type of fields in vacuum and by characterizing the distortion produced by the matter on the fields, as given by macroscopic quantities: the dielectric constant ε Max and the permeability μ Max (the shorthand “Max” represents “Maxwell”). It took many years before Lorentz—who had the atomic theory of matter at his disposal—made Maxwell’s theory rigorously understood by averaging properties of microscopic fields on a macroscopic scale.

  34. In the present review, we shall limit ourselves to the sourceless case, i.e., the so-called vacuum gravitational fields. A generalization to include matter is straightforward and presents no difficulties.

  35. Indeed, as we shall see, in the case E α β = D α β and B α β = H α β , Q α β ν μ can be identified with Weyl’s tensor and (10) reduces to Einstein’s equation in the vacuum.

  36. This equivalence is only complete if we impose as initial date the set R μ ν = 0 on a given space-like hypersurface.

  37. We are certainly not considering a usual Minkowski space-time. Here, we are considering a more complex structure that takes into account the fluctuations of space-time (as in quantum gravidynamics). These fluctuations are assumed to be represented on an average by the macroscopic quantities ε and μ, as stated in the introduction.

References

  1. M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions (Dover Publications, New York, 1974), p. 504

    Google Scholar 

  2. J. Acacio de Barros, N. Pinto-Neto, M.A. Sagioro-Leal, Phys. Lett. A. 241, 229 (1998)

    ADS  Google Scholar 

  3. R. Adler, M. Bazin, M. Schiffer, Introduction to General Relativity, Chap. 6 (McGraw-Hill, New York, 1975)

    Google Scholar 

  4. G. Ares de Parga, O. Chavoya, J.L. López Bonilla, J. Math. Phys. 30, 1294 (1989)

    MathSciNet  MATH  ADS  Google Scholar 

  5. E. Ayon-Beato, A. Garcia, Phys. Rev. Lett. 80, 5056 (1998)

    ADS  Google Scholar 

  6. E. Ayon-Beato, A. Garcia, Gen. Relat. Gravit. 31, 629 (1999a)

    MathSciNet  MATH  ADS  Google Scholar 

  7. E. Ayon-Beato, A. Garcia, Phys. Lett. B. 464, 25 (1999b)

    MathSciNet  MATH  ADS  Google Scholar 

  8. J Bardeen. Proceedings of GR5, (Tbilisi, U.S.S.R, 1968), p. 174

  9. J. Bardeen, Phys. Rev. D. 22, 1882 (1980)

    MathSciNet  ADS  Google Scholar 

  10. C. Barrabes, V.P. Frolov, Phys. Rev. D. 53, 3215 (1996)

    MathSciNet  ADS  Google Scholar 

  11. H. Bateman, The Mathematical Analysis of Electrical and Optical Wave Motion (Dover, New York, 1955)

    MATH  Google Scholar 

  12. L. Bel, Colloques Internationaux du Centre National de la Recherche Scientifique (Paris, 1962), p. 119

  13. V.A. Belinski, I.M. Khalatnikov, E.M. Lifshtz, Usp. Fiz. Nauk. 102, 463 (1970)

    ADS  Google Scholar 

  14. V.A. Belinski, I.M. Khalatnikov, E.M. Lifshtz, Adv. Phys. 19, 525 (1880)

    ADS  Google Scholar 

  15. I. Ben-Dayan, M. Gasperini, G. Marozzi, F. Nugier, G. Veneziano, arXiv:astro-ph.CO/1202.1247v2 (2012)

  16. Z. Bialynicka-Birula, I. Bialynicka-Birula, J. Opt. Soc. Am. B. 4, 10 (1987)

    Google Scholar 

  17. J. Bicak, J. Griffiths, Ann. Phys. (N.Y.) 252, 180 (1996)

    MathSciNet  MATH  ADS  Google Scholar 

  18. N.D. Birrel, P.C.W. Davis. Quantum Fields in Curved Space (Cambridge University Press, Cambridge, 1982)

    Google Scholar 

  19. E. Bittencourt, Master’s Thesis, CBPF/Rio de Janeiro (2009)

  20. H.J. Blome, C. Chicone, F.W. Hehl, B. Mashhoon, arXiv:gr-qc/1002.1425 (2010)

  21. P.R. Brady, S. Droz, W. Israel, S.M. Morsink, Class. Quantum Grav. 13, 2211 (1996)

    MathSciNet  MATH  ADS  Google Scholar 

  22. R. Brandenberger, in Proceedings of the VIII Brazilian School of Cosmology and Gravitation, ed. by M. Novello (Editions Frontiéres, Singapore, 1996)

  23. R. Brandenberger, R. Khan, W.H. Press, Phys. Rev. D. 28, 1809 (1983)

    MathSciNet  ADS  Google Scholar 

  24. R. Brandenberger, V. Mukhanov, A. Sornborger, Phys. Rev. D. 48, 1629 (1993)

    MathSciNet  ADS  Google Scholar 

  25. M. Bruni, P.K.S. Dunsby, G.F.R. Ellis, Astrophys. J. 395, 34 (1992)

    ADS  Google Scholar 

  26. M. Bruni, S. Matarrese, O. Pantano, Phys. Rev. Lett. 74, 1916 (1995)

    ADS  Google Scholar 

  27. A. Cabo, E. Ayon-Beato, arXiv:gr-qc/9704073 (1997)

  28. R.R. Caldwell, R. Dare, P.J. Steinhardt, Phys. Rev. Lett. 80, 1582 (1998)

    ADS  Google Scholar 

  29. M. Campista, Ph. D. Thesis, CBPF/Rio de Janeiro (2010)

  30. M. Campista, M. Novello, J.M. Salim, Int. J. Mod. Phys. A. 25, 3095 (2010)

    MATH  ADS  Google Scholar 

  31. A. Campos, B.L. Hu, Phys. Rev. D. 58, 125021 (1998)

    ADS  Google Scholar 

  32. C. Cartier, R. Durrer, E.J. Copeland, Phys. Rev. D. 67, 103517 (2003)

    ADS  Google Scholar 

  33. C. Chicone, B Mashhoon, arXiv:gr-qc/1111.4702 (2008)

  34. Y. Choquet-Bruhat, R. Geroch, Comm. Math. Phys. 14, 329 (1969)

    MathSciNet  MATH  ADS  Google Scholar 

  35. Y. Choquet-Bruhat, General Relativity and ’s Equations (Oxford University Press, New York, 2009)

    Google Scholar 

  36. R. Coqueraux, A. Grossmann, Ann. Phys. (N.Y.) 143, 296 (1982)

    ADS  Google Scholar 

  37. M. Cvetic, Phys. Rev. Lett. 71, 815 (1993)

    MathSciNet  MATH  ADS  Google Scholar 

  38. M. Dabrowski, J. Stelmach, Astron. J. 92, 1272 (1986)

    ADS  Google Scholar 

  39. V.A. De Lorenci, R. Klippert, M. Novello, J.M. Salim, Phys. Rev. D. 65, 063501 (2002). arXiv:gr-qc/9806076v2 (1998)

    ADS  Google Scholar 

  40. R. Debever, Cahiers de Physique 168 (1964)

  41. P.A.M. Dirac, Can. J. Math. 2, 129 (1950)

    MathSciNet  MATH  Google Scholar 

  42. G. Dunne, T. Hall, Phys. Rev. D. 58, 105022 (1998)

    ADS  Google Scholar 

  43. G. Dunne, Int. J. Mod. Phys. A. 12, 1143 (1997)

    MathSciNet  ADS  Google Scholar 

  44. P.K.S. Dunsby, B.A.C.C. Bassett, G.F.R. Ellis, Class. Quantum Grav. 14, 1215 (1997)

    MathSciNet  MATH  ADS  Google Scholar 

  45. D. Edwards, Astrophys. Space Sci. 24, 563 (1973)

    ADS  Google Scholar 

  46. A. Einstein, Ann. Phys. 354, 769 (1916)

    Google Scholar 

  47. A. Einstein, Rev. Mod. Phys. 20, 35 (1048)

    ADS  Google Scholar 

  48. G.F.R. Ellis. Proceedings of the International School of Physics, Enrico Fermi Course XLVII (London Academic 104, 1971)

  49. G.F.R. Ellis, in Cargèse Lectures in Physics Vol. VI, ed. by E. Schatzmann (Gordon and Breach, London, 1971)

  50. G.F.R. Ellis, M. Bruni. Phys. Rev. D. 40, 1804 (1989)

    MathSciNet  ADS  Google Scholar 

  51. G.F.R. Ellis, H. van Elst, in NATO science series. Ser. C, Math. Phys. Sci., vol. 541, ed. by M. Lachièze-Rey (Kluwer Academic, 1999), p. 1

  52. J.C. Fabris, J.M. Salim, S.L. Sautu, Mod. Phys. Lett. A. 13, 953 (1998)

    MathSciNet  ADS  Google Scholar 

  53. M. Fierz, Helv. Phys. Acta. 12, 379 (1939)

    Google Scholar 

  54. R. Garcia-Salcedo, N. Breton, Int. J. Mod. Phys. A. 15, 4341 (2000)

    MATH  ADS  Google Scholar 

  55. M. Gasperini, G. Marozzi, F. Nugier, G. Veneziano, JCAP. 07, 008 (2011)

    ADS  Google Scholar 

  56. M. Gasperini, G. Veneziano, Phys. Rep. 373, 1 (2003)

    MathSciNet  ADS  Google Scholar 

  57. M Gasperini, G Veneziano. arXiv: abs/hep-th/0207130 (2002)

  58. M. Giovannini, M. Shaposhnikov, Phys. Rev. D. 57, 2186 (1998)

    ADS  Google Scholar 

  59. S.W. Goode, Phys. Rev. D. 39, 2882 (1989)

    MathSciNet  ADS  Google Scholar 

  60. E. Goulart, Ph.D. Thesis, CBPF/Rio de Janeiro (2007)

  61. E. Goulart, M. Novello, Grav. Cosmol. 14, 321 (2008)

    MathSciNet  MATH  ADS  Google Scholar 

  62. I.S. Gradshteyn, I.M. Ryzhik. Table of Integrals, Series and Products (Academic, London, 1965)

    Google Scholar 

  63. L.P. Grishchuk, Y.V. Sidorov, Phys. Rev. D. 42, 3413 (1990)

    MathSciNet  ADS  Google Scholar 

  64. S. Hawking, G.F.R. Ellis, The Large Scale Structure of Space-Time (Cambridge University Press, Cambridge, 1977)

    Google Scholar 

  65. S. Hawking, Ap. J. 145, 544 (1966)

    ADS  Google Scholar 

  66. W. Heisenberg, H. Euler, Z. Phys. 98, 714 (1936)

    ADS  Google Scholar 

  67. M. Hindmarsh, A. Everett, Phys. Rev. D. 58, 103505 (1998)

    ADS  Google Scholar 

  68. P.A. Hogan, Astrophys. J. 333, 64 (1988)

    MathSciNet  ADS  Google Scholar 

  69. P.A. Hogan, G.F.R. Ellis, Class. Quantum Grav. 14, A171 (1997)

    MathSciNet  MATH  ADS  Google Scholar 

  70. P.A. Hogan, E.M. O’Shea, Phys. Rev. D. 65, 124017 (2002)

    MathSciNet  ADS  Google Scholar 

  71. P.A. Hogan, A. Trautman, in Gravitation and Geometry: Papers in Honour of I. Robinson, ed. by W. Rindler, A. Trautman (Bibliopolis, Naples, 1987)

  72. J.H. Horne, G.T. Horowitz, Nucl. Phys. B368, 444 (1992)

    MathSciNet  ADS  Google Scholar 

  73. J.C. Hwang, E.T. Vishniac, Ap. J. 353, 1 (1990)

    MathSciNet  ADS  Google Scholar 

  74. J. Ibáñez, R.J. van den Hoogen, A.A. Coley, Phys. Rev. D. 51, 928 (1995)

    MathSciNet  ADS  Google Scholar 

  75. W. Israel, Ann. Phys. 100, 310 (1976)

    MathSciNet  ADS  Google Scholar 

  76. K. Jedamzik, V. Jatalinić, A.V. Olinto, Phys. Rev. D. 57, 3264 (1998)

    ADS  Google Scholar 

  77. B.J.T. Jones, Rev. Mod. Phys. 48, 107 (1976)

    ADS  Google Scholar 

  78. P. Jordan, J. Ehlers, R. Sachs, Akad. Wiss. Lt. Meinz Abh. Math. Nat. Kl. 1, 3 (1961)

    Google Scholar 

  79. P. Jordan, J. Ehlers, W. Kundt, Akad. Wiss. Lit. (Mainz) Abhandl. Math.-Nat. Kl. 2. 2, 21 (1960)

    MathSciNet  Google Scholar 

  80. P. Jordan, J. Ehlers, W. Kundt, Gen. Rel. Grav. 41, 2191 (2009)

    MathSciNet  MATH  ADS  Google Scholar 

  81. M. Joyce, M. Shaposhnikov, Phys. Rev. Lett. 79, 1193 (1997)

    ADS  Google Scholar 

  82. E.G. Kalnins, W. Miller Jr., J. Math. Phys. 32, 698 (1991)

    MathSciNet  MATH  ADS  Google Scholar 

  83. R. Klippert, Ph.D. Thesis, CBPF/Rio de janeiro (1998)

  84. R. Klippert, V.A. De Lorenci, M. Novello, J.M. Salim, Phys. Lett. B. 472, 27 (2000)

    ADS  Google Scholar 

  85. L. Kofman, A. Linde, A.A. Starobinsky, Phys. Rev. D. 56, 3258 (1997)

    ADS  Google Scholar 

  86. C. Lanczos, Rev. Mod. Phys. 34, 379 (1962)

    MathSciNet  MATH  ADS  Google Scholar 

  87. L.D. Landau, E.M. Lifshitz. Fluid Mechanics (Pergamon Press, New York, 1982)

    Google Scholar 

  88. A. Lichnerowicz, Ann. Mat. Pura Appl. 50, 1 (1960)

    MathSciNet  MATH  Google Scholar 

  89. E.M. Lifshitz, I.M. Khalatnikov, Adv. Phys. 12, 185 (1963)

    MathSciNet  ADS  Google Scholar 

  90. M. Marklund, P.K.S. Dunsby, G Brodin, Phys. Rev. D. 62, 101501 (2000)

    ADS  Google Scholar 

  91. M. Mars, M.M. Martin-Prats, J.M.M. Senovila, Clas. Quantum Grav. 13, L51 (1996)

    MATH  ADS  Google Scholar 

  92. R. Maartens, B.A. Bassett, Class. Quant. Grav. 15, 705 (1998)

    MathSciNet  MATH  ADS  Google Scholar 

  93. B. Mashhoon, Ann. Phys. 17, 705 (2008). arXiv:gr-qc/0805.2926 (2008)

  94. B. Mashhoon, arXiv:gr-qc/1101.3752 (2011)

  95. A.L. Matacz, Phys. Rev. D. 49, 788 (1994)

    MathSciNet  ADS  Google Scholar 

  96. S. Matarrese, O. Pantano, D. Saez, Phys. Rev. D. 47, 1311 (1993)

    ADS  Google Scholar 

  97. S. Matarrese, O. Pantano, D. Saez, Phys. Rev. Lett. 72, 320 (1994)

    MathSciNet  MATH  ADS  Google Scholar 

  98. P.G. Miedema, Phys. Rev. D. 50, 2431 (1994)

    ADS  Google Scholar 

  99. R. Moessner, M. Trodden. 51, 2801 (1995). ibid

  100. M.C. Motta da Silva, Ph.D. Thesis, CBPF/rio de Janeiro (1998)

  101. V. Mukhanov, R. Brandenberger, Phys. Rev. Lett. 68, 1969 (1992)

    MathSciNet  MATH  ADS  Google Scholar 

  102. V.F. Mukhanov, H.A. Feldmann, R.H. Brandenberger, Phys. Rep., Vol. 215 (1992)

  103. V.F. Mukhanov, S. Winitzki, Introduction to Quantum Effects in Gravity (Cambridge University Press, Cambridge, 2007)

    MATH  Google Scholar 

  104. G.L. Murphy, Phys. Rev. D. 8, 4231 (1973)

    ADS  Google Scholar 

  105. H. Mutoh, T. Hirai, K. Maeda, Phys. Rev. D. 55, 3276 (1997)

    ADS  Google Scholar 

  106. E.T. Newman, R. Penrose, J. Math. Phys. 3, 566 (1962)

    MathSciNet  ADS  Google Scholar 

  107. H. Noh, J. Hwang, Phys. Rev. D. 52, 1970 (1995)

    ADS  Google Scholar 

  108. M. Novello, Inter. Atomic Energy Agency Ic. 75 (61), 1 (1975)

    Google Scholar 

  109. M. Novello, in Proceedings of the II Brazilian School of Cosmology and Gravitation, ed. by M. Novello, J. Sansom, Cia, R. de Janeiro (Portuguese, 1980)

  110. M. Novello, E. Bittencourt, Gravit. Cosmol. 17(3), 230 (2011)

    MathSciNet  MATH  ADS  Google Scholar 

  111. M. Novello, S.E.P. Bergliaffa, Phys. Rep. 463, 127 (2008). arXiv:astro-ph/0802.1634v1 (2008)

    MathSciNet  ADS  Google Scholar 

  112. M. Novello, S.L.S. Duque, Phys. A. 168, 1073 (1990)

    Google Scholar 

  113. M. Novello, L.R. de Freitas, Int. J. Mod. Phys. A. 13, 363 (1998)

    MATH  ADS  Google Scholar 

  114. M. Novello, N. Pinto-Neto, Fortschr. Phys. 40, 173 (1992a)

    MathSciNet  Google Scholar 

  115. M. Novello, N. Pinto-Neto, Fortschr. Phys. 40, 195 (1992b)

    MathSciNet  Google Scholar 

  116. M. Novello, L. Oliveira, J.M. Salim, E. Elbaz, Int. J. Mod. Phys. D. 1, 641 (1993)

    MathSciNet  ADS  Google Scholar 

  117. M. Novello, J.M. Salim, Phys. Rev. D. 20, 377 (1979)

    MathSciNet  ADS  Google Scholar 

  118. M. Novello, J.M. Salim, Fund. Cosmic Phys. 8, 201 (1983)

    ADS  Google Scholar 

  119. M. Novello, J.M. Salim, CBPF-NF-021/85, Rio de Janeiro (1985)

  120. M. Novello, J.M. Salim, in The Tenth Marcel Grossmann Meeting, ed. by M. Novello, S.E.P. Bergliaffa, R. Ruffini (World Scientific, Cingapura, 2006)

  121. M. Novello, J.M. Salim, H. Heintzmann, in J. Leite Lopes Festschrift, ed. by S. Joffily, (1988)

  122. M. Novello, J.M. Salim, M.C. Motta da Silva, S.E. Jorás, R. Klippert, Phys. Rev. D. 51, 450 (1995a)

    MathSciNet  ADS  Google Scholar 

  123. M. Novello, J.M. Salim, M.C. Motta da Silva, S.E. Jorás, R. Klippert, Phys. Rev. D. 52, 730 (1995b)

    MathSciNet  ADS  Google Scholar 

  124. M. Novello, J.M. Salim, M.C. Motta da Silva, R. Klippert, Phys. Rev. D. 54, 2578 (1996)

    MathSciNet  ADS  Google Scholar 

  125. M. Novello, J.M. Salim, M.C. Motta da Silva, R. Klippert, Phys. Rev. D. 61, 124025 (2000)

    MathSciNet  ADS  Google Scholar 

  126. M. Novello, A.L. Velloso, Gen. Rel. Grav. 19, 1251 (1987)

    MathSciNet  MATH  ADS  Google Scholar 

  127. H.P. De Oliveira, J.M. Salim, S.L. Sautu, Class. Quantum Grav. 14, 2833 (1997)

    MATH  ADS  Google Scholar 

  128. D.W. Olson, Phys. Rev. D. 14, 327 (1976)

    ADS  Google Scholar 

  129. A. Papapetrou, Lectures on General Relativity (D.Reidel Publishing Company, Dordrecht, 1974)

    MATH  Google Scholar 

  130. P.J.E. Peebles, The Large-Scale Structure of the Universe (Princeton University Press, Princeton, 1980)

    Google Scholar 

  131. P. Peter, N. Pinto-Neto, Phys. Rev. D. 65, 023513 (2001)

    ADS  Google Scholar 

  132. F.A.E. Pirani, Phys. Rev. 105, 1089 (1957)

    MathSciNet  MATH  ADS  Google Scholar 

  133. E.M. O’Shea, Phys. Rev. D. 69, 064038 (2004)

    MathSciNet  ADS  Google Scholar 

  134. E.M. O’Shea. arXiv:gr-qc/0401106 (2004)

  135. T. Regge, J.A. Wheeler, Phys. Rev. 108, 1063 (1957)

    MathSciNet  MATH  ADS  Google Scholar 

  136. H.P. Robertson, Rev. Mod. Phys. 5, 62 (1933)

    ADS  Google Scholar 

  137. I. Robinson, J. Math. Phys. 2, 290 (1961)

    ADS  Google Scholar 

  138. I. Robinson, A. Trautman, J. Math. Phys. 24, 1425 (1983)

    MathSciNet  MATH  ADS  Google Scholar 

  139. M.P. Ryanm Jr., Lecture on VI Brazilian School of Cosmology and Gravitation (1991)

  140. B.L. Schumaker, Phys. Rep. 135, 317 (1986)

    MathSciNet  ADS  Google Scholar 

  141. J.M. Salim, Ph.D. thesis, CBPF/Rio de Janeiro (1982)

  142. J.M. Salim, H.P. de Olivera, Acta Phys. Pol. B. 19, 649 (1988)

    ADS  Google Scholar 

  143. J.M. Salim, S.L. Sautu, Class. Quantum Grav. 13, 353 (1996)

    MathSciNet  MATH  ADS  Google Scholar 

  144. A. Saa, E. Gunzig, L. Brenig, V. Faraoni, T.M. Rocha Filho, A. Figueiredo, arXiv:gr-qc/0012105 (2000)

  145. A. Sagnotti, B. Zwiebach, Phys. Rev. D. 24, 305 (1981)

    MathSciNet  ADS  Google Scholar 

  146. W. de Sitter, Proc. K. Ned. Akad. Wet. 19, 1217 (1917)

    Google Scholar 

  147. J.M. Stewart, Class. Quantum Grav. 7, 1169 (1990)

    MATH  ADS  Google Scholar 

  148. J.M. Stewart, M. Walker, Proc. R. Soc. Lond. A. 341, 49 (1974)

    MathSciNet  ADS  Google Scholar 

  149. K. Subramanian, J.D. Barrow. ibid. 58, 883502 (1998)

    Google Scholar 

  150. T. Tajima, S. Cable, K. Shibata, R.M. Kulsrud, Astrophys. J. 390, 309 (1992)

    ADS  Google Scholar 

  151. C. Thompson, O. Blaes, Phys. Rev. D. 57, 3219 (1998)

    MathSciNet  ADS  Google Scholar 

  152. R.C. Tolman, P. Ehrenfest, Phys. Rev. 36, 1791 (1930)

    ADS  Google Scholar 

  153. A. Trautman, Recent Developments in General Relativity (PWN, Warsaw, 1962)

    Google Scholar 

  154. C.G. Tsagas, Class. Quant. Grav. 22, 393 (2005)

    MathSciNet  MATH  ADS  Google Scholar 

  155. C.G. Tsagas, A. Challinor, R. Maartens, Phys. Rep. 465, 61 (2008)

    MathSciNet  ADS  Google Scholar 

  156. A.A. Tseytlin, Phys. Lett. B. 363, 223 (1995)

    MathSciNet  ADS  Google Scholar 

  157. G. Veneziano. arXiv:hep-th/0002094 (2000)

  158. C.V. Vishveshwara, Phys. Rev. D. 1, 2870 (1970)

    ADS  Google Scholar 

  159. J. Wainwright, B.J. Marshman, Phys. Lett. 72A, 275 (1979)

    MathSciNet  ADS  Google Scholar 

  160. R.M. Wald, General relativity (Univ. Chicago Press, Chicago, 1984)

    MATH  Google Scholar 

  161. H. Weyl, Space, time and matter (Dover Publications INC., USA, 1922)

    Google Scholar 

  162. D.L. Wiltshire, Class. Quantum Grav. 28, 164006 (2011)

    MathSciNet  ADS  Google Scholar 

  163. D.L. Wiltshire, P.R. Smale, T. Mattsson, R. Watkins. arXiv:astro-ph.CO/1201.5371v1(2012)

  164. V.D. Zakharov, Gravitational Waves in Einstein’s Theory (Halsted Press, Jerusalem, 1973)

    MATH  Google Scholar 

Download references

Acknowledgments

The authors would like to thank all the participants of the Pequeno Seminário at CBPF for valuable contribution to certain parts of this review. We also thank CNPq, FAPERJ, CAPES (BEX 13956/13-2) FINEP for their financial support. MN acknowledge the ICRANet-Pescara staff for the hospitality during some stages of this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to E. Bittencourt.

Additional information

M. Novello is Cesare Lattes ICRANet Professor

Appendix A: An Example of Further Developments in the QM-Formalism

Appendix A: An Example of Further Developments in the QM-Formalism

This appendix presents an example in which the QM-formalism is modified and the analogy with the electromagnetism is strengthened. Modifications of this kind indicate alternative, more intuitive ways to solve problems appearing in general-relativity theory. We present an enlightening example next.

Electrodynamics is the paradigm of field theory. Its theoretical and experimental properties have been simulated and sought for in many other theories and in particular in the analysis of gravitational phenomena. Much work has been done along this line, which discusses the resemblance between electrodynamics and gravidynamics.Footnote 31 However, it seems possible to further improve this similarity, as we shall show.

In this vein, we review here a modification of Einstein’s theory of general relativity under certain special states of the geometry of the space-time. Since the original proposal of Einstein’s geometrization of gravitational processes, many physicists have discussed alternative models of gravitation. The kind of theory we shall analyze here is given by means of the metric properties—represented by a symmetrical metric tensor g μ ν (x)—and by two other functions, ε(x) and μ(x), which are independent of the metric,Footnote 32 but have intimate connection to the space-time.

For pedagogical reasons, we find it convenient to limit our considerations to the case in which both ε and μ are constants. The meaning one should attribute to these two constants comes from direct analogy with the dielectric and permeability constants of a given medium in electrodynamics.

We shall simplify the model by merely stating that ε and μ can be provisionally identified with the characteristics of certain states of tensions, in free space-time, due to an average procedure on (quantum) properties of gravitation.

In other words, ε and μ are interpreted as the result—in a macroscopic level—of some sort of averaging microscopic field fluctuation.Footnote 33 This is perhaps not difficult to assume if we can say exactly how the equations of motion of gravity phenomena must be modified by them, as we shall do later. We remark that we are not supporting this interpretation but merely suggesting it as a provisional sursis of the model.

We shall describe gravitational interactionFootnote 34 by means of a fourth-rank tensor Q α β μ ν . We shall set up its algebraic properties and give its dynamics. It is possible to separate this tensor, for an observer moving with four-velocity V μ, into four second-order symmetric trace-free tensors E α β , B α β , D α β and H α β . The principal result is then obtained by showing that it is possible to select a class of observers with velocity μ in such a way as to have equations of motion for Q α β μ ν similar to the Maxwell equations for the electrodynamics. That is, for E α β , D α β , B α β and H α β separated into two groups: one containing only E α β and B α β (and their derivatives) and the other containing only H α β and D α β (and their derivatives). These equations have the same formal structure of Maxwell’s equations in a given general medium. We therefore come to the conclusion that the present theory has a class of privileged observers in which gravitational field equations admit this simple separated form. Any other observer, which is in motion with respect to μ, mixes the terms E α β , D α β , H α β and B α β into the equations. This situation could be thought of as defining a new type of ether. However, unlike the ether of the pre-Einstein epoch, our ether is not a substance, but it is only a preferred frame of observation.

In the remainder of this presentation, we discuss in some detail a very particular situation of these tensors, that is, the case in which they can be reduced to two tensors plus two constants: ε and μ. Then, we show that Einstein’s theory is obtained from this for a particular set of values of ε and μ, that is, ε = μ = 1. It is in this sense that we can call this theory a generalization of Einstein’s gravidynamics.

1.1 A1 The Q-Field

1.1.1 A1.1 Definitions

Let us define in a four-dimensional Riemannian manifold a fourth-rank tensor Q α β μ ν described by an observer V μ in terms of four second-order tensors E α β , D α β , H α β and B α β .We set, by analogy with the irreducible decomposition of the Weyl tensor, that

$$\begin{array}{@{}rcl@{}} Q_{\alpha\beta}{~}^{\mu\nu}=V_{[\alpha}D_{\beta ]}{~}^{[\mu} V^{\nu]}+V_{[\alpha}E_{\beta ]}{~}^{[\mu} V^{\nu]}+\delta^{[\mu}_{[\alpha} E^{\nu ]}_{\beta ]} \\-\eta_{\alpha\beta\rho\sigma}V^{\rho}B^{\sigma[\mu}V^{\nu]}-\eta^{\mu\nu\rho\sigma}V_{\rho}H_{\sigma[\alpha}V_{\beta]}.{} \end{array} $$
(A1)

The tensors E α β , D α β , B α β and H α β , represented below by X α β , satisfy the following properties:

$$\begin{array}{@{}rcl@{}} &&{}X^{\alpha}{}{~}_{\alpha}=0, \end{array} $$
(A2a)
$$\begin{array}{@{}rcl@{}} &&{}X^{\alpha \beta}V_{\alpha}=0, \end{array} $$
(A2b)
$$\begin{array}{@{}rcl@{}} &&{}X_{\alpha\beta}=X_{\beta \alpha}. \end{array} $$
(A2c)

We can write D α β , E α β , etc. in terms of Q α β μ ν and projections on V μ like, for instance

$$ D_{\alpha \beta}=-Q_{\varepsilon \alpha \mu \beta}V^{\varepsilon}V^{\mu} $$

and so on.

1.1.2 A1.2 Algebraic Properties

From the definition (A1) of Q α β μ ν , we directly obtain the following properties:

$$\begin{array}{@{}rcl@{}} Q_{\alpha \beta}{~}^{\mu\nu}&=&-Q_{\beta\alpha}{~}^{\mu\nu}, \end{array} $$
(A3)
$$\begin{array}{@{}rcl@{}} Q_{\alpha \beta}{~}^{\mu \nu}&=&-Q_{\alpha \beta}{~}^{\nu \mu}, \end{array} $$
(A4)
$$\begin{array}{@{}rcl@{}} Q^{\alpha}_{\ \ \beta \alpha \nu}&=&E_{\beta \nu}-D_{\beta \nu}, \end{array} $$
(A5)
$$\begin{array}{@{}rcl@{}} &&{\kern-3.6pc}Q^{\alpha}{}{~}_{\alpha}=0 \end{array} $$
(A6)

1.1.3 A1.3 Dynamics

By analogy with Einstein’s equations in vacuum, we impose on Q α β μ ν the equation of motionFootnote 35

$$ Q^{\alpha \beta \mu \nu}{}{~}_{;\nu}=0 $$
(A7)

Now, we shall use the above properties to project the system of (A7) parallel and orthogonal to the rest frame of a selected observer with four-velocity μ from the whole class of V μ. We impose that the congruence generated by μ satisfy the properties.

$$\begin{array}{@{}rcl@{}} &&{}\ell^{\mu}\ell_{\mu}=+1, \end{array} $$
(A8a)
$$\begin{array}{@{}rcl@{}} &&{}w_{\alpha \beta}=\frac{1}{2}\ h_{[\alpha}{\ }^{\lambda}\ h_{\beta ]}{\ }^{\varepsilon}\ell_{\lambda;\varepsilon }=0, \end{array} $$
(A8b)
$$\begin{array}{@{}rcl@{}} &&{}\theta_{\alpha \beta}=\frac{1}{2}\ h_{(\alpha}{\ }^{\lambda}\ h_{\beta )}{\ }^{\varepsilon}\ell_{\lambda;\varepsilon }=0, \end{array} $$
(A8c)
$$\begin{array}{@{}rcl@{}} &&{}\dot{\ell}^{\mu}=\ell^{\mu}{}{~}_{;\nu}\ell^{\nu}=0. \end{array} $$
(A8d)

where h μ ν is the projector in the plane orthogonal to μ, that is

$$ h_{\mu \nu}=g_{\mu \nu}-\ell_{\mu}\ell_{\nu} $$
(A9)

So, the congruence generated by μ is geodesic, irrotational, nonexpanding and shear-free. The reason for selecting such a particular class of observers will become clear later. Then, (A7) takes the form

$$\begin{array}{@{}rcl@{}} &&{}D_{\alpha \mu;\nu}h^{\mu \nu}h^{\alpha}_{\ \ \varepsilon}=0, \end{array} $$
(A10a)
$$\begin{array}{@{}rcl@{}} &&{}\dot{D}_{\alpha \mu}h^{\alpha}_{(\sigma}\ h{\ }^{\mu}_{\varepsilon )}+h^{\alpha}_{(\sigma}\ \eta_{\varepsilon )} {\ }^{\nu \rho \tau}\ell_{\rho}H_{\tau \alpha ;\nu}=0, \end{array} $$
(A10b)
$$\begin{array}{@{}rcl@{}}\ &&{}B_{\alpha \mu ;\nu}h^{\mu \nu}h^{\alpha}_{\varepsilon}=0, \end{array} $$
(A10c)
$$\begin{array}{@{}rcl@{}} &&{}\dot{B}^{\mu \nu}h_{\mu (\sigma}\ h{\ }_{\lambda )\nu}-h_{(\sigma}^{\ \alpha}\ \eta {\ }_{\lambda )}{~}^{\nu \rho \tau}\ell_{\rho}E_{\tau \alpha; \nu}=0 \ , \end{array} $$
(A10d)

in which a parenthesis means symmetrization.

This set of equations has a striking resemblance with Maxwell’s macroscopic equations of electrodynamics. Indeed, we can formally understand the above set as having the form [79, 80]

$$\begin{array}{@{}rcl@{}} &&{}\nabla \cdot \vec D=0\ , \end{array} $$
(A11a)
$$\begin{array}{@{}rcl@{}} &&{}\dot{\vec{D}}-\nabla\times \vec{H}=0\ , \end{array} $$
(A11b)
$$\begin{array}{@{}rcl@{}} &&{}\nabla \cdot \vec{B}=0\ , \end{array} $$
(A11c)
$$\begin{array}{@{}rcl@{}} &&{}\dot{\vec{B}}+\nabla\times \vec{E}=0\ , \end{array} $$
(A11d)

where the symbol → is put over D, E, etc. only to represent its tensorial character; the ∇ operator represents the generalizations of the usual well-known differential operators.

We can therefore understand the reason for selecting the above privileged set of observers, given by the tangential vector μ. Only for this class of frames does (A7) take the form (A10aA10d). Any other observer which is in motion with respect to μ will mix into the equations of motion of the set of tensors (E α β ,B α β ) with the set of tensors (D α β ,H α β ). So, it is in this sense that there is a natural selection of observers, with respect to the equation of motion satisfied by Q α β μ ν .

1.1.4 A1.4 ε and μ States of Tension

A particular class of states of space-time occurs in the case in which there is a linear function relating the tensors B α β with H α β and E α β with D α β by the intermediary of two constants, ε and μ.

We set

$$B_{\alpha \lambda}=\mu H_{\alpha \lambda},$$
(A12a)
$$D_{\alpha \lambda}=\varepsilon H_{\alpha \lambda}. $$
(A12b)

If we put expressions (A12b) into definition (A1) of Q α β μ ν , a straightforward calculation shows that it is possible to write Q α β μ ν in terms of the Weyl tensor W α β μ ν and its “electric” and “magnetic” parts \(\mathcal {E}_{\alpha \beta }\) and \(\mathcal {H}_{\alpha \beta }\), if we identify the tensor E α β with \(\mathcal {E}_{\alpha \beta }\) and H α β with \(\mathcal {H}_{\alpha \beta }\). Then, we can write

$$\begin{array}{@{}rcl@{}} Q_{\alpha \beta} = W_{\alpha \beta}^{\ \ \ \mu \nu}+(\varepsilon -1)\ell_{[\alpha}\ \mathcal{E}_{\beta ]}^{[\mu }\ \ell^{\nu ]}\\+ (1-\mu )\eta_{\alpha \beta \rho \sigma}\mathcal{H}^{\sigma [\mu}\ \ell^{\nu ]}\ell^{\rho}\ , \end{array} $$
(A13)

where

$$\begin{array}{@{}rcl@{}} W_{\alpha \beta}^{\ \ \ \mu \nu} &=& 2\ell_{[\alpha}\ \mathcal{E}_{\beta ]}{~}^{[\mu}\ell^{\nu]}+\delta^{[\mu}_{[\alpha}\ \mathcal{E}^{\nu ]}_{\beta ]}{\kern27pt} \\ &&-\eta_{\alpha \beta \lambda \sigma}\ell^{\lambda}\mathcal{H}^{\sigma [\mu}\ell^{\nu ]}- \eta^{\mu \nu \rho \sigma}\ell_{\rho}\mathcal{H}_{\sigma [\alpha}\ \ell_{\beta ]}{} \end{array} $$
(A14)

and, consequently,

$$ \mathcal{E}_{\alpha \beta}=-W_{\alpha \mu \beta \nu}\ell^{mu}\ell^{\nu}, $$
(A15)
$$ \mathcal{H}_{\alpha \beta}=\frac{1}{2}\ \eta_{\alpha \mu}{~}^{\rho\sigma}W_{\rho \sigma \beta \lambda}\ell^{\mu}\ell^{\lambda} $$
(A16)

The resulting equations of motion (13)–(14) turn into the set:

$$\begin{array}{@{}rcl@{}} &&{}\mathcal{E}_{\alpha \mu ||\nu}h^{\mu \nu}h^{\alpha}_{\ \ \varepsilon}=0, \end{array} $$
(A17a)
$$\begin{array}{@{}rcl@{}} &&{}\varepsilon\mathcal{\dot{E}}_{\alpha \mu}h^{\alpha}_{(\sigma}\ h^{\mu}_{\varepsilon)}+h^{\alpha}_{(\sigma}\ \eta_{\varepsilon )}^{\nu \rho \tau}\ell_{\rho} \mathcal{H}_{\tau \alpha || \nu}=0, \end{array} $$
(A17b)
$$\begin{array}{@{}rcl@{}} &&{}\mathcal{H}_{\alpha \mu ||\nu}h^{\mu \nu}h^{\alpha}_{\varepsilon}=, \end{array} $$
(A17c)
$$\begin{array}{@{}rcl@{}} &&{}\mu \mathcal{\dot{H}}_{\alpha \mu}h^{\alpha}_{(\sigma}\ h^{\mu}_{\varepsilon )}-h^{\alpha}_{(\sigma}\ \eta_{\varepsilon)}^{\nu \rho \tau}\ell_{\rho}\mathcal{E}_{\tau \alpha || \nu}=0. \end{array} $$
(A17d)

By the same argument that guided us to (A10aA10d), we see from the above set that we can identify ε as being the gravitational analogue of the dielectric constant of electrodynamics and μ as being the permeability of space-time.

Now, we recognize in (A17d) Einstein’s equations for the free gravitational field for the particular case in which ε = μ = 1.Footnote 36

So, it seems natural to interpret (A17d) for the general case (ε, μ different from unity) as the equations for the gravitational field on states of space-time that are macroscopically characterized (in the sense discussed in the introduction) by the two constants ε and μ.

1.1.5 A1.5 Conformal Behavior of Q α β μ ν

A conformal transformation of the metric g α μ is given by the map

$$\begin{array}{@{}rcl@{}} &&g_{\mu \nu}(x) \longrightarrow \tilde{g}_{\mu \nu}(x)={\Omega}^{2}(x)g_{\mu \nu}(x) \\ &&g^{\mu \nu}(x) \longrightarrow \tilde{g}^{\mu \nu}(x)={\Omega}^{-2}(x)g^{\mu \nu}(x) \end{array} $$
(A18)

Since we can set \(\eta _{\alpha \beta }^{\ \ \ \mu \nu }\) as independent of the conformal transformation,

$$\tilde{\eta}_{\alpha \beta}^{\ \ \ \mu \nu}=\eta_{\alpha \beta}^{\ \ \ \mu \nu} $$

It is then easy to see that the electric and magnetic parts of Weyl tensor remain unchanged,

$$\begin{array}{@{}rcl@{}} &&\tilde{\mathcal{E}}_{\mu \nu}=-\tilde{W}_{\mu \rho \nu \sigma}\tilde{\ell}^{\rho}\tilde{\ell}^{\dot{\sigma}}= \mathcal{E}_{\mu\nu}\ , \end{array} $$
(A19a)
$$\begin{array}{@{}rcl@{}} &&\tilde{\mathcal{H}}_{\mu \nu}=-\frac{1}{2}\ \tilde{\eta}_{\mu \alpha}^{\ \ \ \rho \sigma} \tilde{W}_{\rho \sigma \nu \lambda}\tilde{\ell}^{\alpha}\tilde{\ell}^{\lambda}=\mathcal{H}_{\mu \nu}, \end{array} $$
(A19b)

where we have used conformal transformation of the velocity μ as usual,

$$ \tilde{\ell}^{\mu}={\Omega}^{-1}\ell^{\mu} $$
(A20)

As a consequence of the above mapping, Q α β μ ν behaves, under the conformal transformation, as the Weyl tensor,

$$ \tilde{Q}_{\alpha \beta \mu \nu}(x)={\Omega}^{2}(x)Q_{\alpha \beta \mu \nu}(x). $$
(A21)

1.2 A2 Gravitational Energy in an εμ State of Tension

There have been many discussions, since Einstein’s [46, 47] paper, concerning the definition of the energy of a given gravitational field. We do not intend to discuss this subject here but we shall limit ourselves to considering one reasonably successful suggestion by Bel [12] for the form of the energy-momentum tensor of gravitational radiation.

The point of departure [12] comes from the similitude of the equation of motion of gravity and electrodynamics. He defines a fourth-rank tensor T αβμν given in terms of quadratic components of the field (identified with the Riemann tensor) and written in terms of the Weyl tensor W αβμν.

Bel’s super-energy tensor takes the form:

$$ T^{\alpha \beta \mu \nu}=\frac{1}{2}\left\{W^{\alpha \rho \mu \sigma}W^{\beta}{}{~}_{\rho}{~}^{\nu}{}{~}_{\sigma}+ C^{\ast \alpha \rho \mu \sigma} C^{\ast \beta}{}{~}_{\rho}{~}^{\nu}{}{~}_{\sigma}\right\}\ , $$
(A22)

where is the dual operator. Note that the symmetry of the Weyl tensor (\(C^{\,\ast }_{\alpha \beta \mu \nu } = W^{\,\,\,\,\,\,\,\,\,\ast }_{\alpha \beta \mu \nu } = \,^{\ast }W_{\alpha \beta \mu \nu }\)) does not hold for Q α β μ ν . This is related to the lack of Q α β μ ν Q μ ν α β symmetry. Indeed, we have that

$$\begin{array}{@{}rcl@{}} Q^{\ast}_{\alpha\beta}{~}^{\mu \nu}&=&W^{\ast}_{\alpha \beta}{~}^{\mu \nu}+\frac{1}{2}(\varepsilon -1) \eta_{\alpha \beta \rho \sigma}\ell^{[\rho}\ \mathcal{E}^{\sigma ]}{~}^{[\mu}\ \ell^{\nu ]}\\ &&+\frac{1}{2}\ (1-\mu )\eta_{\alpha \beta \rho \sigma}\eta^{\rho \sigma \varepsilon \sigma}\ell_{\varepsilon} \mathcal{H}_{\tau}^{[\mu}\ \ell^{\nu]}{\kern23pt} \end{array} $$

and

$$\begin{array}{@{}rcl@{}} Q_{\alpha\beta}{~}^{\mu^{\ast} \nu}&=&C^{\,\ast}_{\alpha \beta}{~}^{\mu \nu}+\frac{1}{2}(\varepsilon -1) \eta^{\mu \nu}_{\ \ \rho \sigma}\ell^{[\alpha}\ \mathcal{E}^{\beta ]}{~}^{[\rho}\ \ell^{\sigma ]}\\ &&+\frac{1}{2}\ (1-\mu )\eta^{\mu \nu}{}{~}_{\rho \sigma}\eta_{\alpha \beta \varepsilon \tau}\ell^{\varepsilon} \mathcal{H}^{\tau}{~}^{[\rho}\ \ell^{\sigma ]}{\kern17pt} \end{array} $$

Then we have that

$$\begin{array}{@{}rcl@{}} Q_{\alpha \beta \mu\nu}^{\,\,\,\,\,\,\,\,\ast}\ell^{\beta}\ell^{\nu}&=&C^{\,\ast}_{\alpha \beta \mu \nu}\ell^{\beta}\ell^{\nu}=\mathcal{H}_{\alpha \mu}\ ,\\ Q^{\,\ast}_{\alpha\beta \varepsilon \sigma}\ell^{\beta}\ell^{\sigma}&=&\mu\ \mathcal{H}_{\alpha \varepsilon}\ . \end{array} $$

This T αβμν tensor has properties that are very similar to the Minkowski energy-momentum tensor ofelectrodynamics. The scalar constructed with T αβμν and the tangent vector ν, for instance, takes the form

$$ U_{(T)}=T^{\alpha \beta \mu \nu}\ell_{\alpha}\ell_{\beta}\ell_{\mu}\ell_{\nu} $$
(A23)

and gives the “energy” of the field

$$ U_{(T)}=\frac{1}{2}(\mathcal{E}^{2}+\mathcal{H}^{2}), $$
(A24)

where

$$\begin{array}{@{}rcl@{}} &&\mathcal{E}^{2}=\mathcal{E}_{\alpha \beta}\mathcal{E}^{\alpha \beta}. \end{array} $$
(A25a)
$$\begin{array}{@{}rcl@{}} &&\mathcal{H}^{2}=\mathcal{H}_{\alpha \beta}\mathcal{H}^{\alpha \beta}. \end{array} $$
(A25b)

In the context of the present extended theory, for a space-time in the state εμ of tension, we are led to modify T αβμν into Θαβμν defined in an analogous manner by the equality

$$ {\Theta}^{\alpha \beta \mu \nu} = \frac{1}{2}\left\{Q^{\alpha \rho \mu \sigma}C^{\beta}{}{~}_{\rho}{~}^{\nu}{}{~}_{\sigma} + Q^{\alpha^{\ast}\rho\mu\sigma} {~}^{\ast}C^{\beta}{}{~}_{\rho}{~}^{\nu}{}{~}_{\sigma}\right\}. $$
(A26)

Then, the energy U (ε, μ) as viewed by an observer μ will be given by the relation

$$ U_{(\varepsilon ,\mu )}=\theta^{\alpha \beta \mu \nu}\ell_{\alpha}\ell_{\beta}\ell_{\mu}\ell_{\nu}=\frac{1}{2} \left(\varepsilon\mathcal{E}^{2}+\mu \mathcal{H}^{2}\right) $$
(A27)

in complete analogy with the electrodynamical case in a general medium.

We would like to make an additional remark by presenting two special properties of Θαβμν

$$\begin{array}{@{}rcl@{}} &&{}{\Theta}^{\alpha}_{\ \ \beta \alpha \mu}=\frac{1}{2}(1-\varepsilon )\varepsilon^{\rho \sigma}C_{\beta \rho \mu \sigma}\ , \end{array} $$
(A28a)
$$\begin{array}{@{}rcl@{}} &&{}{\Theta} = {\Theta}^{\alpha \mu}{}{~}_{\alpha \mu}=0 \ . \end{array} $$
(A28b)

Property (A28a) states that not all traces of Θαβμν are null for a general state of tension of space-time and that the non-null parts of the contracted tensor are independent of the “permeability” μ. The second property (A28b) states that the scalar obtained by taking the trace of Θαβμν twice is null, independent from the state of tension of the space-time.

1.3 A3 The Velocity of Propagation of Gravitational Disturbances in States of Tension

To determine the velocity of gravitational waves in εμ states of space-time, let us perturb the set of equations (A10aA10d). The perturbation will be represented by the map:

$$\begin{array}{@{}rcl@{}} &&\mathcal{E}_{\mu \nu}\longrightarrow \mathcal{E}_{\mu \nu}+\delta \mathcal{E}_{\mu \nu}\ ,\\ &&\mathcal{H}_{\mu \nu}\longrightarrow \mathcal{H}_{\mu \nu}+\delta \mathcal{H}_{\mu \nu} \, \end{array} $$
(A29a) (A29b)

where \(\delta \mathcal {E}_{\mu \nu }\), \(\delta \mathcal {H}_{\mu \nu }\) are null quantities. Then, (A10aA10d) are transformed into the perturbed set of equations

$$\delta \mathcal{E}_{\alpha}{~}^{\beta}{}{~}_{;\beta}\approx 0 \ ,$$
(A30a)
$$\varepsilon \delta \dot{\mathcal{E}}_{\alpha \mu}+\frac{1}{2}\ h_{(\alpha}^{\lambda}\eta_{\mu )}{~}^{\rho\sigma \tau}\ell_{\rho}\,\delta \mathcal{H}_{\tau \lambda;\rho}\approx 0 \ , $$
(A30b)
$$\delta \mathcal{H}_{\alpha}{~}^{\beta}{}{~}_{;\beta}\approx 0 \ , $$
(A30c)
$$\mu \ \delta \dot{\mathcal{H}}-\frac{1}{2}\ h_{(\alpha}^{\lambda}\eta_{\mu )}{~}^{\rho \sigma \tau}\mathcal{E}_{\tau \lambda;\rho}\approx 0\ , $$
(A30d)

where the covariant derivative is taken in the background—and we limit ourselves to the linear terms of the perturbation.

Now, let us specialize the background to be the flat Minkowski space-time.Footnote 37 In this case, the covariant derivatives are the usual derivation and we can use commutative property to write:

$$ \varepsilon \delta {\mathcal{\ddot{E}}}_{\alpha \beta}+\frac{1}{2}\ h_{(\alpha}^{\lambda}\eta_{\beta)}^{\rho \sigma \tau} \ell_{\sigma}\delta \mathcal{\dot{H}}_{\tau \lambda,\rho}\approx 0 \, $$
(A31)

by taking the derivative of (A30b) projected in the privileged direction μ.

Multiplying (A30d) by the factor

$$\frac{1}{2 \mu}\ h_{(\alpha}^{\nu} \eta_{\beta )}{~}^{\gamma\sigma\tau}\ell_{\tau}\ \frac{\partial}{\partial x^{\sigma}}\ , $$

we find that

$$\begin{array}{@{}rcl@{}} &&\frac{1}{2}h_{(\alpha }^{\nu}\ \eta_{\beta )}{~}^{\gamma\sigma \tau}\ell_{\tau}\,\delta \dot{\mathcal{H}}_{\gamma \nu, \sigma}\\ &&{\kern6pt}-\frac{1}{4\mu}\ h_{(\alpha }^{\nu}\ \eta_{\beta )}{~}^{\sigma \tau \gamma}\ \ell_{\tau}\ell_{\rho}h^{\varepsilon}{}{~}_{(\,\gamma}\ \eta_{\nu)}{~}^{\psi \rho \phi}\delta \varepsilon_{\psi \varepsilon ,\psi ,\sigma}\approx 0.\\ \end{array} $$
(A32)

Substituting (A31) into (A31) we finally find that

$$ \delta \ddot{\mathcal{E}}_{\alpha \mu}-\frac{1}{\varepsilon \mu}\nabla^{2}\delta\mathcal{E}_{\alpha \mu}=0 \ , $$
(A33)

where ∇2 is the Laplacian operator defined in the three-dimensional space orthogonal to μ.

In the same way, an analogous wave equation can be obtained for \(\mathcal {H}_{\alpha \mu }\). From (A33), we obtain the expected result: the velocity of propagation of gravitational waves in ε, μ states of tension of space-time is equal to \(1/\sqrt {\varepsilon \mu }\).

The set of privileged observers that we dealt with here may be enlarged by somehow weakening the defining conditions [see (A8aA8d)]. This point deserves further investigation.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Novello, M., Bittencourt, E. & Salim, J.M. The Quasi-Maxwellian Equations of General Relativity: Applications to Perturbation Theory. Braz J Phys 44, 832–894 (2014). https://doi.org/10.1007/s13538-014-0239-1

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13538-014-0239-1

Keywords

Navigation