Skip to main content
Log in

Gravito-electromagnetic analogies

  • Research Article
  • Published:
General Relativity and Gravitation Aims and scope Submit manuscript

Abstract

We reexamine and further develop different gravito-electromagnetic analogies found in the literature, and clarify the connection between them. Special emphasis is placed in two exact physical analogies: the analogy based on inertial fields from the so-called “1+3 formalism”, and the analogy based on tidal tensors. Both are reformulated, extended and generalized. We write in both formalisms the Maxwell and the full exact Einstein field equations with sources, plus the algebraic Bianchi identities, which are cast as the source-free equations for the gravitational field. New results within each approach are unveiled. The well known analogy between linearized gravity and electromagnetism in Lorentz frames is obtained as a limiting case of the exact ones. The formal analogies between the Maxwell and Weyl tensors are also discussed, and, together with insight from the other approaches, used to physically interpret gravitational radiation. The precise conditions under which a similarity between gravity and electromagnetism occurs are discussed, and we conclude by summarizing the main outcome of each approach.

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

Similar content being viewed by others

Notes

  1. We want to emphasize this point, which, even today, is not clear in the literature. Equations (1.1) apply to the instant where the two particles have the same (or infinitesimally close, in the gravitational case) tangent vector. When the particles have arbitrary velocities, both in electromagnetism and gravity, their relative acceleration is not given by a simple contraction of a tidal tensor with a separation vector; the equations include more terms, see [1, 2, 55]. There is however a difference: whereas Eq. (1.1a) requires strictly \(\delta \mathbf {U}=\mathbf {U}_{2}-\mathbf {U}_{1}=0\), see [1, 2, 98], Eq. (1.1b) allows for an infinitesimal \(\delta U\propto \delta x\), as can be seen from Eq. (6) of [55]. That means that (1.1b) holds for infinitesimally close curves belonging to an arbitrary geodesic congruence (it is in this sense that in e.g., [54, 97] \(\delta U\) is portrayed as “arbitrary”—it is understood to be infinitesimal therein, as those treatments deal with congruences of curves).

  2. The characterization of the Riemann tensor by these three spatial rank 2 tensors is known as the “Bel decomposition”, even though the explicit decomposition (15) is not presented in any of Bel’s papers (e.g., [5]). To the author’s knowledge, an equivalent expression (Eq. (4.6) therein) can only be found at [86].

  3. We thank João Penedones for drawing our attention to this point.

  4. By rotation we mean here absolute rotation, i.e, measured with respect to a comoving Fermi–Walker transported frame. As one can check from the connection coefficients (29) below, in such frame (\(\Omega _{\alpha \beta }=0\)) we have \(d^{2}\delta x^{\hat{i}}/d\tau ^{2}=D^{2}\delta x^{\hat{i}}/d\tau ^{2}\). See also in this respect [110].

  5. If the two particles were connected by a “rigid” rod then the symmetric part of the electric tidal tensor would also, in general, torque the rod; hence in such system we would have a rotation even in the gravitational case, see [25] pp. 154–155. The same is true for a quasi-rigid extended body; however, even in this case the effects due to the symmetric part are very different from the ones arising from electromagnetic induction: first, the former do not require the fields to vary along the particle’s worldline, they exist even if the body is at rest in a stationary field; second, they vanish if the body is spherical, which does not happen with the torque generated by the induced electric field, see [6].

  6. This example is particularly interesting in this discussion. In the electromagnetic analogous problem, a magnetic dipole in (initially) radial motion in the Coulomb field of a point charge experiences a force; that force, as shown in [6], comes entirely from the antisymmetric part of the magnetic tidal tensor, \(B_{\alpha \beta }=B_{[\alpha \beta ]}\); it is thus a natural realization of the arguments above that \(\mathbb {H}_{\alpha \beta }=0\) in the analogous gravitational problem.

  7. Note however that in some literature, e.g., [110], the term “congruence adapted” is employed with a different meaning, designating any tetrad field whose time axis is tangent to the congruence, without any restriction on the transport law for the spatial triad (namely without requiring the triads to co-rotate with the congruence). Hence “adapted” therein means what, in our convention, we would call adapted to each individual observer.

    Fig. 1
    figure 1

    The reference frame: a congruence of of time-like curves—the observers’ worldlines—of tangent \(\mathbf {u}\); each observer carries an orthonormal tetrad \(\mathbf {e}_{\hat{\alpha }}\) such that \(\mathbf {e}_{\hat{0}}=\mathbf {u}\) and the spatial triad \(\mathbf {e}_{\hat{i}}\) spans the observer’s local rest space. The triad \(\mathbf {e}_{\hat{i}}\) rotates, relative to Fermi–Walker transport, with some prescribed angular velocity \(\vec {\Omega }\)

  8. e.g., the trivial variation from point to point of the triads associated with a non-rectangular coordinate system in flat spacetime. These do not encode inertial forces, nor do they necessarily vanish in an inertial frame.

  9. This corresponds to a generalized version, for arbitrary orthonormal frames, of Eqs. (6.13) or (6.18) of [27], which in their scheme would follow from a “derivative” of the type (5.3), but allowing for an arbitrary \(\Omega ^{\alpha }\), rather than the two choices \(\Omega ^{\alpha }=0\) and \(\Omega ^{\alpha }=\omega ^{\alpha }\) (“fw” and “cfw” in their notation, respectively), cf. Eq. (2.16). On the other hand, their Lie transport option (“lie”) in (5.3), which does not preserve orthonormality of the axes, is not encompassed in our derivative (54).

  10. Somewhat erroneously, as the tetrads do rotate with respect to the local compass of inertia, since they are not Fermi–Walker transported in general [54, 101, 102].

  11. In [26] the term involving \(\Omega ^{\alpha }\) in Eq. (56) above is cast not as part of a gravitomagnetic, but of a “Coriolis” acceleration (“\(a_{\mathrm{C}}^{\alpha }\)”). Therein, what is cast as “gravitomagnetic” (“\(a_{\mathrm{d}}^{\alpha }\)”), are the terms involving \(K_{(\alpha \beta )}\) and \(\omega ^{\alpha }\).

  12. In the case of Kerr spacetime, these are the observers whose worldlines are tangent to the temporal Killing vector field \(\xi =\partial /\partial t\), i.e., the observers of zero 3-velocity in Boyer-Lindquist coordinates. This agrees with the convention in [28, 54, 95, 99]. We note however that the denomination “static observers” is employed in some literature (e.g., [82, 83]) with a different meaning, where it designates hypersurface orthogonal time-like Killing vector fields (which are rigid, vorticity-free congruences, existing only in static spacetimes).

  13. This is manifest in the algebraic Bianchi identities. The generalization of Eq. (95) for non-congruence adapted frames is \(\star \tilde{R}_{\ ji}^{j}=2\epsilon _{ikj}\omega ^{j}\Omega ^{k}-2K_{(ik)}\omega ^{k}\); the first term is not zero in general when \(\vec {\Omega }\ne \vec {\omega }\) and/or both \(K_{(ij)}\) and \(\omega ^{k}\) are different from zero.

  14. Eqs. (93)-(94) are equivalent to Eq. (7.3) of [27]; therein they are obtained through a different procedure, not by projecting the identity \(\star R_{\, \, \,\,\,\, \gamma \beta }^{\gamma \alpha }=0\Leftrightarrow R_{[\alpha \beta \gamma ]\delta }=0\), but instead from the splitting of the identity \(d^{2}\mathbf {u}=0\Leftrightarrow u_{[\alpha ;\beta \gamma ]}=0\). Noting that \(u_{[\alpha ;\beta \gamma ]}=-R_{[\alpha \beta \gamma ]\lambda }u^{\lambda }\), we see that the latter is indeed encoded in the time–time and space–time parts (with respect to \(u^{\alpha }\)) of the former.

  15. Our conventions relate to the ones in [104] by identifying our \(\{K_{(ij)},\theta \}\) with \(\{-K_{ij},-K\}\) therein.

  16. One could also obtain it directly from the covariant version Eqs. (54) and (56), by setting therein \(\Omega ^{\alpha }=\omega ^{\alpha }=H^{\alpha }/2\), noting that, to linear order \(\Gamma ^i_{0k}=\varepsilon ^i_{\,jk}H^j/2+\partial \xi ^i_{\,\,\,k}/\partial t\), and using (122), as done below.

  17. To make contact with the notation in [34, 35], we note that the restriction to the spatial directions of both \(\tilde{\nabla }\) and \(\nabla ^{\perp }\) yields the 3-D connection “\(\bar{\nabla }\)” of [35] (“\(D\)” of [34]).

  18. The wave equations in [107109] are obtained using also the Ricci identities \(2\nabla _{[\gamma }\nabla _{\beta ]}X_{\alpha }=R_{\delta \alpha \beta \gamma }X^{\delta }\), which couple the electromagnetic fields to the curvature tensor; this coupling is shown to lead to amplification phenomena, suggested therein as a possible explanation for the observed (and unexplained) large-scale cosmic magnetic fields.

  19. We thank J. Penedones for discussions on this point.

References

  1. Costa, L.F., Herdeiro, C.A.R.: Phys. Rev. D 78, 024021 (2008)

    Article  MathSciNet  ADS  Google Scholar 

  2. Costa, L.F., Herdeiro, C.A.R.: gr-qc/0612140

  3. Corda, C.: Phys. Rev. D 83, 062002 (2011)

    Article  ADS  Google Scholar 

  4. Nichols, D., et al.: Phys. Rev. D 84, 124014 (2011)

    Article  ADS  Google Scholar 

  5. Bel, L., Acad, C.R.: Sci. Paris 246, 3015 (1958)

    MATH  Google Scholar 

  6. Costa, L.F., Natário, J., Zilhão, M.: [ arXiv:1207.0470]

  7. Ciufolini, I., Wheeler, J.A.: Gravitation and Inertia, Princeton Series in Physics. Princeton University, Princeton, NJ (1995)

  8. Ohanian, H.C., Ruffini, R.: Gravitation and Spacetime, 2nd ed. W.W. Norton, New York (1994)

  9. Harris, E.G.: Am. J. Phys. 59, 421 (1991)

    Article  ADS  Google Scholar 

  10. Clark, S.J., Tucker, R.W.: Class. Quant. Grav. 17, 4125 (2000)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  11. Ruggiero, M.L., Tartaglia, A.: Il Nuovo Cimento B 117, 743 (2002)

    ADS  Google Scholar 

  12. Wald, R.M.: General Relativity. University of Chicago Press, Chicago (1984). Sec. 4.4

    Book  MATH  Google Scholar 

  13. Wald, R.M.: Phys. Rev. D 6, 406 (1972)

    Article  ADS  Google Scholar 

  14. Carroll, S.M.: Spacetime and Geometry. Addison-Wesley, Boston (2003). Sec. 7.2.

    Google Scholar 

  15. Thorne, Kip S.: In: Fairbank, J.D., Deaver Jr., B.S., Everitt, C.W.F., Michelson, P.F. (eds.) Near Zero: New Frontiers of Physics. W. H. Freeman and Company, NY (1988)

  16. Gralla, S., Harte, A.I., Wald, R.M.: Phys. Rev. D 81, 104012 (2010)

    Article  ADS  Google Scholar 

  17. Mashhoon, B.: In: Pascual-Sanchez, J.-F., Floria, L., San Miguel, A. (eds.) Reference Frames and Gravitomagnetism, p. 121. World Scientific, Singapore (2001). gr-qc/0011014

  18. Landau, L., Lifshitz, E.: The Classical Theory of Fields, 4th edn. Elsevier, Amsterdam (1975)

    Google Scholar 

  19. Oliva, W.: Geometric Mechanics, Lect. Not. Math. 1798, Springer, Appendix C (2002)

  20. Natário, José: Gen. Relativ. Gravit 39, 1477 (2007)

    Article  MATH  ADS  Google Scholar 

  21. Costa, J., Natário, J.: J. Math. Phys. 46, 082501 (2005)

    Article  MathSciNet  ADS  Google Scholar 

  22. Mena, F.C., Natário, J.: J. Geom. Phys. 59, 448 (2009)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  23. Lynden-Bell, D., Nouri-Zonoz, M.: Rev. Mod. Phys. 70, 427 (1998)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  24. Nouri-Zonoz, M.: Phys. Rev. D 60, 024013 (1999)

    Article  MathSciNet  ADS  Google Scholar 

  25. Thorne, Kip S., Price, R.H., Macdonald, D.A.: Black Holes, the Membrane Paradigm. Yale University Press, New Haven (1986)

    Google Scholar 

  26. Semerák, O.: Il Nuovo Cim. B 110, 973 (1995)

    Article  ADS  Google Scholar 

  27. Jantzen, R.T., Carini, P., Bini, D.: Ann. Phys. 215, 1 (1992)

    Article  MathSciNet  ADS  Google Scholar 

  28. Jantzen, R.T., Carini, P., Bini, D.: GEM: The User Manual (2004). http://www34.homepage.villanova.edu/robert.jantzen/gem/gem_grqc

  29. Costa, L.F.O., Natário, J., Wylleman, L.: Work in progress (for a brief summary of the main results). J. Phys. Conf. Ser. 314, 012072 (2011)

  30. Pascual Sánchez, J.-F.: Il Nuovo Cimento B 115, 725 (2000)

  31. Matte, A.: Canadian J. Math. 5, 1 (1953)

    Article  MathSciNet  MATH  Google Scholar 

  32. Bel, L.: Cahiers de Physique 16, 59 (1962)

    MathSciNet  ADS  Google Scholar 

  33. Bel, L.: Gen. Relativ. Gravit 32, 2047 (2000)

  34. Tchrakian, D.H.: Gen. Relativ. Gravit 6, 151 (1975)

    Article  MathSciNet  ADS  Google Scholar 

  35. Campbell, W., Morgan, T.: Am. J. Phys. 44, 356 (1976)

    Article  ADS  Google Scholar 

  36. Maartens, R., Bassett, B.: Class. Quant. Grav. 15, 705 (1998)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  37. Ellis, G.F.R., Maartens, R., MacCallum, M.A.H.: Relativistic Cosmology. Cambridge University Press, Cambridge (2012)

    Book  MATH  Google Scholar 

  38. Ellis, G.F.R.: General relativity and cosmology. In: Sachs, B.K. (ed.) Proceedings of the International School of Physics Enrico Fermi, Course XLVII (1971)

  39. Ellis, G.F.R., Hogan, P.A.: Gen. Relativ. Gravit 29, 235 (1997)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  40. Bonnor, W.B.: Class. Quant. Grav. 12, 499 (1995)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  41. Cherubini, C., Bini, D., Capozziello, S., Ruffini, R.: Int. J. Mod. Phys. D 11, 827 (2002)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  42. McIntosh, C.B.G., Arianhod, R., Wade, S.T., Hoenselaers, C.: Class. Quant. Grav. 11, 1555 (1994)

    Article  MATH  ADS  Google Scholar 

  43. Wylleman, L., Van den Bergh, N.: Phys. Rev. D 74, 084001 (2006)

    Article  ADS  Google Scholar 

  44. Bel, L.: Ann. de l’ I. H. P. 17, 37 (1961)

    MathSciNet  MATH  Google Scholar 

  45. Zakharov, V.D.: Gravitational Waves in Einstein’s Theory, Eng. Trans. by R. N. Sen. Halsted Press, Wiley, New York (1973)

  46. Dadhich, Naresh: Gen. Relativ. Gravit 32, 1009 (2000)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  47. Van Elst, H., Uggla, C.: Class. Quant. Grav. 14, 2673 (1997)

    Article  MATH  ADS  Google Scholar 

  48. Mathisson, M.: Acta Phys. Pol. 6, 163 (1937)

    Google Scholar 

  49. Mathisson, M.: Gen. Relativ. Gravit 42, 1011 (2010)

  50. Dixon, W.G.: J. Math. Phys. 8, 1591 (1967)

    Article  ADS  Google Scholar 

  51. Papapetrou, A.: Proc. R. Soc. London A 209, 248 (1951)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  52. Pirani, F.A.E.: Acta Phys. Pol. 15, 389 (1956)

    MathSciNet  ADS  Google Scholar 

  53. Costa, L.F., Herdeiro, C.: Relativity in fundamental astronomy: dynamics, reference frames, and data analysis. Klioner, S.A., Seidelmann, P.K., Soffel, M.H. (eds.) Proceedings of the International Astronomical Union, vol. 5, S261, pp. 31–39. Cambridge University Press, Cambridge (2010). Preprint [ arXiv:0912.2146]

  54. Misner, Charles W., Thorne, KipS, Wheeler, John A.: Gravitation. W. H. Freeman and Company, San Francisco (1973)

    Google Scholar 

  55. Ciufolini, I.: Phys. Rev. D 34, 1014 (1986)

    Article  MathSciNet  ADS  Google Scholar 

  56. Massa, E., Zordan, C.: Meccanica 10, 27 (1975)

    Article  MathSciNet  Google Scholar 

  57. Massa, E.: Gen. Relativ. Gravit 5, 555 (1974)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  58. Massa, E.: Gen. Relativ. Gravit 5, 573 (1974)

    Article  MathSciNet  ADS  Google Scholar 

  59. Bini, D., Carini, P., Jantzen, R.T., Wilkins, D.: Phys. Rev. D 49, 2820 (1994)

    Article  ADS  Google Scholar 

  60. Ciufolini, I., Pavlis, E.: Nature 431, 958 (2004)

    Article  ADS  Google Scholar 

  61. Ciufolini, I., Pavlis, E., Peron, R.: New Astron. 11, 527 (2006)

    Article  ADS  Google Scholar 

  62. Review of Gravity Probe B. National Academy Press, Washington (1995)

  63. Everitt, F., et al.: Phys. Rev. Lett. 106, 221101 (2011). http://einstein.stanford.edu/

  64. Ciufolini, I., et al.: Space Sci. Rev. 148, 71 (2009). http://www.lares-mission.com/

  65. Kaplan, J., Nichols, D.: Kip Thorne. Phys. Rev. D 80, 124014 (2009)

    Article  ADS  Google Scholar 

  66. Damour, T., Soffel, M., Xu, C.: Phys. Rev. D. 43, 3273 (1991)

    Article  MathSciNet  ADS  Google Scholar 

  67. Deser, S., Franklin, J.: Am. J. Phys. 75, 281 (2007)

    Article  ADS  Google Scholar 

  68. Bailey, Q.G.: Phys. Rev. D 82, 065012 (2010)

    Article  ADS  Google Scholar 

  69. Semerák, O.: Gen. Relativ. Gravit 30, 1203 (1998)

    Article  MATH  ADS  Google Scholar 

  70. Drukker, N., Fiol, B., Simon, J.: JCAP 0410, 012 (2004)

    Article  MathSciNet  ADS  Google Scholar 

  71. Nordtvedt, K.: Phys. Rev. D 7, 2347 (1973)

    Article  ADS  Google Scholar 

  72. Hawking, S.W.: Astrophys. J. 145, 544 (1966)

    Article  ADS  Google Scholar 

  73. Senovilla, J.M.M.: Class. Quantum Grav. 17, 2799 (2000)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  74. Komar, A.: Phys. Rev. 164, 1595 (1967)

    Article  ADS  Google Scholar 

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

    Article  MathSciNet  MATH  ADS  Google Scholar 

  76. Garecki, J.: Acta. Phys. Pol. B 8, 159 (1977)

    MathSciNet  Google Scholar 

  77. Teyssandier, P.: [ gr-qc/9905080]

  78. Lozanovski, C., Mcintosh, C.B.G.: Gen. Relativ. Gravit 31, 1355 (1999)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  79. Lozanovski, C., Aarons, M.: Class. Quant. Grav. 16, 4075 (1999)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  80. Hawking, S., Ellis, G.: The Large Scale Structure of Spacetime. Cambridge University Press, Cambridge (1974)

    Google Scholar 

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

    Article  MathSciNet  MATH  ADS  Google Scholar 

  82. Wylleman, L., Beke, D.: Phys. Rev. D 81, 104038 (2010)

    Article  ADS  Google Scholar 

  83. Dahia, F., Silva, P.F.: [ arXiv:1004.3937]

  84. Campbell, W., Macek, J., Morgan, T.: Phys. Rev. D 15, 2156 (1977)

    Article  MathSciNet  ADS  Google Scholar 

  85. Ciufolini, I.: Nature 449, 41 (2007)

    Article  ADS  Google Scholar 

  86. Gómez-Lobo, A. García-Parrado.: Class. Quant. Grav. 25, 015006 (2008)

  87. Ferrando, J.J., Sáez, J.A.: Class. Quant. Grav. 29, 075012 (2012)

    Article  ADS  Google Scholar 

  88. Rizzi, G., Ruggiero, M.L.: Relativity. In: Rizzi, G., Ruggiero, M.L. (eds.) Rotating Frames. Kluwer Academic Publishers, Dordrecht (2004). preprint [gr-qc/0305084]

    Chapter  Google Scholar 

  89. Rizzi, G., Ruggiero, M.L.: Gen. Relativ. Gravit 35, 1743 (2003)

    MathSciNet  ADS  Google Scholar 

  90. Ruggiero, M.L.: Gen. Relativ. Gravit 37, 1845 (2005)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  91. Bel, L.: Colloques Internationaux du Centre national de la reserche scientifique, p. 119. Paris (1962)

  92. Soffel, M., Klioner, S., Muller, J., Biskupek, L.: Phys. Rev. D 78, 024033 (2008)

    Article  ADS  Google Scholar 

  93. Murphy Jr, T.W., Nordtvedt, K., Turyshev, S.G.: Phys. Rev. Lett. 98, 071102 (2007)

    Article  ADS  Google Scholar 

  94. Nordtvedt, K.: Int. J. Theor. Phys. 27, 2347 (1988)

    Article  Google Scholar 

  95. Bini, D., Geralico, A., Jantzen, R.: Class. Quant. Grav. 22, 4729 (2005)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  96. Mason, D., Pooe, C.: J. Math. Phys. 28, 2705 (1987)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  97. Dixon, W.G.: Proc. Roy. Soc. Lond. A 314, 499 (1970)

    Article  ADS  Google Scholar 

  98. Balakin, A., Van Holten, J., Kerner, R.: Class. Quant. Grav. 17, 5009 (2000)

    Article  MATH  ADS  Google Scholar 

  99. Bardeen, J.M.: Astrophys. J. 162, 71 (1970)

    Article  MathSciNet  ADS  Google Scholar 

  100. Bardeen, J.M., Press, W.H., Teukolsky, S.A.: Astrophys. J. 178, 347 (1972)

    Article  ADS  Google Scholar 

  101. Semerák, O.: Gen. Relativ. Gravit 25, 1041 (1993)

    Article  MATH  ADS  Google Scholar 

  102. Semerák, O.: Class. Quant. Grav. 13, 2987 (1996)

    Article  MATH  ADS  Google Scholar 

  103. Iorio, L., Lichtenegger, H., Ruggiero, M.L., Corda, C.: Astrophys. Space Sci 331, 351 (2010)

    Article  ADS  Google Scholar 

  104. Eric Gourgoulhon, 3+1 Formalism and Bases of Numerical Relativity, [gr-qc/0703035]

  105. Arnowitt, R., Deser, S., Misner, C.W.: In: Witten, L. (ed.) Gravitation: An Introduction to Current Research, pp. 227–264. Wiley, New York (1962). preprint [gr-qc/0405109]

  106. Baumgarte, T.W., Shapiro, S.L.: Phys. Rep. 376, 41 (2003)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  107. Tsagas, C.: Class. Quant. Grav. 22, 393 (2005)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  108. Tsagas, C.: Phys. Rev. D 81, 043501 (2010)

    Article  ADS  Google Scholar 

  109. Barrow, J., Tsagas, C., Yamamoto, K.: Phys. Rev. D 86, 023533 (2012)

    Article  ADS  Google Scholar 

  110. Bini, D., de Felice, F., Geralico, A.: Class. Quant. Grav. 23, 7603 (2006)

    Article  MATH  ADS  Google Scholar 

  111. Kay, David C.: Tensor Calculus. McGraw-Hill, New York (1988)

    Google Scholar 

  112. Goldstein, H., Poole, C., Safko, J.: Classical Mechanics, 3rd edn. Addison-Wesley, New York (2000)

    Google Scholar 

Download references

Acknowledgments

We thank J. Penedones, the anonymous referees and A. Editor for useful comments and remarks; we also thank A. García-Parrado and J. M. M. Senovilla for correspondence and useful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to L. Filipe O. Costa.

Appendix A: Inertial forces—simple examples in flat spacetime

Appendix A: Inertial forces—simple examples in flat spacetime

In Sect.  we have seen that the inertial forces felt in a given frame arise from two independent contributions of different origin: the kinematics of the observer congruence (that is, from the derivatives of the temporal basis vector of the frame, \(\mathbf {e}_{\hat{0}}=\mathbf {u}\), where \(\mathbf {u}\) is the observers’ 4-velocity), and the transport law for the spatial triads \(\mathbf {e}_{\hat{i}}\) along the congruence. In order to illustrate these concepts with simple examples, we shall consider, in flat spacetime, the straightline geodesic motion of a free test particle (4-velocity \(\mathbf {U}\)), from the point of view of three distinct frames: a) a frame whose time axis is the 4-velocity of a congruence of observers at rest, but whose spatial triads rotate uniformly with angular velocity \(\vec {\Omega }\); b) a frame composed of a congruence of rigidly rotating observers (vorticity \(\vec {\omega }\)), but carrying Fermi–Walker transported spatial triads (\(\vec {\Omega }=0\)); c) a rigidly rotating frame, that is, a frame composed of a congruence of rigidly rotating observers, carrying spatial triads co-rotating with the congruence \(\vec {\Omega }=\vec {\omega }\) (i.e., “adapted” to the congruence, see Sect. 3.1). This is depicted in Fig. 2.

Fig. 2
figure 2

A test particle in uniform motion in flat spacetime from the point of view of three different frames: (a) a frame composed of observers at rest, but carrying spatial triads that rotate with uniform angular velocity \(\vec {\Omega }\); (b) a frame consisting of a congruence of rigidly rotating observers (vorticity \(\vec {\omega }\)), but each of them carrying a non-rotating spatial triad (i.e., that undergoes Fermi–Walker transport); (c) a rigidly rotating frame (a frame adapted to a congruence of rigidly rotating observers); the spatial triads co-rotate with the congruence, \(\vec {\Omega }=\vec {\omega }\). Note: by observer’s rotation we mean their circular motion around the center; and by axes rotation we mean their rotation (relative to FW transport) about the local tetrad’s origin

In the first case there we have a vanishing gravitoelectric field \(\vec {G}=0\), and a gravitomagnetic field \(\vec {H}=\vec {\Omega }\) arising solely from the rotation (with respect to Fermi–Walker transport) of the spatial triads; thus the only inertial force present is the gravitomagnetic force \(\vec {F}_{\mathrm{GEM}}=\gamma \vec {U}\times \vec {\Omega }\), cf. Eq. (42), with \(\gamma \equiv -U^{\alpha }u_{\alpha }\). In the frame b), there is a gravitoelectric field \(\vec {G}=\vec {\omega }\times (\vec {r}\times \vec {\omega })\) due the observers acceleration, and also a gravitomagnetic field \(\vec {H}=\vec {\omega }\), which originates solely from the vorticity of the observer congruence. That is, there is a gravitomagnetic force \(\gamma \vec {U}\times \vec {\omega }\) which reflects the fact that the relative velocity \(v^{\alpha }=U^{\alpha }/\gamma -u^{\alpha }\) (or \(\vec {v}=\vec {U}/\gamma \), in the observer’s frame, where \(\vec {u}=0\)) between the test particle and the observer it is passing by changes in time. The total inertial forces are in this frame

$$\begin{aligned} \vec {F}_{\mathrm{GEM}}=\gamma \left[ \gamma \vec {\omega }\times (\vec {r}\times \vec {\omega })+\vec {U}\times \vec {\omega }\right] . \end{aligned}$$

In the frame (c), which is the relativistic version of the classical rigid rotating frame, one has the effects of (a) and (b) combined: a gravitoelectric field \(\vec {G}=\vec {\omega }\times (\vec {r}\times \vec {\omega })\), plus a gravitomagnetic field \(\vec {H}=\vec {\omega }+\vec {\Omega }=2\vec {\omega }\), the latter leading to the gravitomagnetic force \(2\gamma \vec {U}\times \vec {\omega }\), which is the relativistic version of the well known Coriolis acceleration, e.g., [112]. The total inertial force is in this frame

$$\begin{aligned} \vec {F}_{\mathrm{GEM}}=\gamma \left[ \gamma \vec {\omega }\times (\vec {r}\times \vec {\omega })+2\vec {U}\times \vec {\omega }\right] \end{aligned}$$

which is the relativistic generalization of the inertial force in e.g., Eq. (4.91) of [112]. Moreover, in this case, as discussed in Sects. 3.2.1 and 3.2.2, the \(\Gamma _{\hat{j}\hat{k}}^{\hat{i}}\) in Eq. (43) are the connection coefficients of the (Levi–Civita) 3-D covariant derivative with respect to the metric \(h_{ij}\) (defined by Eq. (59)) defined on the space manifold associated to the quotient of the spacetime by the congruence; \(\vec {U}\) is the vector tangent to the 3-D curve (see Fig. 2c) obtained by projecting the particle’s worldline on the space manifold, and \( \vec {F}_{\mathrm{GEM}}=\tilde{D}\vec {U}/d\tau \) is simply the covariant 3-D acceleration of that curve.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Costa, L.F.O., Natário, J. Gravito-electromagnetic analogies. Gen Relativ Gravit 46, 1792 (2014). https://doi.org/10.1007/s10714-014-1792-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10714-014-1792-1

Keywords

Navigation