Einstein–Cartan theory

In theoretical physics, the Einstein–Cartan theory, also known as the Einstein–Cartan–Sciama–Kibble theory, is a classical theory of gravitation similar to general relativity but relaxing the assumption that the affine connection has vanishing antisymmetric part (torsion tensor), so that the torsion can be coupled to the intrinsic angular momentum (spin) of matter, much in the same way in which the curvature is coupled to the energy and momentum of matter. In fact, the spin of matter in curved spacetime requires that torsion is not constrained to be zero but is a variable in the principle of stationary action. Regarding the metric and torsion tensors as independent variables gives the correct generalization of the conservation law for the total (orbital plus intrinsic) angular momentum to the presence of the gravitational field. The theory was first proposed by Élie Cartan in 1922[1] and expounded in the following few years.[2] Dennis Sciama[3] and Tom Kibble[4] independently revisited the theory in the 1960s, and an important review was published in 1976.[5] Albert Einstein became affiliated with the theory in 1928 during his unsuccessful attempt to match torsion to the electromagnetic field tensor as part of a unified field theory. This line of thought led him to the related but different theory of teleparallelism.[6]

Einstein–Cartan theory has been historically overshadowed by its torsion-free counterpart and other alternatives like Brans–Dicke theory because torsion seemed to add little predictive benefit at the expense of the tractability of its equations. Since the Einstein–Cartan theory is purely classical, it also does not fully address the issue of quantum gravity. In the Einstein–Cartan theory, the Dirac equation becomes nonlinear[7] and therefore the superposition principle used in usual quantization techniques would not work. Recently, interest in Einstein–Cartan theory has been driven toward cosmological implications, most importantly, the avoidance of a gravitational singularity at the beginning of the universe.[8][9] The theory is considered viable and remains an active topic in the physics community.[10]

Field equations

The Einstein field equations of general relativity can be derived by postulating the Einstein–Hilbert action to be the true action of spacetime and then varying that action with respect to the metric tensor. The field equations of Einstein–Cartan theory come from exactly the same approach, except that a general asymmetric affine connection is assumed rather than the symmetric Levi-Civita connection (i.e., spacetime is assumed to have torsion in addition to curvature), and then the metric and torsion are varied independently.

Let represent the Lagrangian density of matter and represent the Lagrangian density of the gravitational field. The Lagrangian density for the gravitational field in the Einstein–Cartan theory is proportional to the Ricci scalar:

where is the determinant of the metric tensor, and is a physical constant involving the gravitational constant and the speed of light. By Hamilton's principle, the variation of the total action for the gravitational field and matter vanishes:

The variation with respect to the metric tensor yields the Einstein equations:

where is the Ricci tensor and is the canonical stress-energy-momentum tensor. The Ricci tensor is no longer symmetric because the connection contains a nonzero torsion tensor; therefore, the right-hand side of the equation cannot be symmetric either, implying that must include an asymmetric contribution that can be shown to be related to the spin tensor. This canonical energy-momentum tensor is related to the more familiar symmetric energy-momentum tensor by the Belinfante–Rosenfeld procedure.

The variation with respect to the torsion tensor yields the Cartan equations

where is the spin tensor. Because the torsion equation is an algebraic constraint rather than a partial differential equation, the torsion field does not propagate as a wave, and vanishes outside of matter. Therefore, in principle the torsion can be algebraically eliminated from the theory in favor of the spin tensor, which generates an effective "spin-spin" nonlinear self-interaction inside matter.

Avoidance of singularities

The Einstein–Cartan theory eliminates the general-relativistic problem of the unphysical singularity at the Big Bang.[9] The minimal coupling between torsion and Dirac spinors generates an effective nonlinear spin-spin self-interaction, which becomes significant inside fermionic matter at extremely high densities. Such an interaction replaces the singular Big Bang with a cusp-like Big Bounce at a minimum but finite scale factor, before which the observable universe was contracting. This scenario also explains why the present Universe at largest scales appears spatially flat, homogeneous and isotropic, providing a physical alternative to cosmic inflation.[8]

Torsion allows fermions to be spatially extended[11] instead of "pointlike", which helps to avoid the formation of singularities such as black holes and removes the ultraviolet divergence in quantum field theory. According to general relativity, the gravitational collapse of a sufficiently compact mass forms a singular black hole. In the Einstein–Cartan theory, instead, the collapse reaches a bounce and forms a regular Einstein-Rosen bridge (wormhole) to a new, growing universe on the other side of the event horizon.

See also

References

  1. Élie Cartan. "Sur une généralisation de la notion de courbure de Riemann et les espaces à torsion." C. R. Acad. Sci. (Paris) 174, 593–595 (1922).
  2. Élie Cartan. "Sur les variétés à connexion affine et la théorie de la relativité généralisée." Part I: Ann. Éc. Norm. 40, 325–412 (1923) and ibid. 41, 1–25 (1924); Part II: ibid. 42, 17–88 (1925).
  3. Dennis W. Sciama. "The physical structure of general relativity", Rev. Mod. Phys. 36, 463-469 (1964).
  4. Tom W. B. Kibble. "Lorentz invariance and the gravitational field", J. Math. Phys. 2, 212-221 (1961).
  5. Friedrich W. Hehl, Paul von der Heyde, G. David Kerlick, and James M. Nester. "General relativity with spin and torsion: Foundations and prospects." Rev. Mod. Phys. 48, 393–416 (1976). http://link.aps.org/doi/10.1103/RevModPhys.48.393
  6. Hubert F. M. Goenner. "On the History of Unified Field Theories." Living Rev. Relativity, 7, 2 (2004).
  7. F. W. Hehl and B. K. Datta. "Nonlinear spinor equation and asymmetric connection in general relativity", J. Math. Phys. 12, 1334–1339 (1971).
  8. 1 2 Nikodem J. Popławski, (2010). "Cosmology with torsion: An alternative to cosmic inflation". Phys. Lett. B. 694 (3): 181–185. arXiv:1007.0587Freely accessible. Bibcode:2010PhLB..694..181P. doi:10.1016/j.physletb.2010.09.056.
  9. 1 2 Nikodem Popławski, (2012). "Nonsingular, big-bounce cosmology from spinor-torsion coupling". Phys. Rev. D. 85 (10): 107502. arXiv:1111.4595Freely accessible. Bibcode:2012PhRvD..85j7502P. doi:10.1103/PhysRevD.85.107502.
  10. Friedrich W. Hehl. "Note on the torsion tensor." Letter to Physics Today. March 2007, page 16.
  11. Nikodem J. Popławski, (2010). "Nonsingular Dirac particles in spacetime with torsion". Phys. Lett. B. 690 (1): 73–77. arXiv:0910.1181Freely accessible. Bibcode:2010PhLB..690...73P. doi:10.1016/j.physletb.2010.04.073.

Further reading

This article is issued from Wikipedia - version of the 7/14/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.