Skip to main content
Log in

Scalars, vectors and tensors from metric-affine gravity

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

Abstract

The metric-affine gravity provides a useful framework for analyzing gravitational dynamics since it treats metric tensor and affine connection as fundamentally independent variables. In this work, we show that, a metric-affine gravity theory composed of the invariants formed from non-metricity, torsion and curvature tensors can be decomposed into a theory of scalar, vector and tensor fields. These fields are natural candidates for the ones needed by various cosmological and other phenomena. Indeed, we show that the model accommodates TeVeS gravity (relativistic modified gravity theory), vector inflation, and aether-like models. Detailed analyses of these and other phenomena can lead to a standard metric-affine gravity model encoding scalars, vectors and tensors.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Carroll, S.M.: Spacetime and geometry: an introduction to general relativity. Addison-Wesley, San Francisco (2004)

    MATH  Google Scholar 

  2. Palatini, A.: Rend. Circ. Mat. Palermo 43, 203 (1919)

    Article  MATH  Google Scholar 

  3. Einstein, A.: Sitzung-ber Preuss Akad. Wiss. 414 (1925)

  4. Peldan, P.: Actions for gravity, with generalizations: a review class. Quant. Grav. 11, 1087, (1994). [arXiv:gr-qc/9305011]

  5. Magnano, G.: Are there metric theories of gravity other than General Relativity? (1995). [arXiv:gr-qc/9511027]

  6. Vitagliano, V., Sotiriou, T.P., Liberati, S.: Ann. der Phys. 326, 1259–1273 (2011). [arXiv:1008.0171v2[gr-qc]]

    Google Scholar 

  7. Sotiriou, T.P., Liberati, S.: J. Phys. Conf. Ser. 68:012022 (2007). [arXiv:0611040v1[gr-qc]]

    Google Scholar 

  8. Demir, D.A.: Effects of flavor violation on split supersymmetry, (2005). [arXiv:hep-ph/0410056]

  9. Bertone, G., Hooper, D., Silk, J.: Particle dark matter: evidence, candidates and constraints. Phys. Rept. 405, 279 (2005). [arXiv:hep-ph/0404175]

    Google Scholar 

  10. Komatsu, E. et al.: [WMAP Collaboration]: five-year Wilkinson microwave anisotropy probe observations: cosmological interpretation. Astrophys. J. Suppl. 180, 330 (2009). [arXiv:0803.0547 [astro-ph]]

    Google Scholar 

  11. Tegmark, M. et al.: [SDSS Collaboration]: cosmological constraints from the SDSS luminous red galaxies. Phys. Rev. D 74, 123507 (2006). [arXiv:astro-ph/0608632]

  12. Astier, P. et al.: [The SNLS Collaboration]: the supernova legacy survey: measurement of \(Omega_M\), \(Omega_Lambda\) and \(w\) from the first year data set. Astron. Astrophys. 447, 31 (2006). [arXiv:astro-ph/0510447]

  13. Adriani, O. et al.: [PAMELA Collaboration]: an anomalous positron abundance in cosmic rays with energies 1.5.100 GeV. Nature 458, 607 (2009). [arXiv:0810.4995 [astro-ph]]

    Google Scholar 

  14. Abdo, A.A. et al.: [The Fermi LAT Collaboration]: measurement of the cosmic ray \(e+\) plus \(e-\) spectrum from 20 GeV to 1 TeV with the Fermi large area telescope. Phys. Rev. Lett. 102, 181101 (2009). [arXiv:0905.0025 [astro-ph.HE]]

    Google Scholar 

  15. Hooper, D., Blasi, P., Serpico, P.D.: Pulsars as the sources of high energy cosmic ray positrons. JCAP 0901, 025 (2009). [arXiv:0810.1527 [astro-ph]]

  16. Yuksel, H., Kistler, M.D., Stanev, T.: eV Gamma rays from geminga and the origin of the GeV positron excess. Phys. Rev. Lett. 103, 051101 (2009). [arXiv:0810.2784 [astro-ph]]

    Google Scholar 

  17. Milgrom, M.: A modification of the Newtonian dynamics as a possible alternative to the hidden mass hypothesis. Astrophys. J. 270, 365 (1983)

    Article  ADS  Google Scholar 

  18. Bekenstein, J., Milgrom, M.: Does the missing mass problem signal the breakdown of Newtonian gravity? Astrophys. J. 286, 7 (1984)

    Article  ADS  Google Scholar 

  19. Bernal, T., Capozziello, S., Cristofano, G., Laurentis, M.D.: MOND’s acceleration scale as a fundamental quantity. [arXiv:1110.2580v1[gr-qc]]

  20. Sanders, R.H.: A stratified framework for scalar–tensor theories of modified dynamics. Astrophys. J. 480, 492 (1997). [arXiv:astro-ph/9612099]

    Google Scholar 

  21. Sanders, R.H.: A tensor–vector–scalar framework for modified dynamics and cosmic dark matter. Mon. Not. Roy. Astron. Soc. 363, 459 (2005). [arXiv:astro-ph/0502222]

    Google Scholar 

  22. Bekenstein, J.D.: Relativistic gravitation theory for the MOND paradigm. Phys. Rev. D 70, 083509 (2004). [Erratum-ibid. D 71 069901 (2005) ] [arXiv:astro-ph/0403694]

  23. Bekenstein, J.D.: Modified gravity vs dark matter: relativistic theory for mond. PoS JHW2004, 012 (2005). [arXiv:astro-ph/0412652]

  24. Skordis, C.: The tensor–vector–scalar theory and its cosmology. Class. Quant. Grav. 26,143001 (2009) [arXiv:0903.3602 [astro-ph.CO]]

    Google Scholar 

  25. Zlosnik, T.G., Ferreira, P.G., Starkman, G.D.: The vector–tensor nature of Bekenstein’s relativistic theory of modified gravity. Phys. Rev. D 74, 044037 (2006). [arXiv:gr-qc/0606039]

    Google Scholar 

  26. Zlosnik, T.G., Ferreira, P.G., Starkman, G.D.: Modifying gravity with the aether: an alternative to dark matter. Phys. Rev. D 75, 044017 (2007). [arXiv:astro-ph/0607411]

    Google Scholar 

  27. Skordis, C.: Generalizing TeVeS cosmology. Phys. Rev. D 77, 123502 (2008). [arXiv:0801.1985 [astro-ph]]

  28. Giannios, D.: Spherically symmetric, static spacetimes in TeVeS. Phys. Rev. D 71, 103511 (2005). [arXiv:gr-qc/0502122]

  29. Chiu, M.C., Ko, C.M., Tian, Y.: Theoretical aspects of gravitational lensing in TeVeS astrophys. J 636, 565 (2006). [arXiv:astro-ph/0507332]

  30. Skordis, C.: TeVeS cosmology: covariant formalism for the background evolution and linear perturbation theory. Phys. Rev. D 74, 103513 (2006). [arXiv:astro-ph/0511591]

    Google Scholar 

  31. Diaz-Rivera, L.M., Samushia, L., Ratra, B.: Inflation and accelerated expansion TeVeS cosmological solutions. Phys. Rev. D 73, 083503 (2006). [arXiv:astro-ph/0601153]

    Google Scholar 

  32. Chen, D.M., Zhao, H.S.: Strong lensing probability for testing TeVeS theory. Astrophys. J 650, L9 (2006)

    Article  ADS  Google Scholar 

  33. Sagi, E., Bekenstein, J.D.: Black holes in the TeVeS theory of gravity and their thermodynamics. Phys. Rev. D 77, 024010 (2008). [arXiv:0708.2639 [gr-qc]]

  34. Chen, D.M.: Strong lensing probability in TeVeS theory. JCAP 0801, 006 (2008). [arXiv:0712.1633 [astro-ph]]

  35. Contaldi, C.R., Wiseman, T., Withers, B.: TeVeS gets caught on caustics. Phys. Rev. D 78, 044034 (2008). [arXiv:0802.1215 [gr-qc]]

  36. Bekenstein, J.D., Sagi, E.: Do Newton’s \(G\) and Milgrom’s \(a_0\) vary with cosmological epoch? Phys. Rev. D 77, 103512 (2008). [arXiv:0802.1526 [astro-ph]]

    Google Scholar 

  37. Tamaki, T.: Post-Newtonian parameters in the tensor-vector-scalar theory. Phys. Rev. D 77, 124020 (2008). [arXiv:0803.4309 [gr-qc]]

  38. Ferreira, P.G., Skordis, C., Zunckel, C.: Dark matter, modified gravity and the mass of the Neutrino. Phys. Rev. D 78, 044043 (2008). [arXiv:0806.0116 [astro-ph]]

  39. Mavromatos, N.E., Sakellariadou, M., Yusaf, M.F.: Can TeVeS avoid dark matter on galactic scales? Phys. Rev. D 79, 081301 (2009). [arXiv:0901.3932 [astro-ph.GA]]

  40. Ferreras, I., Mavromatos, N.E., Sakellariadou, M., Yusaf, M.F.: Incompatibility of rotation curves with gravitational lensing for TeVeS. Phys. Rev. D 80, 103506 (2009). [arXiv:0907.1463 [astro-ph.GA]]

    Google Scholar 

  41. Lasky, P.D.: Black holes and neutron stars in the generalized tensor–vector–scalar theory. Phys. Rev. D 80, 081501 (2009). [arXiv:0910.0240 [astro-ph.HE]]

  42. Sagi, E.: Propagation of gravitational waves in the generalized tensor–vector–scalar theory. Phys. Rev. D 81, 064031 (2010). [arXiv:1001.1555 [gr-qc]]

  43. Reyes, R., Mandelbaum, R., Seljak, U., Baldauf, T., Gunn, J.E., Lombriser, L., Smith, R.E.: Confirmation of general relativity on large scales from weak lensing and galaxy velocities. Nature 464, 256 (2010). [arXiv:1003.2185 [astro-ph.CO]]

    Google Scholar 

  44. Banados, M., Gomberoff, A., Rodrigues, D.C., Skordis, C.: A note on bigravity and dark matter. Phys. Rev. D 79, 063515 (2009). [arXiv:0811.1270 [gr-qc]]

  45. Banados, M., Ferreira, P.G., Skordis, C.: Eddington–Born–Infeld gravity and the large scale structure of the universe. Phys. Rev. D 79, 063511 (2009). [arXiv:0811.1272 [astro-ph]]

  46. Liddle, A.R., Lyth, D.H.: Cosmological inflation and large scale structure. Cambrigde University Press, Cambrigde (2000)

    Book  Google Scholar 

  47. Tsujikawa, S.: Introductary review of cosmic inflation [arXiv:0304257v1[hep-ph]]

  48. Chiba, T.: Initial conditions for vector inflation [arXiv:0805.4660v3[gr-qc]]

  49. Maki, T.: On the cosmology of Weyl’s gauge invariant gravity. Acta Physica Polonica B 41(6) (2010)

  50. Ford, L.H.: Inflation driven by a vector field. Phys. Rev. D 40, 4 (1989). [arXiv:9402012[gr-qc]]

  51. Golovnev, A., Mukhanov, V., Vanchurin, V.: JCAP 0806, 009 (2008). [arXiv:0802.2068[astro-ph]]

  52. Germani, C., Kehagias, A.: P-nflation: generating cosmic inflation with p-forms [arXiv:0902.3667v2[astro-ph.CO]]

  53. Koh, S.: Vector field and inflation, [arXiv:0902.3904[hep-th]]

  54. Jacobson, T., Mattingly, D.: Gravty with a preferred frame. Phys. Rev. D 64, 024028 (2001). [arXiv:0007031[gr-qc]]

  55. Carroll, S.M., Lim, E.A.: Lorentz violating vector fields slow the Universe down. Phys. Rev. D 70, 123525 (2004). [arXiv:0407149[hep-th]]

    Google Scholar 

Download references

Acknowledgments

C. N. K. thanks Glenn Starkman and Fred Adams for their comments, criticisms and suggestions. A. A. thanks Selin Soysal for discussions. We would like to thank V. Vitagliano and very conscientious referees for their useful comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Canan N. Karahan.

Appendices

Appendix A: Contraction tensors

Contraction of tensors becomes a tedious operation as their rank becomes larger and larger. Already at the rank-3 level, there arise various possibilities in contracting the indices. Indeed, if one defines

$$\begin{aligned} \mathbb A \bullet \mathbb B \equiv \mathbb A ^{\lambda }_{\alpha \beta } \Xi ^{\alpha \beta \mu \nu }_{\lambda \rho } \mathbb B ^{\rho }_{\mu \nu } \end{aligned}$$
(57)

the contraction tensor \(\Xi ^{\alpha \beta \mu \nu }_{\lambda \rho }\) is found to have the most general form

$$\begin{aligned} \Xi ^{\alpha \beta \mu \nu }_{\lambda \rho }&= g_{\lambda \rho }\left(g^{\alpha \beta } g^{\mu \nu } \oplus g^{\alpha \mu } g^{\beta \nu } \oplus g^{\alpha \nu } g^{\beta \mu }\right)\nonumber \\&\oplus \, \delta _{\lambda }^{\alpha } \left( g^{\beta \nu } \delta _{\rho }^{\mu } \oplus g^{\beta \mu } \delta _{\rho }^{\nu } \oplus g^{\mu \nu } \delta _{\rho }^{\beta }\right)\nonumber \\&\oplus \, \delta _{\lambda }^{\beta } \left( g^{\alpha \nu } \delta _{\rho }^{\mu } \oplus g^{\alpha \mu } \delta _{\rho }^{\nu } \oplus g^{\mu \nu } \delta _{\rho }^{\alpha }\right)\nonumber \\&\oplus \, \delta _{\lambda }^{\mu } \left( g^{\beta \nu } \delta _{\rho }^{\alpha } \oplus g^{\alpha \beta } \delta _{\rho }^{\nu } \oplus g^{\alpha \nu } \delta _{\rho }^{\beta } \right)\nonumber \\&\oplus \, \delta _{\lambda }^{\nu } \left( g^{\beta \mu } \delta _{\rho }^{\alpha } \oplus g^{\alpha \beta } \delta _{\rho }^{\mu } \oplus g^{\alpha \mu } \delta _{\rho }^{\beta } \right) \end{aligned}$$
(58)

where \(\oplus \) implies \(+\) or \(-\) depending on whether symmetric or antisymmetric combinations of the indices are involved. Clearly, \(\oplus \) also contains the appropriate symmetry factors.

As an example, let us take \(\mathbb B ^{\rho }_{\mu \nu } = \mathbb S ^{\rho }_{\mu \nu }\) which is antisymmetric in \((\mu ,\nu )\). In this case, when contracting \(\mathbb S ^{\rho }_{\mu \nu }\) with \(\Xi ^{\alpha \beta \mu \nu }_{\lambda \rho }\) only the anti symmetric part of \(\Xi ^{\alpha \beta \mu \nu }_{\lambda \rho }\) in \((\mu ,\nu )\) matters. In other words, when \(\mathbb B ^{\rho }_{\mu \nu } = \mathbb S ^{\rho }_{\mu \nu }\) we consider only

$$\begin{aligned} \Xi ^{\alpha \beta [\mu \nu ]}_{\lambda \rho }&= \frac{1}{2} \Bigg [ g_{\lambda \rho }\left(g^{\alpha \mu } g^{\beta \nu } - g^{\alpha \nu } g^{\beta \mu }\right)\nonumber \\&\oplus \, \delta _{\lambda }^{\alpha } \left( g^{\beta \nu } \delta _{\rho }^{\mu } - g^{\beta \mu } \delta _{\rho }^{\nu } \right)\nonumber \\&\oplus \, \delta _{\lambda }^{\beta } \left( g^{\alpha \nu } \delta _{\rho }^{\mu } - g^{\alpha \mu } \delta _{\rho }^{\nu } \right) \nonumber \\&\oplus \, \Big [\delta _{\lambda }^{\mu } \left( g^{\beta \nu } \delta _{\rho }^{\alpha } \oplus g^{\alpha \beta } \delta _{\rho }^{\nu } \oplus g^{\alpha \nu } \delta _{\rho }^{\beta } \right)\nonumber \\&- \delta _{\lambda }^{\nu } \left( g^{\beta \mu } \delta _{\rho }^{\alpha } \oplus g^{\alpha \beta } \delta _{\rho }^{\mu } \oplus g^{\alpha \mu } \delta _{\rho }^{\beta } \right)\Big ]\Bigg ] \end{aligned}$$
(59)

which is anti-symmetric in \((\mu , \nu )\).

If \(\mathbb A ^{\lambda }_{\alpha \beta }\) in (57) is antisymmetric in \((\alpha ,\beta )\) then we consider antisymmetric part of (59).

$$\begin{aligned} \Xi ^{[\alpha \beta ][\mu \nu ]}_{\lambda \rho }&= \frac{1}{4} \Big [\delta _{\lambda }^{\alpha } \left( g^{\beta \nu } \delta _{\rho }^{\mu } - g^{\beta \mu } \delta _{\rho }^{\nu } \right) - \delta _{\lambda }^{\beta } \left( g^{\alpha \nu } \delta _{\rho }^{\mu } - g^{\alpha \mu } \delta _{\rho }^{\nu } \right) \Big ]\nonumber \\&+ \frac{1}{4}\Big [\delta _{\lambda }^{\mu } \left( g^{\beta \nu } \delta _{\rho }^{\alpha } - g^{\alpha \nu } \delta _{\rho }^{\beta } \right) - \delta _{\lambda }^{\nu } \left( g^{\beta \mu } \delta _{\rho }^{\alpha } - g^{\alpha \mu } \delta _{\rho }^{\beta } \right)\Big ]\nonumber \\&+ \frac{1}{2} g_{\lambda \rho }\left(g^{\alpha \mu } g^{\beta \nu } - g^{\alpha \nu } g^{\beta \mu }\right) \end{aligned}$$
(60)

For instance, \(\mathbb S \bullet \mathbb S \) will be computed by using this contraction tensor.

However, if \(\mathbb A ^{\lambda }_{\alpha \beta }\) in (57) is symmetric in \((\alpha ,\beta )\) then we have to consider symmetric part of (59).

$$\begin{aligned} \Xi ^{(\alpha \beta )[\mu \nu ]}_{\lambda \rho }&= \frac{1}{4} \Big [\delta _{\lambda }^{\alpha } \left( g^{\beta \nu } \delta _{\rho }^{\mu } - g^{\beta \mu } \delta _{\rho }^{\nu } \right) + \delta _{\lambda }^{\beta } \left( g^{\alpha \nu } \delta _{\rho }^{\mu } - g^{\alpha \mu } \delta _{\rho }^{\nu } \right) \Big ]\nonumber \\&+ \frac{1}{4}\Big [\delta _{\lambda }^{\mu } \left( g^{\beta \nu } \delta _{\rho }^{\alpha } + g^{\alpha \nu } \delta _{\rho }^{\beta } \right) - \delta _{\lambda }^{\nu } \left( g^{\beta \mu } \delta _{\rho }^{\alpha } + g^{\alpha \mu } \delta _{\rho }^{\beta } \right)\Big ]\nonumber \\&+ \frac{1}{2} g^{\alpha \beta } \left( \delta _{\lambda }^{\mu } \delta _{\rho }^{\nu } - \delta _{\lambda }^{\nu } \delta _{\rho }^{\mu }\right) \end{aligned}$$
(61)

For instance, \(\mathbb Q \bullet \mathbb S \) should be computed by using this contraction tensor. In computing

\(\mathbb Q \bullet \mathbb Q \) we should symmetrize in both \((\mu ,\nu )\) and \((\alpha ,\beta )\). Then contraction tensor of \(\mathbb Q \bullet \mathbb Q \) is given

$$\begin{aligned} \Xi ^{(\alpha \beta )(\mu \nu )}_{\lambda \rho }&= g^{\alpha \beta }g^{\mu \nu }g_{\lambda \rho }+\frac{1}{2}\Big [g_{\lambda \rho }\left(g^{\alpha \mu } g^{\beta \nu }+g^{\alpha \nu }g^{\beta \mu }\right)\nonumber \\&+g^{\mu \nu }\left(\delta ^{\alpha }_{\lambda } \delta ^{\beta }_{\rho }+\delta ^{\beta }_{\lambda } \delta ^{\alpha } _{\rho }\right)+ g^{\alpha \beta }\left(\delta ^{\mu }_{\lambda } \delta ^{\nu }_{\rho }+\delta ^{\nu }_{\lambda }\delta ^{\mu }_{\rho }\right)\Big ]\nonumber \\&+ \frac{1}{4}\Big [\delta ^{\alpha }_{\lambda }\left(g^{\beta \nu } \delta ^{\mu }_{\rho }+g^{\beta \mu }\delta ^{\nu }_{\rho }\right) +\delta ^{\beta }_{\lambda }\left(g^{\alpha \nu }\delta ^{\mu }_{\rho } +g^{\alpha \mu }\delta ^{\nu }_{\rho }\right)\Big ]\nonumber \\&+\frac{1}{4}\Big [\delta ^{\mu }_{\lambda } \left(g^{\beta \nu }\delta ^{\alpha }_{\rho }+ g^{\alpha \nu }\delta ^{\beta }_{\rho }\right)+\delta ^{\nu }_{\lambda } \left(g^{\beta \mu }\delta ^{\alpha }_{\rho }+g^{\alpha \mu }\delta ^{\beta }_{\rho }\right)\Big ] \end{aligned}$$
(62)

In addition to these, one can compute contraction of divergence of tensors as

$$\begin{aligned} \nabla ^{{(\!\!)}}A\bullet \nabla ^{{(\!\!)}}B= \nabla ^{{(\!\!)}}_{\lambda }A^{\lambda }_{\alpha \beta }\Xi ^ {\alpha \beta \mu \nu }\nabla ^{{(\!\!)}}_{\rho }B^{\rho }_{\mu \nu } \end{aligned}$$
(63)

\(\Xi ^{\alpha \beta \mu \nu }\) is contraction tensor and defined in general form as

$$\begin{aligned} \Xi ^{\alpha \beta \mu \nu }= g^{\alpha \beta }g^{\mu \nu }\oplus g^{\alpha \mu }g^{\beta \nu }\oplus g^{\alpha \nu }g^{\beta \mu } \end{aligned}$$
(64)

If A is symmetric in \((\alpha ,\beta )\) and B is symmetric in \((\mu ,\nu )\) contraction tensor takes the form

$$\begin{aligned} \Xi ^{(\alpha \beta )(\mu \nu )}= g^{\alpha \beta }g^{\mu \nu }+\frac{1}{2}\left(g^{\alpha \mu }g^{\beta \nu }+g^{\alpha \nu }g^{\beta \mu }\right) \end{aligned}$$
(65)

this contraction tensor can be used to compute \(\nabla \mathbb Q \bullet \nabla \mathbb Q \) because \(\mathbb Q \) is symmetric in \((\alpha \beta )\). To compute \(\nabla \mathbb S \bullet \nabla \mathbb S \), one needs contraction tensor which is antisymmetric both couple of indices.So,

$$\begin{aligned} \Xi ^{[\alpha \beta ][\mu \nu ]}= \frac{1}{2}\left( g^{\alpha \mu }g^{\beta \nu }-g^{\alpha \nu }g^{\beta \mu }\right) \end{aligned}$$
(66)

If one writes contraction tensor of \(\nabla \mathbb Q \bullet \nabla \mathbb S \), finds it as

$$\begin{aligned} \Xi ^{(\alpha \beta )[\mu \nu ]}=0 \end{aligned}$$
(67)

Appendix B: Positive-definite mass matrix

In the text, we mentioned that for a stable theory, each of the three eigenvalues must individually be positive. This leads to non-trivial constraints on the coefficients in (42). In this appendix we shall discuss certain related details. The eigenvalues of (41) follow from the cubic algebraic equation

$$\begin{aligned} -\!\lambda ^{3}+b\lambda ^{2}+c\lambda +d=0 \end{aligned}$$
(68)

where

$$\begin{aligned} b&= a_{VV}+a_{UU}+a_{WW}\nonumber \\ c&= -a_{VV}a_{UU}-a_{VV}a_{WW}-a_{UU}a_{WW}+a^{2}_{UW} +a^{2}_{VU}+a^{2}_{VW}\nonumber \\ d&= a_{VV}a_{UU}a_{WW}+2a_{VU}a_{UW}a_{VW}-a^{2}_{UW} a_{VV}-a_{VW}^{2}a_{UU}-a^{2}_{VU}a_{WW}.\qquad \end{aligned}$$
(69)

The roots of (68) must each be non-negative for guaranteeing absence of instabilities. The analytic expressions for roots are well-known. However, the constraint equations they lead to are too complicated to achieve specific statements about the elements of the mass matrix (41). Nevertheless, in a given specific problem, one can determine the allowed ranges for \(a_{VV}, \ldots , a_{UW}\) at least numerically,

As an algebraically simpler case to exemplify, one can focus on the special case of vanishing discriminant, that is, one considers

$$\begin{aligned} \Delta =18 abcd - 4 b^{3}d + b^{2}c^{2}- 4 ac^{3}-27 a^{2} d^{2} \end{aligned}$$
(70)

so that only two independent eigenvalues are left. Indeed, one has

$$\begin{aligned} \lambda _{1}=-\frac{b}{3a}-\frac{2}{3a} \root 3 \of {\frac{1}{2}[2b^{3}-9abc+27a^{2}d]} \end{aligned}$$
(71)

and

$$\begin{aligned} \lambda _{2}=-\frac{b}{3a}+\frac{1}{3a}\root 3 \of {\frac{1}{2}[2b^{3}-9abc+27a^{2}d]}\,. \end{aligned}$$
(72)

For positive-definite mass matrix, \(\lambda _{1}\) and \(\lambda _{2}\) must each be positive:

$$\begin{aligned} \lambda _{1}> 0 \Rightarrow -\frac{b}{2} < \root 3 \of {\frac{1}{2}[2b^{3}-9abc+27a^{2}d]} \end{aligned}$$
(73)

and

$$\begin{aligned} \lambda _{2}> 0 \Rightarrow b > \root 3 \of {\frac{1}{2}[2b^{3}-9abc+27a^{2}d]}\,. \end{aligned}$$
(74)

These two constraints lead one at once to the bound

$$\begin{aligned} -\frac{b}{2} < \root 3 \of {\frac{1}{2}[2b^{3}-9abc+27a^{2}d]} < b. \end{aligned}$$
(75)

Similar bounds can be derived for general as well as special cases [8]. In general, constraints on various coefficients become more suggestive in some physically relevant special cases. We here thus exemplify two such cases: symmetric and antisymmetric connections.

  1. 1.

    Symmetric Connection : \({(\!\!)}^{\lambda }_{\alpha \beta }={(\!\!)}^{\lambda }_{ \beta \alpha }\) As we have mentioned in the text, in this case, torsion tensor identically vanishes (\(\mathbb S ^{\lambda }_{\alpha \beta }=0\)), and consequently \(\mathtt V _{\alpha }=\mathtt U _{\alpha }\). The theory then reduces to a two-vector theory of \(\mathtt V \) and \(\mathtt W \). From Eq. (40) the mass-squared matrix of vectors is found to be

    $$\begin{aligned} \frac{1}{2} M_{Pl}^2 \left(\begin{array}{cc} a^{\prime }_{VV}+a^{\prime }_{UU}+2 a^{\prime }_{VU} \quad a^{\prime }_{VW}+a^{\prime }_{UW}\\ a^{\prime }_{VW}+a^{\prime }_{UW} \quad a^{\prime }_{WW} \end{array} \right) \end{aligned}$$
    (76)

    where various coefficients are given by

    $$\begin{aligned} a^{\prime }_{VV}&= \frac{1}{18} + \frac{44}{9} c_{Q} ,\nonumber \\ a^{\prime }_{UU}&= a_{WW} \,,\nonumber \\ a^{\prime }_{WW}&= \frac{1}{18} + \frac{14}{9} c_{Q},\nonumber \\ a^{\prime }_{VU}&= -\frac{1}{9} + \frac{20}{9} c_{Q} ,\nonumber \\ a^{\prime }_{VW}&= -\frac{1}{9} + \frac{20}{9} c_{Q} ,\nonumber \\ a^{\prime }_{UW}&= \frac{7}{18} + \frac{14}{9} c_{Q} . \end{aligned}$$
    (77)

    which follow from (40) for vanishing torsion. Clearly, \(c_Q\) is the only variable. The eigenvalues of (76) follow from the quadratic algebraic equation;

    $$\begin{aligned} \lambda ^2+b^{\prime }\lambda +c^{\prime }=0 \end{aligned}$$
    (78)

    where

    $$\begin{aligned} b^{\prime }&= -a_{VV}^{\prime }-a_{UU}^{\prime }-2a_{UU}^{\prime }-a_{WW}^{\prime }\nonumber \\ c^{\prime }&= (a_{VV}^{\prime }+a_{UU}^{\prime }+2a_{UU}^{\prime })a_{WW}^{\prime }- (a_{VW}^{\prime }+a_{UW}^{\prime })^2\,. \end{aligned}$$
    (79)

    From the Eq. (78), one directly determines the discriminant

    $$\begin{aligned} \Delta = \frac{11680}{81}c_{Q}^2+\frac{872}{162}c_{Q}+\frac{109}{324} \end{aligned}$$
    (80)

    and eigenvalues

    $$\begin{aligned} \lambda _{1,2}=\frac{a_{VV}^{\prime }+a_{UU}^{\prime }+2a_{UU}^{\prime }+a_{WW}^{\prime }\pm \sqrt{\Delta }}{2}= \frac{-\frac{1}{18}+\frac{112}{9}c_{Q}\pm \sqrt{\Delta }}{2} \end{aligned}$$
    (81)

    For a physically sensible theory, the eigenvalues must all be positive. By considering the constraint of positive discriminant and roots, one finds two appropriate intervals

    $$\begin{aligned} c_{Q} < - 0.046 \quad c_{Q}> 0.68. \end{aligned}$$
    (82)

    This shows that except for the small interval containing origin, all values of \(c_{Q}\) lead to a stable massive two-vector theory.

  2. 2.

    Anti-symmetric tensorial connection: \(V_{\alpha }=-U_{\alpha }\) and \(W_{\alpha }=0\) In this case we end up with a single-vector theory with mass-squared \(\frac{1}{2}M_{Pl}^2\bar{a}_{VV}\) where \(\bar{a}_{VV}=1/3+8c_{S}+2c_{Q}+8c_{QS}\). This coefficient must be positive and hence

    $$\begin{aligned} 4c_{S}+c_{Q}+4c_{QS}> -\frac{1}{6} \end{aligned}$$
    (83)

    A much more special arises when non-metricity vanishes. In this special case, the coefficients \(c_{Q}\) and \(c_{QS}\) both vanish, an one finds

    $$\begin{aligned} c_{S}>-\frac{1}{24} \end{aligned}$$
    (84)

    as a bound on \(c_S\).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Karahan, C.N., Altaş, A. & Demir, D.A. Scalars, vectors and tensors from metric-affine gravity. Gen Relativ Gravit 45, 319–343 (2013). https://doi.org/10.1007/s10714-012-1473-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10714-012-1473-x

Keywords

Navigation