Skip to main content
Log in

Conserved charges and first law of thermodynamics for Kerr–de Sitter black holes

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

Abstract

Recently, a general method for calculating conserved charges for (black hole) solutions to generally covariant gravitational theories, in any dimensions and with arbitrary asymptotic behaviors has been introduced. Equipped with this method, which can be dubbed as “solution phase space method,” we calculate mass and angular momentum for the Kerr–dS black holes. Furthermore, for any choice of horizons, associated entropy and the first law of thermodynamics are derived. Interestingly, according to insensitivity of the analysis to the chosen cosmological constant, the analysis unifies the thermodynamics of rotating stationary black holes in 4 (and other) dimensions with either AdS, flat or dS asymptotics. We extend the analysis to include electric charge, i.e. to the Kerr–Newman–dS black holes.

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

Notes

  1. The choice of the alphabet H originates from the words “Hamiltonian generator” which we use interchangeably with the “conserved charge”.

References

  1. Hawking, S.W.: Particle creation by black holes. Commun. Math. Phys. 43, 199–220 (1975)

    Article  ADS  MathSciNet  Google Scholar 

  2. Bekenstein, J.D.: Black holes and entropy. Phys. Rev. D 7, 2333–2346 (1973)

    Article  ADS  MathSciNet  Google Scholar 

  3. Bardeen, J.M., Carter, B., Hawking, S.W.: The four laws of black hole mechanics. Commun. Math. Phys. 31, 161–170 (1973)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Zurek, W.H.: Entropy evaporated by a black hole. Phys. Rev. Lett. 49, 1683 (1982)

    Article  ADS  Google Scholar 

  5. Aghapour, S., Hajian, K.: Black hole entropy from entropy of Hawking radiation. arXiv:1603.02866

  6. Strominger, A., Vafa, C.: Microscopic origin of the Bekenstein–Hawking entropy. Phys. Lett. B 379, 99–104 (1996). arXiv:hep-th/9601029

    Article  ADS  MathSciNet  Google Scholar 

  7. Rovelli, C.: Black hole entropy from loop quantum gravity. Phys. Rev. Lett. 77, 3288–3291 (1996). http://arxiv.org/abs/gr-qc/9603063

  8. Ashtekar, A., Baez, J., Corichi ,A., Krasnov, K.: Quantum geometry and black hole entropy. Phys. Rev. Lett. 80, 904–907 (1998). http://arxiv.org/abs/gr-qc/9710007

  9. Meissner, K.A.: Black hole entropy in loop quantum gravity. 21, 5245–5252 (2004). http://arxiv.org/abs/gr-qc/0407052

  10. Wald, R.M.: Black hole entropy is the Noether charge. Phys. Rev. D 48, 3427–3431 (1993). arXiv:gr-qc/9307038

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Iyer, V., Wald, R.M.: Some properties of Noether charge and a proposal for dynamical black hole entropy. Phys. Rev. D 50, 846–864 (1994). arXiv:gr-qc/9403028

    Article  ADS  MathSciNet  Google Scholar 

  12. Carter, B.: The commutation property of a stationary, axisymmetric system. Commun. Math. Phys. 17, 233–238 (1970)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Carter, B.: DeWitt, C., DeWitt, B.S. (eds.) Les Astre Occlus, Proceedings of 1972 Les Houches Summer School, 2nd edn. Gordon and Breach, New York (1973)

  14. Spradlin, M., Strominger, A., Volovich, A.: Les Houches lectures on de Sitter Space. Em Unity from duality: gravity, gauge theory and strings. Proceedings, NATO Advanced Study Institute, Euro Summer School, 76th session, Les Houches, France, July 30–August 31, 2001, páginas, pp. 423–453, (2001). arXiv:hep-th/0110007

  15. Akcay, S., Matzner, R.A.: Kerr–de sitter universe. Class. Quantum Gravity 28, 085012 (2011). arXiv:1011.0479

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Gibbons, G.W., Hawking, S.W.: Cosmological event horizons, thermodynamics, and particle creation. Phys. Rev. D 15, 2738–2751 (1977)

    Article  ADS  MathSciNet  Google Scholar 

  17. Abbott, L.F., Deser, S.: Stability of gravity with a cosmological constant. Nucl. Phys. B 195, 76 (1982)

    Article  ADS  MATH  Google Scholar 

  18. Perlmutter, S., et al.: Measurements of Omega and Lambda from 42 high redshift supernovae. Astrophys. J. 517, 565–586 (1999). arXiv:astro-ph/9812133

    Article  ADS  Google Scholar 

  19. Strominger, A.: The dS/CFT correspondence. JHEP 10, 034 (2001). arXiv:hep-th/0106113

    Article  ADS  MathSciNet  Google Scholar 

  20. Klemm, D.: Some aspects of the de Sitter/CFT correspondence. Nucl. Phys. B 625, 295 (2002). arXiv:hep-th/0106247

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Balasubramanian, V., de Boer, J., Minic, D.: Mass, entropy and holography in asymptotically de Sitter spaces. Phys. Rev. D 65, 123508 (2002). arXiv:hep-th/0110108

    Article  ADS  MathSciNet  Google Scholar 

  22. Ghezelbash, A.M., Mann, R.B.: Action, mass and entropy of Schwarzschild-de Sitter black holes and the de Sitter/CFT correspondence. JHEP 01, 005 (2002). arXiv:hep-th/0111217

    Article  ADS  MathSciNet  Google Scholar 

  23. Dehghani, M.H.: Kerr–de Sitter space-times in various dimension and dS/CFT correspondence. Phys. Rev. D 65, 104003 (2002). arXiv:hep-th/0112002

    Article  ADS  MathSciNet  Google Scholar 

  24. Dehghani, M.H.: Quasilocal thermodynamics of Kerr–de Sitter space-times and the AdS/CFT correspondence. Phys. Rev. D 65, 104030 (2002). arXiv:hep-th/0201128

    Article  ADS  MathSciNet  Google Scholar 

  25. Dehghani, M.H., KhajehAzad, H.: Thermodynamics of Kerr–Newman de Sitter black hole and dS/CFT correspondence. Can. J. Phys. 81, 1363 (2003). arXiv:hep-th/0209203

    Article  ADS  Google Scholar 

  26. Cai, R.G.: Cardy–Verlinde formula and thermodynamics of black holes in de Sitter spaces. Nucl. Phys. B 628, 375 (2002). arXiv:hep-th/0112253

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Teitelboim, C.: Gravitational thermodynamics of Schwarzschild–de Sitter space. Em Meeting on Strings and Gravity: Tying the Forces Together Brussels, Belgium, October 19–21, 2001 (2002). arXiv:hep-th/0203258

  28. Gomberoff, A., Teitelboim, C.: de Sitter black holes with either of the two horizons as a boundary. Phys. Rev. D 67, 104024 (2003). arXiv:hep-th/0302204

    Article  ADS  MathSciNet  Google Scholar 

  29. Deser, S., Tekin, B.: Energy in generic higher curvature gravity theories. Phys. Rev. D 67, 084009 (2003). arXiv:hep-th/0212292

    Article  ADS  MathSciNet  Google Scholar 

  30. Deser, S., Kanik, I., Tekin, B.: Conserved charges of higher D Kerr-AdS spacetimes. Class. Quantum Gravity 22, 3383 (2005). arXiv:gr-qc/0506057

  31. Deser, S., Tekin, B.: New energy definition for higher curvature gravities. Phys. Rev. D 75, 084032 (2007). arXiv:gr-qc/0701140

    Article  ADS  MathSciNet  Google Scholar 

  32. Ghezelbash, A.M., Mann, R.B.: Entropy and mass bounds of Kerr–de Sitter spacetimes. Phys. Rev. D 72, 064024 (2005). arXiv:hep-th/0412300

    Article  ADS  MathSciNet  Google Scholar 

  33. Roy Choudhury, T., Padmanabhan, T.: Concept of temperature in multi-horizon spacetimes: analysis of Schwarzschild–de Sitter metric. Gen. Relativ. Gravit. 39, 1789–1811 (2007). arXiv:gr-qc/0404091

    Article  ADS  MathSciNet  MATH  Google Scholar 

  34. Sekiwa, Y.: Thermodynamics of de Sitter black holes: thermal cosmological constant. Phys. Rev. D 73, 084009 (2006). arXiv:hep-th/0602269

    Article  ADS  MathSciNet  Google Scholar 

  35. Chruciel, P.T., Jezierski, J., Kijowski, J.: Hamiltonian mass of asymptotically Schwarzschild–de Sitter space-times. Phys. Rev. D 87, 12124015 (2013). arXiv:1305.1014

    ADS  Google Scholar 

  36. Chruciel, P.T., Jezierski, J., Kijowski, J.: Hamiltonian dynamics in the space of asymptotically Kerr–de Sitter spacetimes. Phys. Rev. D 92, 084030 (2015). arXiv:1507.03868

    Article  ADS  MathSciNet  Google Scholar 

  37. Dolan, B.P., Kastor, D., Kubiznak, D., Mann, R.B., Traschen, J.: Thermodynamic volumes and isoperimetric inequalities for de Sitter black holes. Phys. Rev. D 87(10), 104017 (2013). arXiv:1301.5926

    Article  ADS  Google Scholar 

  38. McInerney, J., Satishchandran, G., Traschen, J.: Cosmography of KNdS black holes and isentropic phase transitions. arXiv:1509.02343

  39. Kubiznak, D., Simovic, F.: Thermodynamics of horizons: de Sitter black holes. arXiv:1507.08630 [hep-th]

  40. Hajian, K., Sheikh-Jabbari, M.M.: Solution phase space and conserved charges: a general formulation for charges associated with exact symmetries. Phys. Rev. D 93, 4,044074 (2016). arXiv:1512.05584

    Article  Google Scholar 

  41. Hajian, K.: On Thermodynamics and Phase Space of Near Horizon Extremal Geometries (2015). arXiv:1508.03494

  42. Lee, J., Wald, R.M.: Local symmetries and constraints. J. Math. Phys. Springer, 31, 725–743 (1990)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Wald, R.M., Zoupas, A.: A general definition of ’conserved quantities’ in general relativity and other theories of gravity. Phys. Rev. D 61, 084027 (2000). arXiv:gr-qc/9911095

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Barnich, G., Brandt, F.: Covariant theory of asymptotic symmetries, conservation laws and central charges. Nucl. Phys. B 633, 3–82 (2002). arXiv:hep-th/0111246

    Article  ADS  MathSciNet  MATH  Google Scholar 

  45. Hajian, K., Seraj, A., Sheikh-Jabbari, M.M.: Near horizon extremal geometry perturbations: dynamical field perturbations vs. parametric variations. JHEP 1410, 111 (2014). arXiv:1407.1992

    Article  ADS  MathSciNet  MATH  Google Scholar 

  46. Kerr, R.P.: Gravitational field of a spinning mass as an example of algebraically special metrics. Phys. Rev. Lett. 11, 237–238 (1963)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  47. Griffiths, J.B., Podolskỳ, J.: Exact Space-Times in Einstein’s General Relativity. Cambridge University Press, Cambridge (2009)

    Book  MATH  Google Scholar 

  48. Booth, I.S., Mann, R.B.: Cosmological pair production of charged and rotating black holes. Nucl. Phys. B 539, 267–306 (1999). arXiv:gr-qc/9806056

    Article  ADS  Google Scholar 

  49. Crnkovic, C., Witten, E.: Covariant description of canonical formalism in geometrical theories. In: Hawking, S.W., Israel, W. (eds.) Three Hundred Years of Gravitation, pp. 676–684 (1987)

  50. Ashtekar, A., Bombelli, L., Koul, R.: Phase space formulation of general relativity without a 3+1 splitting. Lect. Notes Phys. 278, 356–359 (1987)

    Article  ADS  Google Scholar 

  51. Ashtekar, A., Bombelli, L. Reula, O.: The covariant phase space of asymptotically flat gravitational fields. In: Francaviglia, M. (ed.) Mechanics, Analysis and Geometry: 200 Years after Lagrange, pp 417–450. North-Holland (1991). doi:10.1016/B978-0-444-88958-4.50021-5

  52. Compre, G., Mao, P.-J., Seraj, A., Sheikh-Jabbari, M.M.: Symplectic and Killing symmetries of AdS\(_{3}\) gravity: holographic vs boundary gravitons. JHEP 01, 080 (2016). arXiv:1511.06079

    Article  ADS  MathSciNet  Google Scholar 

  53. Compre, G., Hajian, K., Seraj, A., Sheikh-Jabbari, M.M.: Extremal rotating black holes in the near-horizon limit: phase space and symmetry algebra. Phys. Lett. B 749, 443–447 (2015). arXiv:1503.07861

    Article  ADS  Google Scholar 

  54. Compre, G., Hajian, K., Seraj, A., Sheikh-Jabbari, M.M.: Wiggling throat of extremal black holes. JHEP 10, 093 (2015). arXiv:1506.07181

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgments

I would like to thank Shahin Sheikh-Jabbari, for his crucial contributions to this paper, in addition to helpful discussions on the subject. I would also like to thank Erfan Esmaeili, who motivated me to study Kerr–dS BHs using SPSM. Besides, I thank Ali Seraj for all of the things he has taught me about covariant phase space formulation. This work has been supported by the Allameh Tabatabaii Prize Grant of National Elites Foundation of Iran and the Saramadan grant of the Iranian vice presidency in science and technology.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kamal Hajian.

Appendix: A deeper review on solution phase space method

Appendix: A deeper review on solution phase space method

The goal of this appendix is to provide a conceptual review on SPSM, although reference to the original paper [40] is recommended. Before reviewing SPSM, we need to recap a standard phase space construction, dubbed as covariant phase space formulation [10, 11, 42, 4951].

Covariant phase space formulation: Phase space \(\mathcal {F}(\mathcal {M},\Omega )\) is a manifold \(\mathcal {M}\) equipped with a closed nondegenerate symplectic form \(\Omega \). In classical field mechanics, it is usual to build the phase space canonically, i.e. building the \(\mathcal {M}\) from a subset of field configurations \(\Phi (\mathbf {x})\) and their momentum conjugates defined on some privileged time foliation of spacetime. In this construction, solutions to the equation of motion are some curves on \(\mathcal {M}\) parametrized by the time. Interestingly, in the context of generally covariant gravitational theories, there is a more suitable construction which does not break general covariance by specifying a time foliation. In this construction, \(\mathcal {M}\) is composed of dynamical field configurations all over the spacetime \(\Phi (x^\mu )\). On the other hand, there would not be any field conjugate present. As a result, any solution to the equation of motion in the phase space would be a point on \(\mathcal {M}\), instead of a curve. The tangent space of the manifold is also constituted from a subset of perturbations \(\delta \Phi (x^\mu )\). The symplectic 2-form which makes \(\mathcal {M}\) to be a phase space is constructed from the Lagrangian d-form \(\mathbf {L}\). To this end, picking up the Lee-Wald \((d-1)\)-form \(\varvec{\Theta }\) from the variation of Lagrangian

$$\begin{aligned} \delta \mathbf {L}=\mathbf {E}_{{\Phi }}\delta \Phi +\mathrm {d}\varvec{\Theta }_{\text {LW}}(\delta \Phi ,\Phi ), \end{aligned}$$
(6.1)

the symplectic form would be [10, 11, 42]

$$\begin{aligned} \Omega _{\text {LW}}(\delta _1\Phi ,\delta _2\Phi ,\Phi )\equiv \int _\Sigma \varvec{\omega }_{\text {LW}}(\delta _1\Phi ,\delta _2\Phi ,\Phi )\, \end{aligned}$$
(6.2)

where

$$\begin{aligned} \varvec{\omega }_{\text {LW}}(\delta _1\Phi ,\delta _2\Phi ,\Phi )=\delta _1\varvec{\Theta }_{\text {LW}}(\delta _2\Phi ,\Phi )-\delta _2\varvec{\Theta }_{\text {LW}}(\delta _1\Phi ,\Phi ). \end{aligned}$$
(6.3)

The \(\mathbf {E}_{{\Phi }}\) denotes equation of motion for the field \(\Phi \), the \(\Sigma \) is some codimension-1 (Cauchy) surface and \(\delta _{1,2}\Phi \) are some members of the tangent space. The \(\varvec{\omega }_{\text {LW}}\) is called (pre)symplectic current. Closed-ness of \(\Omega \) is guaranteed by the definition (6.3). In order to make \(\Omega _{\text {LW}}\) independent of the choice of \(\Sigma \), one needs \(\mathrm {d}\varvec{\omega }_{\text {LW}}=0\) and flow of \(\varvec{\omega }_{\text {LW}}\) out of the boundaries \(\partial \Sigma \) vanish. The former is achieved if \(\Phi \) and \(\delta \Phi \) satisfy e.o.m and linearized e.o.m respectively. So, it is standard to request them from the beginning. But achievement of the latter needs extra conditions, usually some boundary conditions on perturbations. An important thing to be mentioned in covariant phase space formulation is the ambiguity of addition an exact \((d-1)\)-form \(\mathrm {d}\mathbf {Y}(\delta \Phi ,\Phi )\) to the \(\varvec{\Theta }_{\text {LW}}(\delta \Phi ,\Phi )\), i.e.

$$\begin{aligned} \varvec{\Theta }_{\text {LW}}(\delta \Phi ,\Phi )\rightarrow \varvec{\Theta }(\delta \Phi ,\Phi )=\varvec{\Theta }_{\text {LW}}(\delta \Phi ,\Phi )+\mathrm {d}\mathbf {Y}(\delta \Phi ,\Phi ) \end{aligned}$$
(6.4)

This ambiguity entails corresponding ambiguities in the \(\Omega \) defined above, through

$$\begin{aligned} \varvec{\omega }(\delta _1\Phi ,\delta _2\Phi ,\Phi )\rightarrow \varvec{\omega }(\delta _1\Phi ,\delta _2\Phi ,\Phi )+ \mathrm {d}\big (\delta _2 \mathbf {Y}(\delta _1 \Phi ,\Phi )-\delta _1 \mathbf {Y}(\delta _2 \Phi ,\Phi )\big ). \end{aligned}$$
(6.5)

Using the symplectic form, one can associate a Hamiltonian generator (interchangeably called conserved charge) to a diffeomorphism+gauge transformation \(\epsilon =\{\xi ,\lambda ^a\}\) as

$$\begin{aligned} \delta H_{\epsilon }(\Phi )&\!\equiv \! \int _\Sigma \big (\delta ^{[\Phi ]}\varvec{\Theta }(\delta _\epsilon \Phi ,\Phi )-\delta _\epsilon \varvec{\Theta }(\delta \Phi ,\Phi )\big )\!=\!\int _{\Sigma }\mathrm {d}\varvec{k}_{\epsilon }(\delta \Phi ,\Phi )\!=\!\oint _{\partial \Sigma }\varvec{k}_{\epsilon }(\delta \Phi ,\Phi ). \end{aligned}$$
(6.6)

The \(\delta ^{[\Phi ]}\) emphasizes that \(\delta \) acts on dynamical fields, not the \(\epsilon \). Moreover, \(\delta _\epsilon \Phi \equiv \mathcal {L}_\xi \Phi +\delta _{\lambda ^a}A^a\) where \(A^a\) are some probable Abelian gauge fields. In the equation above, the integrand in the first integration has been replaced by an exact \((d-1)\)-form \(\mathrm {d}\varvec{k}_\epsilon \). So, the last equation follows from the Stokes theorem. The \((d-2)\)-form \(\varvec{k}_\epsilon \) is explicitly as (see Appendix A in Ref. [40] for detailed derivation)

$$\begin{aligned} \varvec{k}_\epsilon (\delta \Phi ,\Phi )=\delta \mathbf {Q}_\epsilon -\xi \cdot \varvec{\Theta }(\delta \Phi ,\Phi ), \end{aligned}$$
(6.7)

in which \(\mathbf {Q}_\epsilon \) is the Noether–Wald charge density, defined by the relation

$$\begin{aligned} \mathrm {d}\mathbf {Q}_\epsilon \equiv \varvec{\Theta }(\delta _\epsilon \Phi ,\Phi )-\xi \cdot \mathbf {L}. \end{aligned}$$
(6.8)

Hence, by the Eq. (6.7), \(\varvec{k}_\epsilon \) can be found for different theories straightforwardly. Putting it into Eq. (6.6), if the last integral would be finite and nonvanishing, \(\delta H_{\epsilon }(\Phi )\) then corresponds to a conserved charge variation. In order to find the finite conserved charge \(H_{\epsilon }\), integrability over the phase space is needed. This condition is basically \((\delta _1\delta _2-\delta _2\delta _1)H_\epsilon (\Phi )=0\), in which \(\Phi \)s are any field configuration in the presumed phase space \(\mathcal {F}\), and \(\delta _{1,2}\Phi \) are any arbitrary chosen member of its tangent space. Then, it follows that the integrability condition can be explained as [42, 43, 52]

$$\begin{aligned} \oint _{\partial \Sigma } \Big (\xi \cdot \varvec{\omega }(\delta _1\Phi ,\delta _2\Phi ,\Phi )+\varvec{k}_{\delta _1\epsilon }(\delta _2\Phi ,\Phi ) -\varvec{k}_{\delta _2\epsilon }(\delta _1\Phi ,\Phi )\Big )=0. \end{aligned}$$
(6.9)

As far as calculation of conserved charges are concerned, conservation of \(\delta H_\epsilon \) can be guaranteed if \(\epsilon \) is chosen such that \(\varvec{\omega }(\delta \Phi ,\delta _\epsilon \Phi ,\Phi )=0\) on-shell. It is because there would not be any flow out of the boundaries locally, and hence globally. The family of \(\epsilon \)’s with this property, which has been dubbed “symplectic symmetry generators” [53], can be divided to two sets: (1) the ones for which \(\delta _\epsilon \Phi \ne 0\) at least on one of the points of the phase space, (2) the ones for which \(\delta _\epsilon \Phi =0\) all over the phase space. The former set, dubbed as “nonexact symmetry generators”, constitute a closed algebraic structure, and are considered to be responsible for generating the phase space of a solution at given constant thermodynamical variables. We can dub the generated phase space as “statistical phase space”. Hence, they open a road towards understanding microstates of the system (see [5254] for works in this direction). The latter set are dubbed “exact symmetry generators” and are considered as generators of the set of solutions in different thermodynamical variables [40]. The generated phase space has been called “solution phase space” which we describe below. It has been conjectured that the phase space associated with the geometries without propagating degrees of freedom are composed of the combination of statistical and solution phase spaces [40].

Solution phase space method: This method is specification of the covariant phase space formulation to some specific manifolds and their tangent spaces which endows that method the power of calculability. Consider a family of (black hole) solutions to a generally covariant gravitational theory. Usually, such a family is identified by some isometries and some parameters \(p_j\). The parameters are some arbitrary (but with constrained domain) real numbers appearing in the field configuration of the mentioned solutions. The parameters can be reparametrized, but can not be removed by coordinate transformations. The manifold \(\hat{\mathcal {M}}\) can be chosen to be composed of the members of the family, up to unphysical coordinate/gauge transformations. The symplectic 2-form \(\hat{\Omega }\) would be simply the Lee-Wald symplectic form confined to \(\hat{\mathcal {M}}\). Then, the \(\mathcal {F}_p=(\hat{\mathcal {M}},\hat{\Omega })\) would be a phase space, the “solution phase space”. Hence, any point of the manifold can be identified by \(\hat{\Phi }(x^\mu ,p_j)\). Tangent space of the \(\hat{\mathcal {M}}\) is spanned (up to infinitesimal pure gauge transformations) by “parametric variations” which are found simply by [45]

$$\begin{aligned} \hat{\delta }\Phi \equiv \frac{\partial \hat{\Phi }}{\partial p_j}\delta p_j. \end{aligned}$$
(6.10)

These variations, which are infinitesimal difference of two solutions, satisfy linearized equation of motion. Hence, they respect \(\mathrm {d}\varvec{\omega }_{\text {LW}}(\hat{\delta }_1\Phi ,\hat{\delta }_2\Phi ,\hat{\Phi })=0\).

As it was advertised above, conservation of \(\hat{\delta } H_\epsilon \) is guaranteed if \(\epsilon \) is chosen to be an exact symmetry generators \(\eta \) defined in Eq. (2.1). This results is because of \(\varvec{\omega }_{\text {LW}}(\hat{\delta }\Phi ,\delta _\eta \hat{\Phi },\hat{\Phi })=0\), (which itself is a result of linearity of \(\varvec{\omega }_{\text {LW}}\) in \(\delta _\eta \hat{\Phi }=0\)), preventing flow of \(\varvec{\omega }_{\text {LW}}\) out of the boundaries \(\partial \Sigma \). Along with guaranteeing the conservation, the relation \(\varvec{\omega }_{\text {LW}}(\hat{\delta }\Phi ,\delta _\eta \hat{\Phi },\hat{\Phi })=0\) yields an additional interesting and unexpected result: \(\hat{\delta } H_\eta \) would also be independent of the chosen \(\partial \Sigma \). It is because of vanishing of \(\varvec{\omega }_{\text {LW}}\) all over the \(\Sigma \), and hence, vanishing of \(\varvec{\omega }_{\text {LW}}\) in the region enclosed between two different integrating surfaces \(\partial \Sigma _1\) and \(\partial \Sigma _2\). Then, by the Stokes theorem, and noticing the Eq. (6.6), the claim is proved. Explaining this result in another way, although the integration in calculating \(\hat{\delta } H_\eta \) is over codimension-2 surface \(\partial \Sigma \), but the result would be independent of all coordinates, including the two coordinates which are not integrated on.

Focusing on exact symmetries results in another nice feature for calculation of their conserved charges; discarding the ambiguity \(\mathbf {Y}\). This is because of \(\delta \mathbf {Y}(\delta _\eta \Phi ,\Phi )-\delta _\eta \mathbf {Y}(\delta \Phi ,\Phi )=0\), which is a result of the linearity of the left hand side in \(\delta _\eta \Phi =0\). Using this identity together with Eq. (6.5) in the (6.6), then there would not be any ambiguity in the definition of conserved charges as far as exact symmetries are concerned. Summarizing the last two paragraphs, the charges associated with exact symmetries are conserved, unambiguous, and independent of the chosen described surfaces of integration \(\partial \Sigma \).

So far, the SPSM has provided all materials needed to calculate \(\hat{\delta } H_\eta (p_j)\). The final tasks are checking integrability over \(\hat{\mathcal {M}}\), and (if integrable) performing the integration. The former is feasible simply by replacing \(\delta \Phi \) and \(\epsilon \) in Eq. (6.9) by \(\hat{\delta } \Phi \) and \(\eta \). The latter is abstractly the integration in Eq. (2.4), and pragmatically integrating \(\hat{\delta } H_\eta (p_j)\) over the parameters \(p_j\).

\(\varvec{k}_{\varvec{\xi }}\) for EH- \(\varvec{\Lambda }\) theory: To make the paper self-contained, here we provide the derivation of \(\varvec{k}_\xi \) for the EH-\(\Lambda \) theory, which is described by the Lagrangian density \(\mathcal {L}=\frac{1}{16\pi G} (R-2\Lambda )\). Beginning from the Eq. (6.1), one finds

$$\begin{aligned} \varvec{\Theta }(\delta \Phi ,\Phi )=\star \Big (\frac{1}{16\pi G}(\nabla _\alpha \delta g_{\,\,\mu }^{\alpha }-\nabla _\mu \delta g^\alpha _{\,\,\alpha })\,\mathrm {d}x^\mu \Big ). \end{aligned}$$
(6.11)

In order to find the explicit form of the \(\varvec{k}_\xi \) through Eq. (6.7), in addition to the equation above, the calculation of \(\delta \mathbf {Q}_\xi \) is also needed. To this end, by the definition (6.8) and using the equations of motion,

$$\begin{aligned} \mathbf {Q}_\xi&=\star \Big ( \frac{-1}{16\pi G }\frac{1}{2!} (\nabla _\mu \xi _\nu -\nabla _\nu \xi _\mu )\,\mathrm {d}x^\mu \wedge \mathrm {d}x^\nu \Big ) \end{aligned}$$
(6.12)
$$\begin{aligned}&=\frac{-1}{16\pi G}\frac{\sqrt{-g}}{(2!(d-2)!)}\epsilon _{\mu \nu \alpha _1\dots \alpha _{d-2}}(\nabla ^\mu \xi ^\nu -\nabla ^\nu \xi ^\mu )\,\mathrm {d}x^{\alpha _1}\wedge \dots \wedge \mathrm {d}x^{\alpha _{d-2}}. \end{aligned}$$
(6.13)

Now by the relations

$$\begin{aligned} \delta \sqrt{-g}=\frac{\sqrt{-g}}{2}\delta g^\alpha _{\,\,\alpha }, \quad \delta \Gamma ^\lambda _{\mu \nu }&= \frac{1}{2}[g^{\lambda \sigma }\big (\nabla _\mu \delta g_{\sigma \nu }+\nabla _\nu \delta g_{\sigma \mu }-\nabla _\sigma \delta g_{\mu \nu }\big )], \end{aligned}$$
(6.14)

one finds

$$\begin{aligned} \delta \mathbf {Q}_\xi =&\frac{-1}{16\pi G}\frac{\sqrt{-g}}{(2!(d-2)!)}\epsilon _{\mu \nu \alpha _1\dots \alpha _{d-2}}\Big (\frac{1}{2}\delta g^\alpha _{\,\,\alpha }(\nabla ^\mu \xi ^\nu )-\delta g^{\mu \beta }(\nabla _\beta \xi ^\nu )\nonumber \\&+\xi ^\alpha \nabla ^\mu \delta g^\nu _{\,\,\alpha }\Big )\,\mathrm {d}x^{\alpha _1}\wedge \dots \wedge \mathrm {d}x^{\alpha _{d-2}}-[\mu \leftrightarrow \nu ]. \end{aligned}$$
(6.15)

in which the notation \(\delta g^{\mu \nu }\equiv g^{\mu \alpha }g^{\nu \beta }\delta g_{\alpha \beta }=-\delta (g^{\mu \nu })\) has been used. Notice that by \(\delta (g^{\mu \nu })\) we meant the direct action of \(\delta \) on \(g^{\mu \nu }\). The next step in calculating the \(\varvec{k}_\xi \) would be finding the second term in (6.7), which is

$$\begin{aligned}&-\xi \cdot \varvec{\Theta }(\delta \Phi ,\Phi )\nonumber \\&\quad =-\xi \cdot \Big (\frac{1}{16\pi G}\frac{\sqrt{-g}}{(d-1)!}\epsilon _{\mu \alpha _{1}\dots \alpha _{d-1}}(\nabla _\alpha \delta g^{\alpha \mu }-\nabla ^\mu \delta g^\alpha _{\,\,\alpha })\mathrm {d}x^{\alpha _1}\wedge \dots \wedge \mathrm {d}x^{\alpha _{d-1}}\Big )\nonumber \\&\quad =\frac{-1}{16\pi G}\frac{\sqrt{-g}}{(d-2)!}\epsilon _{\mu \nu \alpha _{1}\dots \alpha _{d-2}}(\nabla _\alpha \delta g^{\alpha \mu }-\nabla ^\mu \delta g^\alpha _{\,\,\alpha })\xi ^\nu \,\mathrm {d}x^{\alpha _1}\wedge \dots \wedge \mathrm {d}x^{\alpha _{d-2}}\quad \end{aligned}$$
(6.16)
$$\begin{aligned}&\quad =\frac{-1}{16\pi G}\frac{\sqrt{-g}}{2(d-2)!}\epsilon _{\mu \nu \alpha _{1}\dots \alpha _{d-2}}(\nabla _\alpha \delta g^{\alpha \mu }-\nabla ^\mu \delta g^\alpha _{\,\,\alpha })\xi ^\nu \,\mathrm {d}x^{\alpha _1}\wedge \dots \wedge \mathrm {d}x^{\alpha _{d-2}}\nonumber \\&\quad \qquad -[\mu \leftrightarrow \nu ]. \end{aligned}$$
(6.17)

Finally, having found the (6.15) and (6.17), the \(\varvec{k}_\xi ^{\text {EH}}\) can be read as

$$\begin{aligned} \varvec{k}_\xi ^{\text {EH}}&=\frac{-1}{16\pi G}\frac{\sqrt{-g}}{(2!(d-2)!)}\epsilon _{\mu \nu \alpha _1\dots \alpha _{d-2}}\Big (\frac{1}{2}\delta g^\alpha _{\,\,\alpha }(\nabla ^\mu \xi ^\nu )-\delta g^{\mu \beta }(\nabla _\beta \xi ^\nu )+\xi ^\alpha \nabla ^\mu \delta g^\nu _{\,\,\alpha }\nonumber \\&\quad +(\nabla _\alpha \delta g^{\alpha \mu }-\nabla ^\mu \delta g^\alpha _{\,\,\alpha })\xi ^\nu \Big )\,\mathrm {d}x^{\alpha _1}\wedge \dots \wedge \mathrm {d}x^{\alpha _{d-2}}-[\mu \leftrightarrow \nu ]. \end{aligned}$$
(6.18)

By the Hodge duality, we would have \(\varvec{k}_\xi ^{\text {EH}}=\star k_\xi ^{\text {EH}}\), where

$$\begin{aligned} k_\xi ^{\text {EH}\mu \nu }= & {} \frac{-1}{16\pi G}\Big (\frac{1}{2}\delta g^\alpha _{\,\,\alpha }(\nabla ^\mu \xi ^\nu )-\delta g^{\mu \beta }(\nabla _\beta \xi ^\nu )+\xi ^\alpha \nabla ^\mu \delta g^\nu _{\,\,\alpha }\nonumber \\&+\,(\nabla _\alpha \delta g^{\alpha \mu }-\nabla ^\mu \delta g^\alpha _{\,\,\alpha })\xi ^\nu \Big )-[\mu \leftrightarrow \nu ]. \end{aligned}$$
(6.19)

Notice that this result is independent of the cosmological constant \(\Lambda \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hajian, K. Conserved charges and first law of thermodynamics for Kerr–de Sitter black holes. Gen Relativ Gravit 48, 114 (2016). https://doi.org/10.1007/s10714-016-2108-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10714-016-2108-4

Keywords

Navigation