Skip to main content
Log in

Energy in Fourth-Order Gravity

  • Original Paper
  • Published:
Annales Henri Poincaré Aims and scope Submit manuscript

Abstract

In this paper we make a detailed analysis of conservation principles in the context of a family of fourth-order gravitational theories generated via a quadratic Lagrangian. In particular, we focus on the associated notion of energy and start a program related to its study. We also exhibit examples of solutions which provide intuitions about this notion of energy which allows us to interpret it, and introduce several study cases where its analysis seems tractable. Finally, positive energy theorems are presented in restricted situations.

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 are parametrising these solutions in a convenient form for our purposes.

  2. Notice that the Einstein constraint equations imply that \(\Lambda \)-vacuum initial data sets, with \(\Lambda >0\), cannot be AE according to standard definitions.

  3. Such invariance property has been analysed in other important limiting cases in [10]

  4. Notations and conventions are detailed in the appendix Sect. 7.1

  5. The completely antisymmetric symbols \({\mu _{\hat{\bar{g}}}}_{\alpha _0\cdots \alpha _{n}}\) are defined by the relation \(dV_{\hat{\bar{g}}}={\mu _{\hat{\bar{g}}}}_{\alpha _0\cdots \alpha _{n}}\vartheta ^{\alpha _0}\wedge \cdots \wedge \vartheta ^{\alpha _n}\) for some positively oriented co-frame \(\{\vartheta ^{\alpha _0},\cdots ,\vartheta ^{\alpha _n} \}\).

  6. by admissible we mean all metric which actually remain static spherically symmetric. In particular having \(f\le 0\) means that the roles of r and t are exchanged and the space loses its static attribute.

  7. This would be \(g_{ij} = \delta _{ij} + O_2(r^{-\tau })\) and \(K_{ij} = O_1(r^{-\tau - 1})\), with \(\tau > 0\).

  8. The ± sign for \(\dot{N}\) depends on whether we are in the asymptotically expanding or contracting case (see Remark 5.1).

  9. See Theorem 5.2 for a detailed statement.

  10. Since below we shall only be interested in an explicit expression for \(\partial _t X^j\) near infinity, where \(|X|_e \rightarrow 0\), this condition will be satisfied. Within the general Cauchy problem, this does not pose a relevant problem, since the metric e is actually auxiliary. Thus, if \(X\ne 0\), one can chose an equivalent metric given by \(e'= \frac{1}{2 \max |X|_e} e,\) which guarantees that \(|X|_{e'} <1\) over all of M over all of M.

  11. In this case, in the \(o_4(|x|^{-\sigma })\) hypothesis we impose on \(\Omega \) implies that time-derivatives also increase the decay by one order.

References

  1. Abbott, L., Deser, S.: Stability Of Gravity With A Cosmological Constant. Nucl. Phys. B 195, 76 (1982)

    Article  ADS  Google Scholar 

  2. Abraham, R., Marsden, J., Ratiu, T.: Manifolds, Tensor Analysis, and Applications. Springer, Berlin (2001)

    Google Scholar 

  3. Adami, H., et al.: Conserved charges in extended theories of gravity. Phys. Rep. 834–835, 1–85 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  4. Alexakis, S., Mazzeo, R.: Complete Willmore surfaces in \(\mathbb{H} ^3\) with bounded energy: boundary regularity and bubbling. J. Diff. Geom. 101, 1 (2015)

    Google Scholar 

  5. Alty, L.: The generalized Gauss–Bonnet–Chern theorem. J. Math. Phys. 36, 3094 (1995)

    Article  ADS  MathSciNet  Google Scholar 

  6. Arms, J., Marsden, J., Moncrief, V.: The structure of the space of solutions of Einstein’s equations II: several killing fields and the Einstein–Yang–Mills equations. Ann. Phys. 144, 81–106 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  7. Arnowitt, R., Deser, S., Misner, C.: Coordinate invariance and energy expressions in general relativity. Phys. Rev. 122(3), 9971006 (1961)

    Article  MathSciNet  Google Scholar 

  8. Ashtekar, A., Das, S.: Asymptotically anti-de Sitter space-times: conserved quantities. Class. Quant. Grav. 17, L17 (2000)

    Article  ADS  Google Scholar 

  9. Avalos, R., Laurain, P., Marque, N.: Rigidity Theorems for Fourth Order Gravity (2022)

  10. Avalos, R., Laurain, P., Lira, J.: On the positive energy theorem for stationary solutions to fourth-order gravity. Calc. Var. 61, 48 (2022)

    Article  Google Scholar 

  11. Avalos, R., Lira, J.: Reduced thin-sandwich equations on manifolds euclidean at infinity and on closed manifolds: existence and multiplicity. J. Math. Phys. 61, 122501 (2020)

    Article  ADS  MathSciNet  Google Scholar 

  12. Avez, A.: Formule de Gauss–Bonnet–Chem en métrique de signature quelconque. C.R. Acad. Sci. 255, 2049–2051 (1962)

    Google Scholar 

  13. Bartnik, R.: The mass of an asymptotically flat manifold. Comm. Pure Appl. Math. 34, 661–693 (1986)

    Article  MathSciNet  Google Scholar 

  14. Bartnik, R.: Phase space for the Einstein equations. Commun. Anal. Geom. 13(5), 845–885 (2005)

    Article  MathSciNet  Google Scholar 

  15. Berezin, V., Dokuchaeva, V., Eroshenko, Y.: Spherically symmetric conformal gravity and “gravitational bubbles”. Astrophys. J. (1989)

  16. Boulware, D., Deser, S., Stelle, K.: Energy and supercharge in higher derivative gravity. Phys. Lett. B 168(4), 336–340 (1986). https://doi.org/10.1016/0370-2693(86)91640-0

    Article  ADS  MathSciNet  Google Scholar 

  17. Boulware, D., Deser, S., Stelle, K.: Properties of energy in higher derivative gravity theories. Quant. Field Theory Quant. Stat. 2, 101 (1987)

    MathSciNet  Google Scholar 

  18. Branson, T.: Differential operators canonically associated to a conformal structure. Math. Scand. 57, 293–345 (1985)

    Article  MathSciNet  Google Scholar 

  19. Branson, T.: The Functional Determinant. Lobal Analysis Research Center Lecture Note Series 4, Seoul National University (1993)

  20. Bryant, R.: A duality theorem for Willmore surfaces. J. Differ. Geom. 20, 23–53 (1984)

    Article  MathSciNet  Google Scholar 

  21. Burgess, C.: Quantum gravity in everyday life: general relativity as an effective field theory. Living Rev. Rel. 7, 5 (2004)

    Article  Google Scholar 

  22. Carlotto, A.: Four lectures on asymptotically flat riemannian manifolds. In: Einstein Equations: Physical and Mathematical Aspects of General Relativity (2019)

  23. Cederbaum, C., Sakovich, A.: On center of mass and foliations by constant spacetime mean curvature surfaces for isolated systems in General Relativity (2018). arXiv:1901.00028 [math.AP]

  24. Chang, A., Yang, P.: Extremal metrics of zeta function determinants on 4-manifolds. Ann. Math. 2 142(1), 171–212 (1995)

    Article  MathSciNet  Google Scholar 

  25. Chern, S.: On the curvatura integra in a Riemannian manifold. Ann. Math. 2nd Se. (1945)

  26. Choquet-Bruhat, Y.: The Cauchy Problem. Gravitation: An Introduction to Current Research (1962)

  27. Choquet-Bruhat, Y.: Positive-energy theorems. Relativite, groupes et topologie II/Relativity, groups and topology II (1984)

  28. Choquet-Bruhat, Y.: General Relativity and the Einstein Equations. Oxford Mathematical Monographs, pp. xxvi+785. Oxford University Press, Oxford (2009)

  29. Choquet-Bruhat, Y., Christodoulou, D., Francaviglia, M.: Cauchy data on a manifold. Annales de l’I. H. P., section A 29(3), 241–255 (1978)

    MathSciNet  Google Scholar 

  30. Choquet-Bruhat, Y., Isenberg, J., Pollack, D.: The constraint equations for the Einstein-scalar fi eld system on compact manifolds. Class. Quant. Grav. 24, 809 (2007)

    Article  ADS  Google Scholar 

  31. Chruściel, P.: Lectures on energy in general relativity, Kraków, March–April 2010. http://homepage.univie.ac.at/piotr.chrusciel (2013) ()

  32. Chruściel, P.: Lectures on Energy in General Relativity. https://homepage.univie.ac.at/piotr.chrusciel/teaching/Energy/Energy.pdf (2012)

  33. Chruściel, P., Corvino, J., Isenberg, J.: Construction of N-body initial data sets in general relativity. Commun. Math. Phys. 304, 637–647 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  34. Chruściel, P., Delay, E.: On mapping properties of the general relativistic constraints operator in weighted function spaces, with applications. M é m. Soc. Math. de France 94, 1–103 (2003)

    MathSciNet  Google Scholar 

  35. Chruściel, P., Isenberg, J., Pollack, D.: Initial Data Engineering. Commun. Math. Phys. 257, 29–42 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  36. Corvino, J.: Scalar Curvature Deformation and a Gluing Construction for the Einstein Constraint Equations. Commun. Math. Phys. 214, 137 (2000)

    Article  ADS  MathSciNet  Google Scholar 

  37. Corvino, J., Schoen, R.: On the asymptotics for the vacuum Einstein constraint equations. J. Differ. Geom. 73, 185–217 (2006)

    Article  MathSciNet  Google Scholar 

  38. Deser, S., Jackiw, R., Templeton, S.: Three-Dimensional Massive Gauge Theories. Phys. Rev. Lett. 48, 975 (1982)

    Article  ADS  Google Scholar 

  39. Deser, S., Jackiw, R., Templeton, S.: Topologically Massive Gauge Theories. Ann. Phys. 140, 372 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  40. Deser, S., Tekin, B.: Energy in generic higher curvature gravity theories. Phys. Rev. D (3) (2003)

  41. Deser, S., et al.: Critical points of D-dimensional extended gravities. Phy. Rev. D 83, 061502 (2011)

    Article  ADS  Google Scholar 

  42. Dilts, J., et al.: Non-CMC solutions of the Einstein constraint equations on asymptotically Euclidean manifolds. Class. Quant. Grav. 31, 065001 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  43. Djadli, Z., Hebey, E., Ledoux, M.: Paneitz-type operators and applications. Duke Math. J. 104(1), 129–169 (2000)

    Article  MathSciNet  Google Scholar 

  44. Djadli, Z., Malchiodi, A.: Existence of conformal metrics with constant \(Q\)-curvature. Ann. Math. 168, 813–858 (2008)

    Article  MathSciNet  Google Scholar 

  45. Donoghue, J.: General relativity as an effective field theory: the leading quantum corrections. Phy. Rev. D 50(6), 3874 (1994)

    Article  ADS  MathSciNet  Google Scholar 

  46. Dzhunushaliev, V., Schmidt, H.: New vacuum solutions of conformal Weyl gravity. J. Math. Phys. 41, 3007 (2000)

    Article  ADS  MathSciNet  Google Scholar 

  47. Eichmair, M., Metzger, J.: Unique isoperimetric foliations of asymptotically flat manifolds in all dimensions. Invent. Math. 1, 1–40 (2012)

    Google Scholar 

  48. Eichmair, M., et al.: The spacetime positive mass theorem in dimensions less than eight. J. Eur. Math. Soc. 18, 83–121 (2016)

    Article  MathSciNet  Google Scholar 

  49. Esposito, P., Robert, F.: Mountain pass critical points for Paneitz–Branson operators. Calc. Var. Part. Differ. Equ. 15(4), 493–517 (2002)

    Article  MathSciNet  Google Scholar 

  50. Fiedler, B., Schimming, R.: Exact solutions of the Bach field equations of general relativity. Rep. Math. Phys. 17, 15–36 (1980)

    Article  ADS  MathSciNet  Google Scholar 

  51. Fischer, A., Marsden, J.: Linearization stability of nonlinear partial differential equations. Proc. Symposia Pure Math. 27, 219 (1975)

    Article  MathSciNet  Google Scholar 

  52. Fischer, A., Marsden, J., Moncrief, V.: The structure of the space of solutions of Einstein’s equations. I. One Killing field. Annales de l’I. H. P. Section A tome 33(2), 147–194 (1980)

  53. Girbau, J., Bruna, L.: Stability by Linearization of Einstein’s Field Equation. Birkh ä user, Springer, Basel AG (2010)

    Book  Google Scholar 

  54. Graham, C., et al.: Conformally invariant powers of the Laplacian, I: Existence. J. Lond. Math. Soc. 46(2), 557–565 (1992)

    Article  MathSciNet  Google Scholar 

  55. Gursky, M.: The principal eigenvalue of a conformally invariant differential operator, with an application to semilinear elliptic PDE. Commun. Math. Phys. 207, 131–143 (1999)

    Article  ADS  MathSciNet  Google Scholar 

  56. Gursky, M., Malchiodi, A.: A strong maximum principle for the Paneitz operator and a non-local flow for the \(Q\)-curvature. J. Eur. Math. Soc. 17, 2137–2173 (2015)

    Article  MathSciNet  Google Scholar 

  57. Gursky, M., Malchiodi, A.: A strong maximum principle for the Paneitz operator and a non-local flow for the \(Q\)-curvature. J. Eur. Math. Soc. 17, 2137–2173 (2015)

    Article  MathSciNet  Google Scholar 

  58. Hang, F., Yang, P.: Sign of Green’s function of Paneitz operators and the \(Q\) curvature. Int. Math. Res. Not. IMRN 19, 9775–9791 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  59. Hang, F., Yang, P.: \(Q\) curvature on a class of \(3\)-manifolds. Commun. Pure Appl. Math. 69(4), 734–744 (2016)

    Article  MathSciNet  Google Scholar 

  60. Hang, F., Yang, P.: \(Q\) curvature on a class of manifolds with dimension at least \(5\). Commun. Pure Appl. Math. 69, 1452–1491 (2016)

    Article  MathSciNet  Google Scholar 

  61. Henneaux, M., Teitelboim, C.: Asymptotically Anti-De Sitter spaces. Commun. Math. Phys. 98, 391 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  62. Hollands, S., Ishibashi, A., Marolf, D.: Comparison between various notions of conserved charges in asymptotically AdS spacetimes. Class. Quant. Grav. 22, 2881 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  63. Holst, M., Meier, C.: Non-CMC solutions to the Einstein constraint equations on asymptotically Euclidean manifolds with apparent horizon boundaries. Class. Quant. Grav. 32, 025006 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  64. Holst, M., Nagy, G., Tsogtgerel, G.: Rough solutions of the Einstein constraints on closed manifolds without near-CMC conditions. Commun. Math. Phys. 288, 547 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  65. Holst, M., Tsogtgerel, G.: The Lichnerowicz equation on compact manifolds with boundary. Class. Quantum Grav. 30, 205011 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  66. Hooft, G., Veltman, M.: One loop divergencies in the theory of gravitation. Annales de l’Institut Henri Poincar é : Section A, Phys. Theor. 20, 69 (1974)

    ADS  MathSciNet  Google Scholar 

  67. Huisken, G., Yau, S.-T.: Definition of Center of Mass for Isolated Physical Systems and Unique Foliations by Stable Spheres with Constant Mean Curvature. Invent. Math. 124, 281–311 (1996)

    Article  ADS  MathSciNet  Google Scholar 

  68. Humbert, E., Raulot, S.: Positive mass theorem for the Paneitz–Branson operator. Calc. Var. Part. Differ. Equ. 36, 525–531 (2009)

    Article  MathSciNet  Google Scholar 

  69. Isenberg, J.: Constant mean curvature solution of the Einstein constraint equations on closed manifold. Class. Quant. Grav. 12, 2249–2274 (1995)

    Article  ADS  MathSciNet  Google Scholar 

  70. Isenberg, J., Maxwell, D., Pollack, D.: A gluing construction for non-vacuum solutions of the Einstein-constraint equations. Adv. Theor. Math. Phys. 9, 129–172 (2005)

    Article  MathSciNet  Google Scholar 

  71. Isenberg, J., Mazzeo, R., Pollack, D.: Gluing and Wormholes for the Einstein constraint equations. Commun. Math. Phys. 231, 529 (2002)

    Article  ADS  MathSciNet  Google Scholar 

  72. Kaku, M.: Quantization of conformal gravity: another approach to the renormalization of gravity. Nucl. Phys. B 203, 285–296 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  73. Kehagias, A., et al.: Black hole solutions in R 2 gravity. J. High Energy Phys. 2015, 1–20 (2015)

    Article  MathSciNet  Google Scholar 

  74. Kim, W., Kulkarni, S., Yi, S.: Quasilocal conserved charges in a covariant theory of gravity. Phy. Rev. Lett. 111, 081101 (2013)

    Article  ADS  Google Scholar 

  75. Lee, D.: Geometric Relativity. Graduate Studies in Mathematics, vol. 201. American Mathematical Society, Providence, RI (2019)

  76. Lee, J., Parker, T.: The Yamabe Problem. Bull. Am. Math. Soc. 17, 1 (1987)

    Article  MathSciNet  Google Scholar 

  77. Lee, J., Wald, R.: Local symmetries and constraints. J. Math. Phys. 31, 725 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  78. Lohkamp, J.: The Higher Dimensional Positive Mass Theorem I. (2016). arXiv:math/0608795 [math.DG]

  79. J. Lohkamp, The Higher Dimensional Positive Mass Theorem II. (2017). arXiv:1612.07505 [math.DG]

  80. Lü, H., Pope, C.: Critical gravity in four dimensions. Phy. Rev. Lett. 106, 181302 (2011)

    Article  ADS  Google Scholar 

  81. Lü, H., et al.: AdS and Lifshitz black holes in conformal and Einstein–Weyl gravities. Phy. Rev. D 86, 044011 (2012)

    Article  ADS  Google Scholar 

  82. Maldacena, J.: Einstein Gravity from Conformal Gravity. (2011). arXiv:1105.5632 [hep-th]

  83. Mannheim, P.: Making the case for conformal gravity. Found. Phys. 42, 388–420 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  84. Mannheim, P., Kazanas, D.: Exact vacuum solution to conformal Weyl gravity and galactic rotation curves. Astrophys. J. (1989)

  85. Marque, N.: Minimal bubbling for Willmore surfaces. In: International Mathematics Research Notices (2020). rnaa079. https://doi.org/10.1093/imrn/rnaa079

  86. Maxwell, D.: Rough solutions to the Einstein constraint equations on compact manifolds. J. Hyperbolic Differ. Equ. 2(2), 521–546 (2005)

    Article  MathSciNet  Google Scholar 

  87. Maxwell, D.: Solutions of the Einstein constraint equations with apparent horizon boundaries. Commun. Math. Phys. 253, 561–583 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  88. Maxwell, D.: A class of solutions of the vacuum Einstein constraint equations with freely specified mean curvature. Math. Res. Lett. 16(4), 627–645 (2009)

    Article  MathSciNet  Google Scholar 

  89. Moncrief, V.: Space-time symmetries and linearization stability of the Einstein equations. II. J. Math. Phys. 17, 1893 (1976)

    Article  ADS  MathSciNet  Google Scholar 

  90. Ndiaye, C.: Constant \(Q\)-curvature metrics in arbitrary dimension. J. Funct. Anal. 251, 1–58 (2007)

    Article  MathSciNet  Google Scholar 

  91. Nerz, C.: Foliations by stable spheres with constant mean curvature for isolated systems without asymptotic symmetry. Calc. Var. Part. Differ. Equ. 54(2), 1911–1946 (2015)

    Article  MathSciNet  Google Scholar 

  92. Nerz, C.: Foliations by spheres with constant expansion for isolated systems without asymptotic symmetry. J. Differ. Geom. 109(2), 257–289 (2018)

    Article  MathSciNet  Google Scholar 

  93. Noakes, D.: The initial value formulation of higher derivative gravity. J. Math. Phys. 24, 1846 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  94. Paneitz, S.: A quartic conformally covariant differential operator for arbitrary pseudo- Riemannian manifolds (summary). Symmet. Integrabil. Geom. Methods Appl. 4, 1 (2008)

    MathSciNet  Google Scholar 

  95. Regge, T., Teiltelboim, C.: Role of surface integrals in the Hamiltonian formulation of general relativity. Ann. Phys. 88, 286–318 (1974)

    Article  ADS  MathSciNet  Google Scholar 

  96. Schmidt, H.: A new conformal duality of spherically symmetric space-times. Annalen Phys. 9SI, 158–159 (2000)

  97. Schoen, R.: Conformal deformation of a Riemannian metric to constant scalar curvature. J. Differ. Geom 20, 479–595 (1984)

    Article  MathSciNet  Google Scholar 

  98. Schoen, R., Yau, S.-T.: On the proof of the positive mass conjecture in general relativity. Commun. Math. Phys. 65, 4576 (1979)

    Article  MathSciNet  Google Scholar 

  99. Schoen, R., Yau, S.-T.: Proof of the Positive Mass Theorem. II. Commun. Math. Phys. 79, 231–260 (1981)

    Article  ADS  MathSciNet  Google Scholar 

  100. Schoen, R., Yau, S.-T.: Positive Scalar Curvature and Minimal Hypersurface Singularities (2017). arXiv:1704.05490 [math.DG]

  101. Spivak, M.: A Comprehensive Introduction to Differential Geometry, vol. I. PUBLISH OR PERISH, INC (1999)

  102. Starobinsky, A.: A new type of isotropic cosmological models without singularity. Phys. Lett. B 91, 99–102 (1980)

    Article  ADS  Google Scholar 

  103. Stelle, K.: Renormalization of higher-derivative quantum gravity. Phy. Rev. D 16(4), 953 (1977)

    Article  ADS  MathSciNet  Google Scholar 

  104. Stelle, K.: Classical Gravity with Higher Derivatives. Gen. Rel. Grav. 9(4), 355–371 (1978)

    Article  ADS  MathSciNet  Google Scholar 

  105. Wald, R., Zoupas, A.: A general definition of “conserved quantities’’ in general relativity and other theories of gravity. Phys. Rev. D 61, 084027 (2000)

    Article  ADS  MathSciNet  Google Scholar 

  106. Willmore, T.J.: Riemannian geometry. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, pp. xii+318 (1993)

  107. Witten, E.: A new proof of the positive energy theorem. Commun. Math. Phys. 80, 381–402 (1981)

    Article  ADS  MathSciNet  Google Scholar 

  108. Zwiebach, B.: Curvature squared terms and string theories. Phys. Lett. B 156, 315 (1985)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the CAPES-COFECUB and CAPES-PNPD for their financial support and Paul Laurain for insightful discussions on this topic. This article was submitted for publication when the first author was employed by the University of Ceará, and the third author by the University of Potsdam. The first author would also like to thank the University of Tübingen and the Alexander von Humboldt Foundation for partial financial support related to the research presented in this paper. Similarly, the third author would like to thank the Université de Lorraine.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to R. Avalos.

Additional information

Communicated by Mihalis Dafermos.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 Geometric Conventions

1.1.1 Curvature Conventions

To avoid any ambiguity let us pinpoint the curvature conventions we follow in this text: the curvature tensor is defined as:

$$\begin{aligned} R(X,Y)Z=\nabla _X\nabla _YZ - \nabla _Y\nabla _XZ - \nabla _{[X,Y]}Z, \end{aligned}$$

with its components ordered as follows:

$$\begin{aligned} R^{i}_{jkl}&=dx^{i}(R(\partial _k,\partial _l)\partial _j),\\&=\partial _k\Gamma ^{i}_{lj}-\partial _l\Gamma ^{i}_{kj} + \Gamma ^{i}_{ku}\Gamma ^u_{jl}- \Gamma ^{i}_{lu}\Gamma ^u_{jk}.\\ \end{aligned}$$

From this we get the canonical Ricci, Einstein, and scalar tensors:

$$\begin{aligned} \textrm{Ric}_{ij}\doteq R^l_{ilj}. \end{aligned}$$

1.1.2 Differential Forms

In this section we will establish our conventions concerning differential forms and operations related to them. First of all, given a k-form \(\omega \in \Omega ^k(V)\) on a d-dimensional manifold V and a local coordinate system \(\{x^i\}_{i=1}^d\), we fix \(\omega _{i_1\cdots i_k}\doteq \omega (\partial _{i_1},\cdots ,\partial _{i_k})\) as its components relative to the basis \(\{dx^{j_1}\wedge \cdots dx^{j_k}\}_{j_1<\cdots <j_k}\), and therefore we may locally write

$$\begin{aligned} \omega =\sum _{i_1<\cdots <i_{k}}\omega _{i_1\cdots i_k}dx^{i_1}\wedge \cdots dx^{i_k}=\frac{1}{k!}\omega _{i_1\cdots i_k}dx^{i_1}\wedge \cdots dx^{i_k}, \end{aligned}$$

where in the last equality the summation is not restricted to \(i_1<\cdots <i_k\). In this setting we have several well-known operations. To start with, given a semi-Riemannian manifold \((V^d,g)\) we have an induced semi-Riemannian metric \(g^{(k)}\) on each space \(\Omega ^k(V)\) which is given by

$$\begin{aligned} g^{(k)}(\alpha ,\beta )\doteq \sum _{i_1<\cdots <i_k}\alpha _{i_1\cdots i_k}\beta ^{i_1\cdots i_k}=\frac{1}{k!}\alpha _{i_1\cdots i_k}\beta ^{i_1\cdots i_k}, \text { for any } \alpha ,\beta \in \Omega ^k(V) \end{aligned}$$

where above \(\beta ^{i_1\cdots i_k}\doteq g^{i_1j_1}\cdots g^{i_kj_k}\beta _{j_1\cdots j_k}\). It is not difficult to see that if \(\{e_1,\cdots ,e_d\}\) is a g-orthonormal basis for \(T_pV\) at a point \(p\in V\), and if \(\{e^1,\cdots ,\) \(e^d\}\) stands for its dual basis, then \(\{e^i_{1}\wedge \cdots \wedge e^{i_k}\}_{i_1<\cdots <i_k}\) is a \(g^{(k)}\)-orthonormal basis for the fibre of \(\Omega ^k(V)\) over p (see, for instance, [2, Proposition 7.2.11]). In particular, it holds that

$$\begin{aligned} g^{(k)}(e^i_{1}\wedge \cdots \wedge e^{i_k},e^i_{1}\wedge \cdots \wedge e^{i_k})=c_{i_1}\cdots c_{i_k} \end{aligned}$$

where above \(c_{i_{j}}\doteq g(e_{i_j},e_{i_j})=\pm 1\).

Also, in this setting, we introduce the Hodge-star operator, which is defined pointwise as a linear operator on each fibre. For this one needs to introduce a volume form. Thus, let us again consider an orientable semi-Riemannian manifold \((V^d,g)\) and denote the associated Riemannian volume form by \(dV_g\), which locally reads as \(dV_g=\sqrt{|\textrm{det}(g)|}dx^1\wedge \cdots \wedge dx^d\). Given our vector bundle \(\Omega ^k(V)\xrightarrow []{\pi } V\), one defines a linear operator

$$\begin{aligned} \star _{g}:\Omega ^{k}(V)\mapsto \Omega ^{d-k}(V) \end{aligned}$$

by the requirement that for all \(\alpha ,\beta \in \Omega ^{k}(V)\) it holds that

$$\begin{aligned} \beta \wedge \star _{g}\alpha =g^{(k)}(\beta ,\star _g \alpha )dV_g. \end{aligned}$$

Above, we add the subscript g on \(\star \) to highlight the dependence on this operator on the choice of metric g. This can be seen to be well-defined and, in fact, in local (oriented) coordinates one can see that the action of \(\star \) on \(\alpha \in \Omega ^k(V)\) is given by (see, for instance, [2, Proposition 7.2.12 and Example 7.2.14(D)]):

$$\begin{aligned} \star _g\alpha&=\sum _{\underset{j_{k+1}<\cdots<j_d}{j_1<\cdots<j_k}}\alpha ^{j_1\cdots j_k}\epsilon _{j_1\cdots j_k j_{k+1}\cdots j_d}\sqrt{|\textrm{det}(g)|}dx^{j_{k+1}}\wedge \cdots \wedge dx^{j_d},\\ (\star _g\alpha )_{j_{k+1}\cdots j_{d}}&=\sum _{j_1<\cdots <j_k}\alpha ^{j_1\cdots j_k}\epsilon _{j_1\cdots j_k j_{k+1}\cdots j_d}\sqrt{|\textrm{det}(g)|}, \end{aligned}$$

where above \(\epsilon _{j_1\cdots j_d}\) stands for the antisymmetric Levi-Civita symbol and the sums which appeal to Einstein summation convention are understood as over all possible values of the indices. Sometimes, we shall write \(\mu _{j_1\cdots j_d}=\epsilon _{j_1\cdots j_d}\) \(\sqrt{|\textrm{det}(g)|}\) which is defined via the relation

$$\begin{aligned} dV_g=\mu _{j_1\cdots j_d}dx^{j_1}\wedge \cdots \wedge dx^{j_d} \text { (no summation) }. \end{aligned}$$

Let us now recall that, given \(\alpha \in \Omega ^k(V)\), the exterior differential \(d:\Omega ^k(V)\mapsto \Omega ^{k+1}(V)\) is characterised by its local action given by

$$\begin{aligned} d\alpha =\sum _{i_1<\cdots <i_{k}}\partial _{i}\alpha _{i\cdots i_{k+1}}dx^{i}\wedge dx^{i_1}\wedge \cdots \wedge dx^{k}. \end{aligned}$$
(72)

In particular, given a 1-form \(\alpha \in \Omega ^1(V)\), we see that

$$\begin{aligned} d\alpha&=\sum _{j=1}^d\partial _{i}\alpha _{j}dx^{i}\wedge dx^{j}=\partial _{i}\alpha _{j}dx^{i}\wedge dx^{j}=\sum _{i<j}(\partial _i\alpha _j-\partial _j\alpha _i)dx^i\wedge dx^j,\\ d\alpha _{ij}&=\partial _i\alpha _j-\partial _j\alpha _i. \end{aligned}$$

Now, given a semi-Riemannian manifold \((V^d,g)\) we can consider the formal adjoint \(d^{*}:\Omega ^{k+1}(V)\mapsto \Omega ^{k}(V)\), which we shall denote by \(\delta _{g}\) and where we shall highlight its dependence on g, which arises through the canonical choice of \(dV_g\) as the volume form. One can thus globally write

$$\begin{aligned} \delta _g\alpha =(-1)^{dk+1+\textrm{Ind}(g)}\star _gd\star _g\alpha \text { for all } \alpha \in \Omega ^{k+1}(V), \end{aligned}$$

where \(\textrm{Ind}(g)\) denotes the index of the semi-Riemannian metric g. For instance, if g is Lorentzian, we have \(\textrm{Ind}(g)=1\). One can then compute that, in local coordinates, the following formula holds:

$$\begin{aligned} \delta _g\alpha _{i_1\cdots i_k}=-\nabla ^i\alpha _{ii_1\cdots i_k}, \end{aligned}$$

where above \(\nabla \) stands for the Riemannian connection associated to g.

Also, given \(X\in \Gamma (TV)\), let us introduce the interior product \(X\lrcorner :\Omega ^k(V)\mapsto \Omega ^{k-1}(TV)\) on any manifold \(V^d\), which is defined by

$$\begin{aligned}&X\lrcorner \alpha (X_1,\cdots ,X_{k-1})\doteq \alpha (X,X_1,\cdots ,X_{k-1}) \text { for any } \alpha \in \Omega (V)\\&\quad \text { and } \forall \,\, X_1,\cdots ,X_{k-1}\in \Gamma (TV). \end{aligned}$$

An important formula linking the exterior derivative, the interior product, and the Lie derivative is given by Cartan’s famous magic formula:

$$\begin{aligned} \pounds _X\alpha =d(X\lrcorner \alpha ) + X\lrcorner (d\alpha ) \,\, \forall \,\, \alpha \in \Omega ^k(V) \text { and } X\in \Gamma (TV). \end{aligned}$$

The above formula plays an important role in the application of Stokes’ theorem to operators in divergence form. Since in the core of this paper we will be interested in certain flux formalae which are derived in this manner within the Lorentzian setting, let us below briefly highlight a few differences with the more usual Riemannian setting.

Let \((V^{n+1}=M^n\times \mathbb {R}, \bar{g})\) be a Lorentzian manifold, parametrise the \(\mathbb {R}\) factor with a coordinate t. Assume \(\partial _t\) is time-like and denote by \(g_t\) the induced Riemannian metric on each \(M_t\doteq M\times \{t\}\). Assume furthermore that \((M^n,g_t)\) are orientable Riemannian manifolds and denote by \(dV_{g_t}\) their corresponding volume forms. With all this, we have a natural orientation for V: if at \(p\in M\) \(\{e_1,\cdots ,e_n\}\) denotes a positive basis for \(T_pM\), then \(\{\partial _t,e_1,\cdots ,e_n\}\) denotes a positive basis for \(T_{(p,t)}V\). Thus, if \(\{x^i\}_{i=1}^n\) is a positively oriented coordinate system for M, then \(dV_{\bar{g}}=\sqrt{|\textrm{det}(\bar{g})|}dt\wedge dx^1\wedge \cdots \wedge dx^n\) denotes our volume form.

Let \(\Omega \subset M\) be a compact subset with smooth boundary and define the subset \(\mathcal {C}_T=\Omega \times [0,T]\). This subset is Stokes regular, in the sense that it is regular enough so as to apply Stokes’ theorem over it. Now, let \(X\in \Gamma (TV)\) and from the formula

$$\begin{aligned} L_XdV_{\bar{g}}=\textrm{div}_{\bar{g}}XdV_{\bar{g}}=d(X\lrcorner dV_{\bar{g}}) \end{aligned}$$

one gets that

$$\begin{aligned} \int _{\mathcal {C}_T}\textrm{div}_{\bar{g}}XdV_{\bar{g}}&=\int _{\partial \mathcal {C}_T}\mathcal {J}^{*}(X\lrcorner dV_{\bar{g}}) \end{aligned}$$
(73)

where above \(\mathcal {J}^{*}:\partial \mathcal {C}_T\mapsto \mathcal {C}_T\) denotes the inclusion. We can split \(\partial \mathcal {C}_{T}=\Omega _0\cup \Omega _T\cup L\), where \(\Omega _0=\Omega \times \{0\}\), \(\Omega _T=\Omega \times \{T\}\) and \(L=\partial \Omega \times [0,T]\). On each of these hypersurfaces we denote the inclusion into V by \(\mathcal {J}^{*}_0,\mathcal {J}^{*}_T\) and \(\mathcal {J}^{*}_L\), respectively. Now, let n denote the future-pointing unit normal to each t-constant hypersurface \(M_t=M\times \{t\}\). Then, writing \(X=-\bar{g}(X,n)n + X^{\top }\), we find

$$\begin{aligned} \mathcal {J}^{*}_0(X\lrcorner dV_{\bar{g}})=\mathcal {J}^{*}_0(dV_{\bar{g}}(-\bar{g}(X,n)n + X^{\top },\cdot ))=-\bar{g}(X,n)\mathcal {J}^{*}_0(dV_{\bar{g}}(n,\cdot )). \end{aligned}$$

Notice that in (73) \(\Omega _0\) is oriented with its Stokes induced orientation, where the outward-pointing unit normal corresponds to \(-n\) and thus \(\{e_1,\cdots ,e_n\}\) is a positive basis for \(\Omega _0\) at p iff \(\{-n,e_1,\cdots ,e_n\}\) is positive for \(\mathcal {C}_T\). This implies that the induced Stokes orientation for \(\Omega _0\) is actually opposite to its intrinsic orientation. On the other hand, we see that in the case of \(\Omega _T\) these two orientations agree. All this implies that

$$\begin{aligned} \mathcal {J}^{*}_0(X\lrcorner dV_{\bar{g}})=-\bar{g}(X,n)dV_{\bar{g}_0} \, ,\, \mathcal {J}^{*}_T(X\lrcorner dV_{\bar{g}})=-\bar{g}(X,n)dV_{\bar{g}_T} \end{aligned}$$

and, using intrinsic orientations for \(\Omega \),

$$\begin{aligned} \int _{\mathcal {C}_T}\textrm{div}_{\bar{g}}XdV_{\bar{g}}&=-\int _{\Omega _0}\mathcal {J}^{*}_0(X\lrcorner dV_{\bar{g}}) + \int _{\Omega _T}\mathcal {J}^{*}_T(X\lrcorner dV_{\bar{g}}) + \int _{L}\mathcal {J}^{*}_L(X\lrcorner dV_{\bar{g}}),\nonumber \\&=\int _{\Omega _0}\bar{g}(X,n)dV_{\bar{g}_0} - \int _{\Omega _T}\bar{g}(X,n)dV_{\bar{g}_T} + \int _{L}\mathcal {J}^{*}_L(X\lrcorner dV_{\bar{g}}), \end{aligned}$$
(74)

where \({J}^{*}_L(X\lrcorner dV_{\bar{g}})=g_t(X,\nu )\nu \lrcorner dV_{\bar{g}}=g_t(X,\nu )dL\), with \(\nu \) the outward pointing unit normal vector field to L, which is to be understood with its induced Stokes orientation.

1.1.3 Extrinsic Geometry

In this section we shall quickly fix our conventions for the extrinsic curvature. Thus, let \(M^n\hookrightarrow (V^{n+1},\bar{g})\) be an immersed hypersurface in a time-oriented Lorentzian manifold. We define the second fundamental form of M as

$$\begin{aligned} \mathbb{I}\mathbb{I}:\Gamma (TM)\times \Gamma (TM)&\mapsto \Gamma (TM^{\perp })\\ (X,Y)&\mapsto (\bar{\nabla }_{\bar{X}}\bar{Y})^{\perp } \end{aligned}$$

where \(\bar{X}\) and \(\bar{Y}\) denote arbitrary extensions (respectively) of X and Y to V, \(\bar{\nabla }\) denotes the Riemannian connection associated to \(\bar{g}\), and \(TM^{\perp }\) denotes the normal bundle of M. Associated to the second fundamental form, we have the extrinsic curvature, here denoted by \(K\in \Gamma (T^0_2M)\), which we define with respect to the future-pointing unit normal to M. Thus, K is defined by

$$\begin{aligned} K(X,Y)\doteq \bar{g}(\mathbb{I}\mathbb{I}(\bar{X},\bar{Y}),n)=\bar{g}(\bar{\nabla }_{\bar{X}}\bar{Y},n), \,\, \forall \, X,Y\in \Gamma (TM). \end{aligned}$$

Also, we define \(\tau \doteq \textrm{tr}_gK\) as the (not normalised) mean curvature of the immersion, and therefore we find that

$$\begin{aligned} \tau =-\textrm{div}_{\bar{g}}n. \end{aligned}$$

Finally, let us notice that if \(V^{n+1}=M^n\times \mathbb {R}\), with \(\mathbb {R}\) parametrised by a coordinate t and the time orientation given by \(\partial _t\), then defining the associated lapse N and shift X by

$$\begin{aligned} N=-\bar{g}(\partial _t,n),\,\, X=\partial _t-Nn. \end{aligned}$$

we may write \(n=\frac{1}{N}(\partial _t-X)\). This implies that

$$\begin{aligned} -K(U,W)&=\bar{g}(\bar{U},\bar{\nabla }_{\bar{W}}n) =\frac{1}{N}\left( \bar{g}(\bar{U},\bar{\nabla }_{\bar{W}}\partial _t) - \bar{g}(\bar{U},\bar{\nabla }_{\bar{W}}X)\right) ,\\&=\frac{1}{N}\left( \partial _t(\bar{g}(\bar{U},\bar{W})) - \bar{g}(\bar{\nabla }_{\partial _t}\bar{U},\bar{W}) + \bar{g}(\bar{U},[\bar{W},\partial _t]) - g(U,\nabla _{W}X)\right) \end{aligned}$$

Therefore,

$$\begin{aligned} -2NK(U,W) =\partial _t(\bar{g}(\bar{U},\bar{W})) - \bar{g}(\bar{U},[\partial _t,\bar{W}]) - \bar{g}(\bar{W},[\partial _t,\bar{U}]) - \pounds _{X}g(U,W). \end{aligned}$$

That is,

$$\begin{aligned} K(U,W)=-\frac{1}{2N}\left( \partial _t(\bar{g}(\bar{U},\bar{W})) - \bar{g}(\bar{U},[\partial _t,\bar{W}]) - \bar{g}(\bar{W},[\partial _t,\bar{U}]) - \pounds _{X}g(U,W)\right) \end{aligned}$$

Notice that locally this reduces to a simple expression given by

$$\begin{aligned} K_{ij}=-\frac{1}{2N}\left( \partial _t\bar{g}_{ij} - \pounds _{X}g_{ij}\right) , \end{aligned}$$

from which we sometimes write \(K=-\frac{1}{2N}\left( \partial _tg - \pounds _{X}g \right) \).

1.2 Proof of Proposition 4.1

Since the sketch presented has already dealt with the \(2 \alpha + \beta = 0\) case, we will assume in the following that \(\chi := 2 \alpha + \beta \ne 0\). Employing the same Maple procedure as presented in Fig. 1, we can see (see Fig. 3) that, as announced in the sketch of the proof: \(M= -m -\frac{\Lambda }{3}r^3 +C_1 r^{f(\alpha ,\beta )} + C_2 r^{g(\alpha ,\beta )}\), with

$$\begin{aligned} f&= \frac{6 \alpha +3 \beta + \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{2\left( 2 \alpha + \beta \right) } \\ g&= \frac{6 \alpha +3 \beta - \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{2\left( 2 \alpha + \beta \right) } \end{aligned}$$

and

$$\begin{aligned} \frac{\left( 2 \alpha + \beta \right) ^2 r }{24 \left( \beta + 4 \alpha \right) }A_{22}&= C_1\left( h^+_1 r^{t_1^+(\alpha , \beta ) }+h^+_2r^{t_2^+(\alpha , \beta ) }+h^+_3 r^{t_3^+(\alpha , \beta ) } + h^+_4 r^{t_4^+(\alpha , \beta ) } \right) \\&\quad + C_2\left( h^-_1 r^{t_1^-(\alpha , \beta ) }+h^-_2 r^{t_2^-(\alpha , \beta ) }+h^-_3 r^{t_3^-(\alpha , \beta ) }+ h^-_4 r^{t_4^-(\alpha , \beta ) } \right) , \end{aligned}$$

with

$$\begin{aligned} t_1^-&=\frac{18 \alpha +9 \beta - \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{4 \alpha + 2 \beta }\nonumber \\ t_2^-&=\frac{10 \alpha +5 \beta - \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{4 \alpha + 2 \beta }\nonumber \\ t_3^-&=\frac{6 \alpha +3 \beta - \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{4 \alpha + 2 \beta }\nonumber \\ t_4^-&=\frac{6 \alpha +3\beta - \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{2 \alpha + \beta }\nonumber \\ t_1^+&=\frac{18 \alpha +9 \beta +\sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{4 \alpha + 2 \beta }\nonumber \\ t_2^+&=\frac{10 \alpha +5 \beta + \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{4 \alpha + 2 \beta }\nonumber \\ t_3^+&=\frac{6 \alpha +3 \beta +\sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{4 \alpha + 2 \beta }\nonumber \\ t_4^+&=\frac{6 \alpha +3\beta +\sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{2 \alpha + \beta }. \end{aligned}$$
(75)

The \(4 \alpha + \beta =0\) case will be treated separately as a special case.

Fig. 3
figure 3

Necessary conditions on M

Remark 7.1

We must point out that \(100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2\) is not necessarily positive. We will consider first that \(100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2 \ge 0\), before explaining that the situation is highly similar in the opposite case.

Since \(100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2 = 25 \left( 2 \alpha + \beta \right) ^2 -8 \beta \left( 2 \alpha + \beta \right) = 25 \chi ^2 - 8 \beta \chi \), we will favour working with \((\chi , \beta )\) instead of \((\alpha , \beta )\). We will thus write:

$$\begin{aligned} f&=\frac{3}{2}+\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2} \nonumber \\ g&= \frac{3}{2}-\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}, \end{aligned}$$
(76)

and

$$\begin{aligned} t_1^-&=\frac{9}{2} -\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2} \nonumber \\ t_2^-&=\frac{5}{2} -\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}\nonumber \\ t_3^-&=\frac{3}{2} -\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}\nonumber \\ t_4^-&=3 -\sqrt{25 -8 \frac{\beta }{\chi } }\nonumber \\ t_1^+&=\frac{9}{2} +\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}\nonumber \\ t_2^+&=\frac{5}{2} +\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}\nonumber \\ t_3^+&=\frac{3}{2} +\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}\nonumber \\ t_4^+&=3 +\sqrt{25 -8 \frac{\beta }{\chi } }. \end{aligned}$$
(77)

This of course works under the assumption that \(\chi \) is positive. However, if \(\chi \le 0\), \( \frac{ \sqrt{ 100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2} }{\chi } = \textrm{sg}(\chi ) \sqrt{25 -8 \frac{\beta }{\chi }}\). Up to exchanging f and g, and the \(t_i^+\) and the \(t_i^-\), the coefficients remain the same.

Using these formulas, one can deduce that there exist a finite number of \([\alpha , \beta ] \in \mathbb {R}\mathbb {P}^1\) for which one of the \(t_i^\pm \) equals another \(t_j^\pm \). We present those values in the table, Fig. 4 and detail a few representative cases:

  • \(t_1^- =t_2^-, t_3^-\) clearly has no solution.

  • \(t_1^-=t_4^- \) is equivalent to \(\frac{9}{2}-\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}=3 -\sqrt{25 -8 \frac{\beta }{\chi } }\) which is rephrased as \(\sqrt{25 -8 \frac{\beta }{\chi } }=-3\). There are then no solutions.

  • \(t_1^-=t_1^+\) if and only if \(\sqrt{25 -8 \frac{\beta }{\chi } }=0\), i.e. \(\frac{\beta }{\chi }=\frac{25}{8}.\)

  • \(t_1^-=t_2^+\) if and only if \(\frac{9}{2} -\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}=\frac{5}{2} +\frac{\sqrt{25 -8 \frac{\beta }{\chi } } }{2}\), i.e. \(\sqrt{25 -8 \frac{\beta }{\chi } }=2\), which means that: \(\frac{\beta }{\chi }=\frac{21}{8}\).

All the other combinations fall into one of these configurations (obviously no solution, no solution because of negative squareroot, solution with null squareroot, solution with positive squareroot).

Fig. 4
figure 4

Values of \(\frac{\beta }{\chi }\) for which \(t_i^\pm =t_j^\pm \)

Outside of those specific values, the \(\left( r^{t_i^\pm } \right) \) form a free family. Thus for the metric to be A flat one must have:

$$\begin{aligned} \left( -\left( \beta + 4 \alpha \right) \sqrt{ 100 \alpha ^2+ 84 \alpha \beta + 17 \beta ^2} + \left( 2 \alpha + \beta \right) \left( 7 \beta + 20 \alpha \right) \right) C_2=0. \end{aligned}$$
(78)

This equation is obtained by looking at the \(r^{\frac{10 \alpha +5 \beta - \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{4 \alpha + 2 \beta }}\) term (last term of the first line in formula (2) in Fig. 3). We will rephrase (78) in terms of \(\beta \) and \(\chi \):

$$\begin{aligned} C_2 \left( - \left( 2-\frac{\beta }{\chi } \right) \sqrt{ 25 - 8 \frac{\beta }{\chi } } +10- 3 \frac{\beta }{\chi } \right) =0, \end{aligned}$$

which implies that either \(C_2=0\) or \(\frac{\beta }{\chi }=0\), i.e. \(\beta =0\).

Similarly, looking at the \(r^{\frac{10 \alpha +5 \beta + \sqrt{100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2}}{4 \alpha + 2 \beta }}\) term (the term of (2) in Fig. 3 between the third and fourth line):

$$\begin{aligned} \left( \left( \beta + 4 \alpha \right) \sqrt{ 100 \alpha ^2+ 84 \alpha \beta + 17 \beta ^2} + \left( 2 \alpha + \beta \right) \left( 7 \beta + 20 \alpha \right) \right) C_1=0. \end{aligned}$$
(79)

We once more rephrase this as:

$$\begin{aligned} C_1 \left( \left( 2-\frac{\beta }{\chi } \right) \sqrt{ 25 - 8 \frac{\beta }{\chi } } +10- 3 \frac{\beta }{\chi } \right) =0, \end{aligned}$$

which implies that \(C_1=0\) or \(\frac{\beta }{\chi }=3\).

Thus outside of \(\frac{\beta }{\chi } = 2, \) \(\frac{25}{8}\) \(\frac{21}{8}\), \(\frac{28}{9}\) 0, 3, one must have \(C_1=C_2=0\), which implies that the metric is Schwarzschild-de Sitter (or AdS). We only have to test these remaining values to conclude. For convenience, and in order to use the same Maple procedure, we will rephrase those in term of \(\alpha \) and \(\beta \). We need to test the cases: \( \beta +4 \alpha =0\), \(50 \alpha + 17 \beta =0\), \(42 \alpha +13 \beta =0\), \(56 \alpha + 19 \beta =0\), \(\beta =0\), \(3 \alpha + \beta =0\). Actually, this last case corresponds to the conformally invariant one and will not be considered here (see Proposition 4.2 for this configuration).

On the Maple results displayed in Fig. 5, one can see that for \( \beta +4 \alpha =0\), \(42 \alpha +13 \beta =0\), \(56 \alpha + 19 \beta =0\) one must have \(C_1=C_2=0\), which is the desired result. In the configuration \(50 \alpha + 17 \beta =0\) however, one obtains only \(C_1+C_2=0\). Nevertheless, since this corresponds to the case where \(f=g\), one concludes that \(M(r)= m + \frac{\Lambda }{3}r^3\) (second line of (4) in Fig. 5), which implies that the metric is indeed Schwarzschild-de Sitter (or AdS).

Fig. 5
figure 5

Cases \( \beta +4 \alpha =0\), \(50 \alpha + 17 \beta =0\), \(42 \alpha +13 \beta =0\), \(56 \alpha + 19 \beta =0\)

In the final case: \(\beta = 0,\) while \(C_1=0\), a priori \(\Lambda =0\), \(C_2 \ne 0\) is an admissible solution, corresponding to the Reissner-Nordström metric. We can check that it is indeed a solution (see Fig. 6) and conclude the proof.

Fig. 6
figure 6

The Reissner-Nordström metric is \(A_{\alpha , 0}\)-flat

Of course the above stands when \(100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2\ge 0\). The reasoning when \(100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2<0\) will be very similar. We will thus give a brief overview of the proof in that case: one simply has to replace f and g by

$$\begin{aligned} f&= \frac{6 \alpha +3 \beta + i\sqrt{\left| 100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2 \right| }}{2\left( 2 \alpha + \beta \right) } \\ g&= \frac{6 \alpha +3 \beta - i\sqrt{\left| 100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2\right| }}{2\left( 2 \alpha + \beta \right) }. \end{aligned}$$

The algebraic operations will remain the same even with complex exponents, and thus A will be written as a sum of (complex) powers of r. One simply has to replace the \(t_i^{\pm }\) by:

$$\begin{aligned} t_1^-&=\frac{18 \alpha +9 \beta - i\sqrt{|100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2|}}{4 \alpha + 2 \beta }\\ t_2^-&=\frac{10 \alpha +5 \beta - i\sqrt{|100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2|}}{4 \alpha + 2 \beta }\\ t_3^-&=\frac{6 \alpha +3 \beta - i\sqrt{|100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2|}}{4 \alpha + 2 \beta }\\ t_4^-&=\frac{6 \alpha +3\beta - i\sqrt{|100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2|}}{2 \alpha + \beta }\\ t_1^+&=\frac{18 \alpha +9 \beta +i\sqrt{|100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2|}}{4 \alpha + 2 \beta }\\ t_2^+&=\frac{10 \alpha +5 \beta + i\sqrt{|100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2|}}{4 \alpha + 2 \beta }\\ t_3^+&=\frac{6 \alpha +3 \beta +i\sqrt{|100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2|}}{4 \alpha + 2 \beta }\\ t_4^+&=\frac{6 \alpha +3\beta +i\sqrt{|100 \alpha ^2 + 84 \alpha \beta + 17 \beta ^2|}}{2 \alpha + \beta }. \end{aligned}$$

In this case, the \(t_i^{\pm }\) cannot interfere and thus the \(r^{t_i^{\pm }}\) form a free family. One can then, mutatis mutandis, look at (78) and (79) in the same manner as before, and conclude that \(C_1=C_2=0\), and thus that the metric is Schwarzschild-de Sitter (or AdS).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Avalos, R., Lira, J.H. & Marque, N. Energy in Fourth-Order Gravity. Ann. Henri Poincaré (2024). https://doi.org/10.1007/s00023-024-01440-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00023-024-01440-3

Navigation