Skip to main content
Log in

Mass Inflation and the \(C^2\)-inextendibility of Spherically Symmetric Charged Scalar Field Dynamical Black Holes

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

It has long been suggested that the Cauchy horizon of dynamical black holes is subject to a weak null singularity, under the mass inflation scenario. We study in spherical symmetry the Einstein–Maxwell–Klein–Gordon equations and while we do not directly show mass inflation, we obtain a “mass inflation/ridigity” dichotomy. More precisely, we prove assuming (sufficiently slow) decay of the charged scalar field on the event horizon, that the Cauchy horizon emanating from time-like infinity \(\mathcal {CH}_{i^+}\) can be partitioned as \(\mathcal {CH}_{i^+}= {\mathcal {D}} \cup {\mathcal {S}}\) for two (possibly empty) disjoint connected sets \({\mathcal {D}}\) and \({\mathcal {S}}\) such that

  • \({\mathcal {D}}\) (the dynamical set) is a future set on which the Hawking mass blows up (mass inflation scenario).

  • \({\mathcal {S}}\) (the static set) is a past set isometric to a Reissner–Nordström Cauchy horizon i.e. the radiation is zero on \({\mathcal {S}}\).

As a consequence of this result, we prove that the entire Cauchy horizon \(\mathcal {CH}_{i^+}\) is globally \(\underline{C^2-{ inextendible}}\), extending a previous local result established by the author. To this end, we establish a novel classification of Cauchy horizons into three types: dynamical (\({\mathcal {S}}=\emptyset \)), static (\({\mathcal {D}}=\emptyset \)) or mixed. As a side benefit, we prove that there exists a trapped neighborhood of the Cauchy horizon, thus the apparent horizon cannot cross the Cauchy horizon, which is a result of independent interest. Our main motivation is to prove the \(C^2\) Strong Cosmic Censorship Conjecture for a realistic model of spherical collapse in which charged matter emulates the repulsive role of angular momentum. In our case, this model is the Einstein–Maxwell–Klein–Gordon system on space-times with one asymptotically flat end. As a consequence of the \(C^2\)-inextendibility of the Cauchy horizon, we prove the following statements, in spherical symmetry:

  1. 1.

    Two-ended asymptotically flat space-times are \(C^2\)-future-inextendible i.e. \(C^2\) Strong Cosmic Censorship is true for Einstein–Maxwell–Klein–Gordon, assuming the decay of the scalar field on the event horizon at the expected rate.

  2. 2.

    In the one-ended case, under the same assumptions, the Cauchy horizon emanating from time-like infinity is \(C^2\)-inextendible. This result suppresses the main obstruction to \(C^2\) Strong Cosmic Censorship in spherical collapse.

The remaining obstruction in the one-ended case is associated to “locally naked” singularities emanating from the center of symmetry, a phenomenon which is also related to the Weak Cosmic Censorship Conjecture.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. In the two-ended case, no additional assumption is required. In the one-ended case, we obtain the result assuming additionally the absence of “locally naked singularity” emanating from the center of symmetry, a slightly stronger statement than the Weak Cosmic Censorship Conjecture.

  2. This statement does not prove that the metric is \(C^1\)-inextendible but does give the insight that a breakdown occurs already at the \(C^1\) level.

  3. This is consequence of the Raychaudhuri equation and the dominant energy condition, as the quantity \(\frac{4|\partial _u r|}{\Omega ^2}\) is monotonic c.f. (A.7).

  4. In fact, the space-time is \(C^2\) extendible and the Hawking mass finite if \(f_R\) decays exponentially at a sufficiently fast rate. This phenomenon explains why in the cosmological setting, mass inflation is not expected for a certain range of parameters c.f. [12].

  5. We emphasize however that this blow up was not formulated in either [42, 43] or [37]. This modern formulation is due to Luk and Oh [32].

  6. Essentially, such data decay weakly and are non-oscillating, so do not obey the asymptotics of Conjecture 1.7 (non-generic behavior).

  7. e.g. they are important for the \(C^0\)-inextendibility result of [26].

  8. In fact, the original theorem of [47] only requires \(s>\frac{1}{2}\), but we assume \(s>\frac{3}{4}\) to simplify the proofs in the present paper.

  9. Additionally, if if \(s>1\), we prove in [47] that \((M,g,\phi ,F)\) admits a continuous extension to \(\mathcal {CH}_{i^+}\).

  10. Of course, if \(\mathcal {CH}_{i^+}\) is the ingoing Cauchy horizon of the one-ended case, or \(\mathcal {CH}_{i^+_1}\) in the two-ended case, we can chose \(\tau =u\) and \(\varsigma =v\).

  11. Those limits exist as a soft consequence of the assumptions of Theorem 3.1, see [47].

  12. By this, we mean that \(\rho ^{-1}\) and \(\varpi ^{-1}\) extend continuously to 0 on \(\mathcal {CH}_{i^+}\).

  13. Otherwise, we already know that \(\mathcal {LB} \cap \{u \le u_s\} \subset {\mathcal {T}}\) (recall that \(\mathcal {LB}\) is defined as the future of a space-like trapped curve \(\gamma \), c.f. [47]).

  14. We are using, implicitly, a standard local well-posedness argument for characteristic initial data entirely inside the space-time, i.e. whose closure is disjoint from the boundary of the Penrose diagram.

  15. Precisely Lemma 10.3, in which it is assumed that the rectangle has finite volume. The finiteness of the volume is later obtained by a different argument, using a monotonicity property specific to the uncharged and massless considered in [32].

  16. Making point-wise lower bounds assumptions on the scalar field on the event horizon, and exploiting a special monotonicity property which is specific to the uncharged and massless case.

  17. As noticed in [16], the monotonicity of the Hawking mass in the uncharged and massless case allows to propagate the mass blow-up to the entire Cauchy horizon, which implies that condition (5.2) is violated everywhere on \(\mathcal {CH}_{i^+}\) for the space-times under consideration. This technique does not, however, survive when the field is massive and/or charged.

  18. This additional assumption is required to obtain a lower bound on r, which is not necessarily valid as one approaches \(\{(u_{\infty }(\mathcal {CH}_{i^+}),v=+\infty )\}\) the end-point towards which r may (or may not) approach 0, c.f. [28].

  19. In particular R is of finite space-time volume, but as we shall see, the stronger assumption \(R \subset {\mathcal {T}}\) turns out to be necessary to our method.

  20. Recall that, in retrospect, Lemma 5.8 was concerned by Cauchy horizon of mixed type: in this case, \(\phi \) and Q were necessarily bounded.

  21. The lack of integrability when \(s<1\), even if \(\phi \) is bounded, was a new difficulty in Lemma 5.8 already, compared to Lemma 10.3 of [32].

  22. This is because we require estimates, even weak ones, just to prove that \(\nu \) has a non-zero limit on \(\mathcal {CH}_{i^+}\).

  23. In particular we do not need to use the bootstraps (8.25), (8.26) or (8.27), which makes the estimates relatively easy.

  24. Strictly speaking, the result in [32] is Lemma 11.5, which presents an alternative between two possibilities, one of them of the statement of Proposition 9.1, and the other one is proven to be impossible in Lemma 11.6, using soft estimates present in [28].

  25. The proof of [32], which does not use the equations, can be exactly reproduced in our setting.

  26. Note that Ori considers Cauchy horizons of mixed type in [37].

  27. Note that while the estimates of [47] concern the scalar field case, they can be immediately transposed to the null dust case with no further work, as the Einstein equations are the same, except with fewer terms and moreover the propagation of both dust clouds \(f_R\) and \(f_L\) is trivial.

References

  1. Angelopoulos, Y., Aretakis, S., Gajic, D.: A vector field approach to almost-sharp decay for the wave equation on spherically symmetric, stationary spacetimes. Ann. PDE, volume 4, Article number: 15 (2018)

  2. Angelopoulos, Y., Aretakis, S.: Dejan Gajic Late-time asymptotics for the wave equation on spherically symmetric, stationary spacetimes. Adv. Math. 323, 529–621 (2018)

    Article  MathSciNet  Google Scholar 

  3. Burko, L.: Gaurav Khanna Universality of massive scalar field late-time tails in black-hole spacetimes. Phys. Rev. D 70, 044018 (2004)

    Article  ADS  MathSciNet  Google Scholar 

  4. Burq, N., Planchon, F., Stalker, J., Tahvildar-Zadeh, A.S.: Strichartz estimates for wave and Schroedingerequations with potentials of critical decay. Indiana Univ. Math. J. 53(6), 1667–1682 (2004)

    Article  Google Scholar 

  5. Burq, N., Planchon, F., Stalker, J., Tahvildar-Zadeh, A.S.: Strichartz estimates for the wave and Schroedinger equations with the inverse-square potential. J. Funct. Anal. 203, 519–549 (2003)

    Article  MathSciNet  Google Scholar 

  6. Choquet-Bruhat, Y.: Theoreme d’existence pour certains systemes d’equations aux derivees partielles non linaires. Acta Math. 88, 141–225 (1952)

    Article  ADS  MathSciNet  Google Scholar 

  7. Christodoulou, D.: The formation of black holes and singularities in spherically symmetric gravitational collapse. Commun. Pure Appl. Math. 44(3), 339–373 (1991)

    Article  MathSciNet  Google Scholar 

  8. Christodoulou, D.: Bounded variation solutions of the spherically symmetric Einstein-scalar field equations. Commun. Pure Appl. Math. 46, 1131–1220 (1993)

    Article  MathSciNet  Google Scholar 

  9. Christodoulou, D.: On the global initial value problem and the issue of singularities. Class. Quantum Grav. 16, A23–A35 (1999)

    Article  MathSciNet  Google Scholar 

  10. Christodoulou, D.: The instability of naked singularities in the gravitational collapse of a scalar field. Ann. Math. (2) 149(1), 183–217 (1999)

    Article  MathSciNet  Google Scholar 

  11. Christodoulou, Demetrios.: The formation of black holes in general relativity. European Mathematical Society (2009)

  12. Costa, J.L., Girão, P.M., Natário, J., Silva, J.D.: On the global uniqueness for the Einstein–Maxwell-scalar field system with a cosmological constant Part 3: mass inflation and extendibility of the solutions. Ann PDE 3, 8 (2017)

    Article  MathSciNet  Google Scholar 

  13. Dafermos, M.: Stability and instability of the Cauchy horizon for the spherically symmetric Einstein–Maxwell-scalar field equations. Ann. Math. 158, 875–928 (2003)

    Article  MathSciNet  Google Scholar 

  14. Dafermos, M.: The interior of charged black holes and the problem of uniqueness in general relativity. Commun. Pure Appl. Math. 58, 0445–0504 (2005)

    Article  MathSciNet  Google Scholar 

  15. Dafermos, M.: Spherically symmetric spacetimes with a trapped surface. Class. Quantum Grav. 22(11) (2005)

  16. Dafermos, M.: Black holes without spacelike singularities. Commun. Math. Phys. 332, 729–757 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  17. Dafermos, M., Luk, J.: The interior of dynamical vacuum black holes I: the \(C^0\)-stability of the Kerr Cauchy horizon. (2017). arXiv:1710.01772

  18. Dafermos, M., Rendall, A.: Inextendibility of expanding cosmological models with symmetry. Class. Quantum Grav. 22(23), L143–L147 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  19. Dafermos, M.: Igor Rodnianski A proof of Price’s law for the collapse of a self-gravitating scalar field. Invent. math. 162, 381–457 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  20. Donninger, R.: Joachim Krieger a vector field method on the distorted Fourier side and decay for wave equations with potentials. Mem. Am. Math. Soc. (2015)

  21. Gajic, D., Luk, J.: The interior of dynamical extremal black holes in spherical symmetry. Pure Appl. Anal. 1(2), 263–326 (2019)

    Article  MathSciNet  Google Scholar 

  22. Hiscock, W.: Evolution of the interior of a charged black hole. Phys. Lett. A 83(3), 110–112 (1981)

    Article  ADS  MathSciNet  Google Scholar 

  23. Hod, S., Piran, T.: Late-time evolution of charged gravitational collapse and decay of charged scalar hair-II. Phys. Rev. D 58, 024018 (1998)

    Article  ADS  MathSciNet  Google Scholar 

  24. Hod, S., Piran, T.: Mass-inflation in dynamical gravitational collapse of a charged scalar-field. Phys. Rev. Lett. 81, 1554–1557 (1998)

    Article  ADS  Google Scholar 

  25. Kehle, C.: Yakov Shlapentokh-Rothman a scattering theory for linear waves on the interior of Reissner–Nordström black holes. Ann. Henri Poincaré 20 (2019)

  26. Kehle, C., Van de Moortel, M.: Continuous extendibility and non-linear scattering on the interior of dynamical black holes in the presence of matter, in preparation

  27. Kehle, C., Van de Moortel, M.: The null contraction singularity at the Cauchy horizon of dynamical black holes, in preparation

  28. Kommemi, J.: The global structure of spherically symmetric charged scalar field spacetimes. Commun. Math. Phys. 323, 35 (2013). https://doi.org/10.1007/s00220-013-1759-1

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Koyama, H., Tomimatsu, A.: Asymptotic power-law tails of massive scalar fields in Reissner–Nordström background. Phys. Rev. D 63, 064032 (2001)

    Article  ADS  Google Scholar 

  30. Konoplya, R., Zhidenko, A.: A massive charged scalar field in the Kerr–Newman background I: quasi-normal modes, late-time tails and stability. Phys. Rev. D 88, 024054 (2013)

    Article  ADS  Google Scholar 

  31. Luk, J.: Weak null singularities in general relativity. J. AMS 31, 1–63 (2018)

    MathSciNet  MATH  Google Scholar 

  32. Luk, J.: Sung-Jin Oh Strong Cosmic Censorship in Spherical Symmetry for two-ended Asymptotically Flat Initial Data I. The Interior of the Black Hole Region, Preprint (2017)

  33. Luk, J.: Sung-Jin Oh Strong Cosmic Censorship in Spherical Symmetry for two-ended Asymptotically Flat Initial Data II. The Exterior of the Black Hole Region, Preprint (2017)

  34. McNamara, J.: Instability of black hole inner horizons. Proc. R. Soc. Lon. A 358, 499–517 (1978)

    Article  ADS  MathSciNet  Google Scholar 

  35. Moschidis, G.: The Einstein–Null dust system in spherical symmetry with an inner mirror: structure of the maximal development and Cauchy stability., Preprint, arXiv:1704.08685, (2017)

  36. Oren, Y., Piran, T.: On the collapse of charged scalar fields. Phys. Rev. D 68, 044013 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  37. Ori, A.: Inner structure of a charged black hole: an exact mass-inflation solution. Phys. Rev. Lett. 67(7), 789–792 (1991)

    Article  ADS  MathSciNet  Google Scholar 

  38. Penrose, R.: Structure of space-time. In: Penrose, R., DeWitt, C.M. (eds.) Battelle Rencontres, p. 22. W. A. Benjamin, New York (1968)

    Google Scholar 

  39. Penrose, R.: Gravitational collapse: the role of general relativity. Riv. Nuovo Cim. 1, 252–276 (1969)

    ADS  Google Scholar 

  40. Planchon, F., Stalker, J., Tahvildar–Zadeh, A.S.: Dispersive estimate for the wave equation with the inverse-square potential. Discrete Contin. Dyn. Syst. 9(6), 1387–1400 (2003)

    Article  MathSciNet  Google Scholar 

  41. Planchon, F., Stalker, J., Tahvildar-Zadeh, A.S.: \(L^p\) estimates for the wave equation with the inverse-square potential. Discrete Contin. Dyn. Syst. 9(2), 427–442 (2003)

    MathSciNet  MATH  Google Scholar 

  42. Poisson, E., Israel, W.: Internal structure of black holes. Phys. Rev. D 63, 1663–1666 (1989)

    ADS  Google Scholar 

  43. Poisson, E., Israel, W.: Inner-horizon instability and mass inflation in black holes. Phys. Rev. Lett. 67(7), 789–792 (1991)

    Article  MathSciNet  Google Scholar 

  44. Price, R.: Nonspherical perturbations of relativistic gravitational collapse. I. Scalar and gravitational perturbations. Phys. Rev. D (3) 5, 2419–2438 (1972)

    Article  ADS  MathSciNet  Google Scholar 

  45. Sbierski, J.: Characterisation of the energy of Gaussian beams on Lorentzian manifolds: with applicationsto black hole spacetimes. Anal. PDE 8(6), 1379–1420 (2015)

    Article  MathSciNet  Google Scholar 

  46. Sbierski, J.: The \(C^0\)-inextendibility of the Schwarzschild spacetime and the spacelike diameter in Lorentzian geometry. J. Differ. Geom. 108(2), 319–378 (2018)

    Article  MathSciNet  Google Scholar 

  47. Van de Moortel, M.: Stability and instability of the sub-extremal Reissner–Nordström black hole interior for the Einstein–Maxwell–Klein–Gordon equations in spherical symmetry. Commun. Math. Phys. 360(1), 103–168 (2018)

    Article  ADS  Google Scholar 

  48. Van de Moortel, M.: Decay of weakly charged solutions for the spherically symmetric Maxwell—Charged-Scalar-Field equations on a Reissner–Nordström exterior space-time, preprint, 2018. arXiv:1804.04297

  49. Van de Moortel , M.: Charged scalar fields on Black Hole space-times, PhD Thesis, University of Cambridge (2018)

  50. Van de Moortel, M.: The breakdown of weak null singularities inside black holes, preprint. arXiv:1912.10890 (2019)

Download references

Acknowledgements

I am very grateful to Jonathan Luk for suggesting this problem and for fruitful discussions. I would like to thank Sung-Jin Oh for this interest in this problem, and for insightful discussions. I am grateful to Mihalis Dafermos and Jonathan Luk for useful comments on the manuscript. Most of this work was completed while I was a Ph.D. student at the University of Cambridge, and a visiting graduate student at Stanford University, whose hospitality I gratefully acknowledge. Finally, I would like to thanks two anonymous reviewers for helpful suggestions which helped clarify some aspects of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Maxime Van de Moortel.

Additional information

Communicated by P. Chrusciel

Publisher's Note

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

Appendix A Construction of a Cauchy Horizon of Mixed Type for the Einstein–Null–Dust Model

Appendix A Construction of a Cauchy Horizon of Mixed Type for the Einstein–Null–Dust Model

In this appendix, we construct an example of a Cauchy horizon of mixed type, following Definition 3.3. In the second part of our development, we prove the blow up of the Hawking mass for mixed and dynamical type Cauchy horizons in the Einstein–Maxwell-null-dust model, following Poisson and Israel [43] andFootnote 26 Ori [37], the first instances of the mass inflation scenario in the literature. Notice that their model is very elementary, as the dust clouds are simply transported linearly in the null directions, and only interact indirectly, via the metric: thus, the “scattering theory” is trivial. Additionally, such a model does not allow for one-ended regular solutions, unlike the charged/massive scalar field model. In this appendix we sketch a mathematical proof of mass inflation, using the methods of [47]; this is also an opportunity to illustrate the terminology of the present paper in the simpler setting of dust, where little analysis is required and monotonicity suffices. The system (1.7), (1.8), (1.9), (1.10), (1.11), (1.12), (1.13) can be expressed in spherical symmetry as

$$\begin{aligned} \partial _{u}\partial _{v} r= & {} \partial _u \lambda =\partial _v \nu = -\frac{\Omega ^2}{2r} \cdot (\frac{\varpi }{r}-\frac{e^2}{r^2}) =\frac{\lambda \nu }{r}-\frac{\Omega ^2}{4r} \cdot (1- \frac{e^2}{r^ 2}) \end{aligned}$$
(A.1)
$$\begin{aligned} \partial _{u}\partial _{v} \log (\Omega ^2)= & {} \frac{\Omega ^2}{2r^2} \cdot ( \frac{2\varpi }{r}-\frac{3e^2}{r^2})= \frac{2 \nu \lambda }{r^{2}}+ \frac{ \Omega ^{2}}{2r^{2}} \cdot (1- \frac{ 2 e^{2}}{r^{2}}),\end{aligned}$$
(A.2)
$$\begin{aligned} \partial _{u} \varpi= & {} \frac{r^2}{2} \kappa ^{-1}| f_R|^{2}, \end{aligned}$$
(A.3)
$$\begin{aligned} \partial _{v} \varpi= & {} \frac{r^2}{2} \iota ^{-1}| f_L|^{2}, \end{aligned}$$
(A.4)
$$\begin{aligned} \partial _v (r f_R)= & {} 0, \end{aligned}$$
(A.5)
$$\begin{aligned} \partial _u (r f_L)= & {} 0, \end{aligned}$$
(A.6)
$$\begin{aligned} \partial _{u}(\frac{-4\nu }{\Omega ^{2}})= & {} \partial _{u}(\kappa ^{-1})=\frac{4r}{\Omega ^{2}}| f_R|^{2}, \end{aligned}$$
(A.7)
$$\begin{aligned} \partial _{v}(\frac{-4\lambda }{\Omega ^{2}})= & {} \partial _v(\iota ^{-1})=\frac{4r}{\Omega ^{2}}|f_L|^{2}, \end{aligned}$$
(A.8)
$$\begin{aligned} F_{\mu \nu }= & {} \frac{e^2 \Omega ^2}{ 2r^2} du \wedge dv. \end{aligned}$$
(A.9)

Proposition A.1

We consider the maximal development of smooth, spherically symmetric and admissible (in the sense of Definition 2.6) two-ended initial data \((M={\mathcal {Q}}^+ \times _r {\mathcal {S}}^2,g_{\mu \nu }, \phi ,F_{\mu \nu })\) satisfying the Einstein–Maxwell-null-dust system. Suppose that Assumptions 12 and 3 of Theorem 3.1 are satisfied, and we also work under the same (Uv) gauge choice. We also choose \(U_{max} \in {\mathbb {R}}\) such that \( [0,U_{max}]\times \{v=+\infty \} \subset \mathcal {CH}_{i^+}\) in the Penrose diagram \({\mathcal {Q}}^+\). We consider data \(f_L^0\) on the event horizon \({\mathcal {H}}^+= \{0\} \times [v_0,+\infty )\) and \(f_R^0\) on an ingoing cone \(\underline{C}_{v_0}= [0,U_{max}]\times \{v_0\}\); we assume that both \(f_R^0\) and \(f_L^0\) are smooth in the (Uv) coordinate system. We also make assumptions analogous to Assumptions 4 and 5 of Theorem 3.1 i.e. for some \(s>\frac{1}{2}\) and some \(C>0\):

$$\begin{aligned} \sup _{v \ge v_0 }|f_L^0|_{|{\mathcal {H}}^+}(v) \cdot v^{s} \le C \cdot , \\ \sup _{0 \le U \le U_{max}}|f_R^0| (U, v_0) \le C. \end{aligned}$$

Then by continuity of \(f^0_R\) on \([0,U_{max})\), there are three possibilities:

  1. (a)

    For all \(0<U_s<U_{max}\), there exists \(0 \le U \le U_s\) such that \(f^0_R(U) \ne 0\). We call this the dynamical case.

  2. (b)

    For all \(U\in [0,U_{max})\), \(f^0_R(U) = 0\). We call this the static case.

  3. (c)

    There exists \(U_T \in (0,U_{max})\) such that for all \(U\in [0,U_{T}]\), \(f^0_R(U) = 0\), and such that for every \(\epsilon >0\), there exists \(U\in (U_{T},U_{T}+\epsilon ]\) such that \(f^0_R(U) \ne 0\). We call this the mixed case.

Then we have, in the three different cases:

  1. (a)

    In the dynamical case, r extends to \(\mathcal {CH}_{i^+}\) as function \(r_{CH}\) which is strictly decreasing on \([0,U_{max})\).

  2. (b)

    In the static case, r extends to \(\mathcal {CH}_{i^+}\) as a constant \(r_{-}>0\).

  3. (c)

    In the mixed case, r extends to \(\mathcal {CH}_{i^+}\) as a function \(r_{CH}\), which is constant on \([0,U_T]\), and strictly decreasing on \((U_T,U_{max} )\).

If we additionally assume that \(f_L^0(v)=C \cdot v^{-s}+o(v^{-s})\), for some \(s>\frac{1}{2}\), where v is defined by gauge (3.1), then we have

  1. (a)

    In the dynamical case, the Hawking mass \(\rho \) blows up on \(\mathcal {CH}_{i^+}\cap [0,U_{max}] \).

  2. (b)

    In the static case, the Hawking mass \(\rho \) is constant on \(\mathcal {CH}_{i^+}\cap [0,U_{max}] \).

  3. (c)

    In the mixed case, the Hawking mass \(\rho \) is constant on \(\mathcal {CH}_{i^+}\cap [0,U_{T}] \) and blows up on \(\mathcal {CH}_{i^+}\cap (U_T,U_{max}]\).

Remark A.1

Note that \(\mathcal {CH}_{i^+}\ne \emptyset \), applying (an easier version of) the argument of [47], which proves the stability of the Reissner–Nordström Cauchy horizon for the Einstein–Maxwell–Klein–Gordon model in spherical symmetry.

Proof

We first prove that the possibilities reduce to the three cases a, b, c. Define the set \(E:=\{U\in [0,U_{max}), f_R^0(U)\ne 0\}\). If \(E=\emptyset \) then we are in case b. If \(E \ne \emptyset \), define \(U_T=\inf E\). There are two possibilities: either \(U_T=0\) or \(U_T>0\). If \(U_T=0\) then we are in case  a by definition of E. If \(U_T>0\) then for all \(0 \le U < U_T\), \(f_R^0(U)= 0\). By continuity of \(f_R^0\) we have also \(f_R^0(U_T)= 0\). Again by definition of E, we are therefore in case c.

We work in the gauge \(\kappa _{|{\mathcal {H}}^+} \equiv 1\) and we denote \(\varphi _R(u)= r(u,v) f_R(u,v)\), \(\varphi _L(v)= r(u,v) f_L(u,v)\) by (A.5), (A.6).

By the stability estimatesFootnote 27 of [47] (see Sect. 4), one can show that for some sub-extremal parameters \(0<|e|<M\),

$$\begin{aligned}&|\varpi (U_{\gamma }(v),v)-M|+ |r(U_{\gamma }(v),v)-r_-(M,e)|\\&\quad + |\partial _v \log (\Omega ^2)(U_{\gamma }(v),v)-2K_-(M,e)| + v \cdot |\partial _v r| (U_{\gamma }(v),v)\lesssim v^{1-2s} \end{aligned}$$

holds on some space-like curve \(\gamma \) terminating at \(i^+\), for v large enough. We will then establish estimates for all \(U \in [0,U_{max}]\): combining (A.1) and (A.2), we get \( \partial _U \partial _v \log (r \Omega ^2) = \frac{ \Omega ^{2}}{2r^{2}} \cdot (1- \frac{ 3 e^{2}}{r^{2}})\). In view of \(|e| > r_-(M,e)\) and the monotonicity of r (which implies that \(r(U,v)\le r(U_{\gamma }(v),v) \) in our region of interest), we have \( \partial _U \partial _v \log (r \Omega ^2)(U,v) <0\) for all \(U \in [0,U_{max}]\). Integrating in U from \(\gamma \), assuming v large enough (in particular \(v\ge v_{\gamma }(U_s)\)), we have, for all \(U \in [0,U_{max}]\):

$$\begin{aligned} \partial _v\log (r\Omega ^2)(U,v) \le \partial _v\log (r\Omega ^2)(U_{\gamma }(v),v) \le 2K_-(M,e) - C \cdot v^{1-2s} \le K_-(M,e)<0, \end{aligned}$$

where the last inequality follows from v being large enough. This proves that for all \(U \in [0,U_{max}]\)

$$\begin{aligned} \Omega ^2(U,v) \lesssim e^{K_-(M,e) \cdot v }. \end{aligned}$$
(A.10)

Using (A.10) with \(\partial _U \partial _v \log (r \Omega ^2) = \frac{ \Omega ^{2}}{2r^{2}} \cdot (1- \frac{ 3 e^{2}}{r^{2}})\) and the fact that \( C(M,e)^{-1} \lesssim |1- \frac{ 3 e^{2}}{r^{2}}| \lesssim C(M,e)\) by monotonicity of r and the fact that r is lower bounded, we get an improved estimate, still for all \(U \in [0,U_{max}]\):

$$\begin{aligned} |\partial _v \log (r \Omega ^2)(U,v) -2K_-(M,e) | \lesssim v^{1-2s}, e^{3K_-(M,e) \cdot v } \lesssim \Omega ^2(U,v) \lesssim e^{K_-(M,e) \cdot v }.\nonumber \\ \end{aligned}$$
(A.11)

Now, we can write (A.1), since \(\nu \le 0\), as \(\partial _v( r|\nu |) =\Omega ^2 \cdot (1- \frac{e^2}{r^2})\). Since r is decreasing in U, \((1- \frac{e^2}{r^2(U,v)})<0\) for all \(U \in [0,U_{max}]\), as \(r<r_-(M,e)<|e|\). Hence \(v \rightarrow r\nu (U,v)\) is strictly decreasing for all fixed U, thus has a limit \(r\nu _{\mathcal {CH}_{i^+}}(U) \in {\mathbb {R}}_-\). Moreover, we establish, using (A.1), the following estimate for all \(U \in [0,U_{max}]\):

$$\begin{aligned} | r\nu (U,v) - r\nu _{\mathcal {CH}_{i^+}}(U) | \lesssim \Omega ^2(U,v) . \end{aligned}$$
(A.12)

Now, writing (A.1) as \(\partial _U(- r\lambda ) =\Omega ^2 \cdot (1- \frac{e^2}{r^2})\), we get decay for \(\lambda \) so we can extend r to \(\mathcal {CH}_{i^+}\) into a function \(r_{\mathcal {CH}_{i^+}} >0\). Therefore, for any fixed \(U<U_{max}\), \(v \rightarrow \nu (U,v)\) has a limit as \(v\rightarrow +\infty \) that we consistently denote \(\nu _{\mathcal {CH}_{i^+}}(U)\).

Since r is lower bounded in this region, we have \(\partial _U (\kappa ^{-1})(U,v) \gtrsim C \cdot \varphi _R^2(U) \cdot e^{|K_-|(M,e) v}\) hence, integrating from \(\gamma \):

$$\begin{aligned} \kappa ^{-1}(U,v)\gtrsim & {} e^{|K_-|(M,e) v} \cdot \int _{U_{\gamma (v)}}^U \varphi _R^2(U')dU'\nonumber \\= & {} e^{|K_-|(M,e) v} \cdot \int _{U_{\gamma (v)}}^U r^2(U',v_0) \cdot (f_R^0(U'))^2dU' \nonumber \\\gtrsim & {} e^{|K_-|(M,e) v} \cdot \int _{U_{\gamma (v)}}^U (f_R^0(U'))^2dU'. \end{aligned}$$
(A.13)

Then, there are our three possibilities, starting with the easiest:

  1. 1.

    Dynamical case: then for all \(U\in [0,U_{max})\), \(\int _{0}^U(f^0_R)^2(U')dU' > 0\). Indeed by (a), for all \(U\in [0,U_{max})\), there exists \(U'<U\) with \((f^0_R(U'))^2 > 0\), hence by continuity there exists \(\epsilon >0\) such that \((f^0_R(U'))^2 > 0\) on \([U',U'+\epsilon ]\), which implies \(\int _{0}^U(f^0_R)^2(U')dU' > 0\) . In view of the limit \(U_{\gamma }(v) \rightarrow 0\) as \(v\rightarrow +\infty \) (recall that \(\gamma \) terminates at \(i^+\)) then \(\lim _{v \rightarrow +\infty } \int _{U_{\gamma (v)}}^U (f_R^0(U'))^2dU' >0\) for all \(U\in [0,U_{max})\), hence \(\kappa ^{-1}(U,v) \rightarrow +\infty \). If \( \nu _{CH}(U)=0\) for some \(0<U < U_{max}\), then (A.12) is contradicted (after dividing both sides of (A.12) by \(r \Omega ^2\)). Hence \(\nu _{\mathcal {CH}_{i^+}}(U)<0\) for all \(U\in [0,U_{max}]\) thus \(r_{\mathcal {CH}_{i^+}}\) is strictly decreasing on \([0,U_{max})\) as desired . Moreover \( \kappa ^{-1}(U,v) \sim \frac{4|\nu _l|(U)}{\Omega ^2(U,v)}\) by (A.12). If additionally we have a lower bound \(f_L^0(v)=C \cdot v^{-s}+o(v^{-s})\) then by the stability estimates of [47], \(-\lambda \sim C \cdot v^{-2s}\), at least for U small enough and in the future of the curve \(\gamma \). Since \(2\rho -r= \kappa ^{-1} \cdot (-\lambda )\), \(\rho (U,v) \rightarrow +\infty \) as \( v \rightarrow +\infty \), for small U. A forciciori, \(\varpi (U,v) \rightarrow +\infty \) for small U, and by the monotonicity of (A.3), \(\varpi (U,v) \rightarrow +\infty \) for all \(U\in [0,U_{max})\). Hence, for all \(U\in [0,U_{max})\), \(\rho (U,v) \rightarrow +\infty \) as \( v \rightarrow +\infty \).

  2. 2.

    Static case: for all \(U\in [0,U_{max})\), \(f^0_R(U) = 0\). Then \(\kappa (U,v) = 1\) as \(\partial _U (\kappa ^{-1})=0\). By (A.12), it means that \(\nu _{\mathcal {CH}_{i^+}}(U)=0\) for all \(U\in [0,U_{max})\) hence \(r_{\mathcal {CH}_{i^+}}\) is constant and by (A.3) \(\varpi _{\mathcal {CH}_{i^+}}\) is constant. Their values are \( r_{\mathcal {CH}_{i^+}} \equiv r_-(M,e)>0\) and \( \varpi _{\mathcal {CH}_{i^+}} \equiv M >0\) by the stability estimates of [47]. This implies that \(\rho \) is constant on \(\mathcal {CH}_{i^+}\): \(\rho (U,v) \rightarrow M-\frac{e^2}{r_-(M,e)}\) as \(v \rightarrow +\infty \) for all \(U\in [0,U_{max})\).

  3. 3.

    Mixed case: \(\int _{0}^{U_T}(f^0_R)^2(U')dU' = 0\) but \(\int _{U_T}^{U}(f^0_R)^2(U')dU' > 0\) for all \(U_T<U<U_{max}\). Indeed, the above statement follows from the same logic as for the dynamical case, see earlier discussion. Then similarly, \(r_{\mathcal {CH}_{i^+}}(U)= r_-(M,e)\), \(\varpi _{\mathcal {CH}_{i^+}}(U)= M\), and \(\rho _{\mathcal {CH}_{i^+}}(U)= M-\frac{e^2}{r_-(M,e)}\) for all \(U \in [0,U_T]\). Yet, \(\kappa ^{-1}(U,v) \rightarrow +\infty \) for all \(U \in (U_T,U_{max})\), as \(\int _{0}^U (f_R^0(U'))^2dU' >0\). Thus \(\nu _{\mathcal {CH}_{i^+}}(U)<0\) for all \(U \in (U_T,U_{max})\) and \(r_{\mathcal {CH}_{i^+}}\) is strictly decreasing on \((U_T,U_{max})\). If additionally we have a lower bound \(f_L^0(v)=C \cdot v^{-s}+o(v^{-s})\), then we have the blow up of \(\rho \), as in the dynamical case.

\(\square \)

Remark A.2

Notice that the proof was considerably easier for the uncharged dust than for a charged scalar field, as we used in this section some special monotonicity properties that are not exploitable in the charged case. In contrast, in the earlier sections, we proved estimates that were harder to obtain but also more robust as they do not rely on monotonicity.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Van de Moortel, M. Mass Inflation and the \(C^2\)-inextendibility of Spherically Symmetric Charged Scalar Field Dynamical Black Holes. Commun. Math. Phys. 382, 1263–1341 (2021). https://doi.org/10.1007/s00220-020-03923-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-020-03923-w

Navigation