Skip to main content
Log in

The positive mass theorem for multiple rotating charged black holes

  • Published:
Calculus of Variations and Partial Differential Equations Aims and scope Submit manuscript

Abstract

In this paper a lower bound for the ADM mass is given in terms of the angular momenta and charges of black holes present in axisymmetric initial data sets for the Einstein–Maxwell equations. This generalizes the mass-angular momentum-charge inequality obtained by Chrusciel and Costa to the case of multiple black holes. We also weaken the hypotheses used in the proof of this result for single black holes, and establish the associated rigidity statement.

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

Similar content being viewed by others

Notes

  1. The notation \(f=o_{l}(r^{-a})\) asserts that \(\lim _{r\rightarrow \infty }r^{a+j}\partial _{r}^{j}f=0\) for all \(j\le l\), and \(f=O_{l}(r^{-a})\) asserts that \(r^{a+j}|\partial _{r}^{j}f|\le C\) for all \(j\le l\). The assumption \(l\ge 5\) is needed for the results in [6, 17].

References

  1. Brill, D.: On the positive definite mass of the Bondi-Weber-Wheeler time-symmetric gravitational waves. Ann. Phys., 7, 466483 (1959)

  2. Cha, Y., Khuri, M.: Deformations of axially symmetric initial data and the mass-angular momentum inequality. Ann. Henri Poincaré 16(3), 841–896 (2015). arXiv:1401.3384

  3. Cha, Y., Khuri, M.: Deformations of charged axially symmetric initial data and the mass-angular momentum-charge inequality. Ann. Henri Poincaré, \(\mathbf{16}\)(12), 2881–2918 (2015). arXiv:1407.3621

  4. Choquet-Bruhat, Y.: General relativity and the Einstein equations, Oxford University Press, Oxford (2009)

  5. Chruściel, P.: On completeness of orbits of Killing vector fields. Classical Quantum Gravity, 10(10), 2091–2101 (1993). arXiv:gr-qc/9304029

  6. Chruściel, P.: Mass and angular-momentum inequalities for axi-symmetric initial data sets. I. Positivity of Mass. Ann. Phys. 323, 2566–2590 (2008). arXiv:0710.3680

  7. Chruściel, P., Costa, J.: Mass, angular-momentum and charge inequalities for axisymmetric initial data. Class. Quantum Gravity 26(23), 235013 (2009). arXiv:0909.5625

  8. Chruściel, P., Galloway, G., Pollack, D.: Mathematical general relativity: a sampler. Bull. Am. Math. Soc. (N.S.) 47(4), 567–638 (2010). arXiv:1004.1016

  9. Chruściel, P., Li, Y., Weinstein, G.: Mass and angular-momentum inequalities for axi-symmetric initial data sets. II. Angular Momentum. Ann. Phys. 323, 2591–2613 (2008). arXiv:0712.4064

  10. Chrusciel, P., Reall, H., Tod, P.: On Israel-Wilson-Perjes black holes. Class. Quantum Gravity 23, 2519–2540 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  11. Costa, J.: Proof of a Dain inequality with charge. J. Phys. A 43(28), 285202 (2010). arXiv:0912.0838

  12. Dain, S.: Proof of the angular momentum-mass inequality for axisymmetric black hole. J. Differ. Geom. 79, 33–67 (2008). arXiv:gr-qc/0606105

  13. Dain, S.: Geometric inequalities for axially symmetric black holes. Class. Quantum Gravity 29(7), 073001 (2012). arXiv:1111.3615

  14. Dain, S., Khuri, M., Weinstein, G., Yamada, S.: Lower bounds for the area of black holes in terms of mass, charge, and angular momentum. Phys. Rev. D 88, 024048 (2013). arXiv:1306.4739

  15. Gibbons, G., Hawking, S., Horowitz, G., Perry, M.: Positive mass theorem for black holes. Commun. Math. Phys. 88, 295–308 (1983)

    Article  MathSciNet  Google Scholar 

  16. Huang, L.-H., Schoen R., Wang, M.-T.: Specifying angular momentum and center of mass for vacuum initial data sets. Commun. Math. Phys. 306(3), 785803 (2011). arXiv:1008.4996

  17. Khuri, M., Sokolowsky, B.: Existence of Brill coordinates for initial data with asymptotically cylindrical ends and applications (2016) (in preparation)

  18. Khuri, M., Weinstein, G.: Rigidity in the positive mass theorem with charge. J. Math. Phys. \(\mathbf{54}\), 092501 (2013). arXiv:1307.5499

  19. Mars, M.: Present status of the Penrose inequality. Class. Quantum Gravity, 26(19), 193001 (2009). arXiv:0906.5566

  20. Nguyen, L.: Singular harmonic maps and applications to general relativity. Comm. Math. Phys. 301(2), 411441 (2011)

  21. Penrose, R.: Naked singularities. Ann. New York Acad. Sci. 224, 125–134 (1973)

    Article  MATH  Google Scholar 

  22. Schoen, R., Zhou, X.: Convexity of reduced energy and mass angular momentum inequalities. Ann. Henri Poincaré 14, 1747–1773 (2013). arXiv:1209.0019

  23. Weinstein, G.: On the force between rotating co-axial black holes. Trans. Am. Math. Soc. 343, 899–906 (1994)

    MathSciNet  MATH  Google Scholar 

  24. Weinstein, G.: N-black hole stationary and axially symmetric solutions of the Einstein/Maxwell equations. Comm. Partial Differ. Equ. 21(9–10), 1389–1430 (1996). arXiv:gr-qc/9412036

  25. Weinstein, G.: Harmonic maps with prescribed singularities into Hadamard manifolds. Math. Res. Lett. 3(6), 835–844 (1996)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marcus Khuri.

Additional information

Communicated by A. Neves.

M. Khuri acknowledges the support of NSF Grant DMS-1308753.

Appendices

Revisiting the heuristic arguments

The heuristic physical arguments which motivate (1.1) go back to Penrose’s original derivation of the Penrose inequality [21]. Typically in such arguments, it is assumed that the end state of gravitational collapse is a single Kerr–Newman black hole. However, a more appropriate assumption for the end state is a finite number of mutually distant Kerr–Newman black holes moving apart with asymptotically constant velocity. This should be the result, if for instance, two distant black holes were initially moving away from each other sufficiently fast. We will now describe the heuristic arguments for the mass-angular momentum-charge inequality in this setting. It appears that this has not been previously considered in the literature.

Let \(m_{i}\), \(\mathcal {J}_i\), \(q_{i}\) denote the ADM masses, angular momenta, and total charges of the end state black holes. Then the total (ADM) mass, angular momentum, and charge of the end state is \(m=\sum m_i\), \(\mathcal {J}=\sum \mathcal {J}_{i}\), \(q=\sum q_i\). In a Kerr–Newman black hole these quantities satisfy the Eq. [14]

$$\begin{aligned} m_{i}^{2}=\frac{A_{i}}{16\pi }+\frac{q_{i}^2}{2}+\frac{\pi (q_{i}^{4}+4\mathcal {J}_{i}^{2})}{A_{i}}, \end{aligned}$$
(5.1)

where \(A_{i}\) denotes horizon area. Moreover, as a function of \(A_{i}\) (keeping \(\mathcal {J}_{i}\) and \(q_{i}\) fixed), the right-hand side is nondecreasing precisely when

$$\begin{aligned} A_{i}\ge 4\pi \sqrt{q_{i}^{4}+4\mathcal {J}_{i}^{2}}, \end{aligned}$$
(5.2)

and this inequality is always satisfied with equality only for extreme black holes. Thus, computing the minimum value of the right-hand side of (5.1) yields

$$\begin{aligned} m_{i}^2\ge \frac{q_{i}^{2}+ \sqrt{q_{i}^4 + 4\mathcal {J}_{i}^2}}{2}, \end{aligned}$$
(5.3)

with equality only for extreme black holes. Let \(m_{0}\), \(\mathcal {J}_{0}\), \(q_{0}\) denote the ADM mass, angular momentum, and total charge of an initial state. Under appropriate hypotheses, such as axisymmetry and the existence of a twist potential, angular momentum is conserved \(\mathcal {J}_{0}=\mathcal {J}=\sum \mathcal {J}_{i}\). Moreover, by assuming that no charged matter is present, the total charge is conserved \(q_0=q=\sum q_i\), and since gravitational waves may only carry away positive energy \(m_0 \ge m = \sum m_i\).

Lemma 5.1

Let \(a_i, b_i\in \mathbb {R}\) and let \(a=\sum a_i\), \(b=\sum b_i\). Then

$$\begin{aligned} \left( a^4+b^2\right) ^{1/4} \le \sum \left( a_i^4 + b_i^2\right) ^{1/4}. \end{aligned}$$
(5.4)

Proof

Let \(c_i=|b_i|^{1/2}\) and \(c=\sum c_i\), then

$$\begin{aligned} |b|^{1/2} \le \left( \sum |b_i|\right) ^{1/2}=\left( \sum c_i^2\right) ^{1/2} \le \sum c_i = c. \end{aligned}$$
(5.5)

Hence \(b^2\le c^4\). We conclude that

$$\begin{aligned} \left( a^4 + b^2\right) ^{1/4} \le \left( a^4+c^4\right) ^{1/4} \le \sum \left( a_i^4+c_i^4\right) ^{1/4} = \sum \left( a_i^4 + b_i^2\right) ^{1/4}. \end{aligned}$$
(5.6)

\(\square \)

Lemma 5.2

Let \(a_i, b_i\in \mathbb {R}\) and let \(a=\sum a_i\), \(b=\sum b_i\). Then

$$\begin{aligned} \sqrt{a^2 + \sqrt{a^4+b^2}} \le \sum \sqrt{a_i^2 + \sqrt{a_i^4+b_i^2}}. \end{aligned}$$
(5.7)

Proof

By Lemma 5.1

$$\begin{aligned} \left( a^4+b^2\right) ^{1/2} \le \left( \sum \left( a_i^4+b_i^2\right) ^{1/4} \right) ^2. \end{aligned}$$
(5.8)

Thus, it follows that

$$\begin{aligned} \sqrt{a^2 + \sqrt{a^4+b^2}} \le \sqrt{\left( \sum a_i\right) ^2 +\left( \sum \left( a_i^4+b_i^2\right) ^{1/4} \right) ^2} \le \sum \sqrt{a_i^2 + \sqrt{a_i^4+b_i^2}} \end{aligned}$$
(5.9)

\(\square \)

Now, let \(a_i=q_i\), and \(b_i=2\mathcal {J}_i\), then we get

$$\begin{aligned} \sqrt{2} m = \sqrt{2} \sum m_i \ge \sum \sqrt{q_i^2 + \sqrt{q_i^4+4\mathcal {J}_i^2}} \ge \sqrt{q^2 + \sqrt{q^4+4\mathcal {J}^2}}. \end{aligned}$$
(5.10)

Squaring both sides yields the desired result (1.1). We conclude that the heuristic arguments are sufficiently robust to support the mass-angular momentum-charge inequality, even for spacetimes with multiple black holes moving apart from one another at high velocities.

The extreme Kerr–Newman and Majumdar–Papapetrou harmonic maps

First we record formulas for the extreme Kerr–Newman harmonic map. Recall that in Boyer-Lindquist coordinates the Kerr–Newmann metric takes the form

$$\begin{aligned} -\frac{\Delta - a^{2}\sin ^{2}\theta }{\Sigma } dt^{2}+ & {} \frac{2a\sin ^{2}\theta }{\Sigma }\left( \widetilde{r}^{2}+a^{2}-\Delta \right) dtd\phi \nonumber \\+ & {} \frac{(\widetilde{r}^{2}+a^{2})^{2} - \Delta a^{2}\sin ^{2}\theta }{\Sigma } \sin ^{2}\theta d\phi ^{2} + \frac{\Sigma }{\Delta } d\widetilde{r}^{2} + \Sigma d\theta ^{2} \end{aligned}$$
(6.1)

where

$$\begin{aligned} \Delta = \widetilde{r}^{2} + a^{2} +q^{2} -2m\widetilde{r}, \quad \Sigma = \widetilde{r}^{2} + a^{2}\cos ^{2}\theta , \end{aligned}$$
(6.2)

and the electromagnetic 4-potential is given by

$$\begin{aligned} \mathbf {A} = -\frac{q_{e}\widetilde{r}}{\Sigma } \left( dt+ a\sin ^{2}\theta d\phi \right) - \frac{q_{b}\cos \theta }{\Sigma } \left( a dt + (\widetilde{r}^{2}+a^{2}) d \phi \right) , \end{aligned}$$
(6.3)

The event horizon is located at the larger of the two solutions to the quadratic equation \(\Delta =0\), namely \(\widetilde{r}_{+}=m+\sqrt{m^{2}-a^{2}-q^{2}}\), where the angular momentum is given by \(\mathcal {J} = ma\). For \(\widetilde{r}>\widetilde{r}_{+}\) it holds that \(\Delta >0\), so that a new radial coordinate may be defined by

$$\begin{aligned} r=\frac{1}{2}(\widetilde{r}-m+\sqrt{\Delta }), \end{aligned}$$
(6.4)

or rather

$$\begin{aligned}&\widetilde{r} = r + m + \frac{m^{2}-a^{2}-q^{2}}{4r}, \quad m^{2} \ne a^{2}+q^{2} \nonumber \\&\widetilde{r} = r + m, \quad m^{2}=a^{2}+q^{2}. \end{aligned}$$
(6.5)

Note that the new coordinate is defined for \(r>0\), and a critical point for the right-hand side of (6.5) (\(m^{2}\ne a^{2} + q^{2}\)) occurs at the horizon, so that two isometric copies of the outer region are encoded on this interval. The coordinates \((r,\theta ,\phi )\) then form a (polar) Brill coordinate system, which is related to the (cylindrical) Brill coordinates via the usual transformation \(\rho =r\sin \theta \), \(z=r\cos \theta \). Finally, the harmonic map \((u_\mathrm{KN},v_\mathrm{KN},\chi _\mathrm{KN},\psi _\mathrm{KN}):\mathbb {R}^{3}{\setminus }\Gamma \rightarrow \mathbb {H}^{2}_{\mathbb {C}}\), \(U_\mathrm{KN}=u_\mathrm{KN}+\log \rho \), which determines the extreme Kerr–Newman solution is given by

$$\begin{aligned}&u_\mathrm{KN}=-\frac{1}{2}\log \left[ \left( \widetilde{r}^{2}+a^{2}+\frac{a^{2}\sin ^{2}\theta (2m\widetilde{r}-q^{2})}{\Sigma }\right) \sin ^{2}\theta \right] ,\nonumber \\&v_\mathrm{KN}=ma\cos \theta (3-\cos ^{2}\theta )-\frac{a(q^{2}\widetilde{r}-ma^{2}\sin ^{2}\theta )\cos \theta \sin ^{2}\theta }{\Sigma },\nonumber \\&\chi _\mathrm{KN}=-\frac{qa\widetilde{r}\sin ^{2}\theta }{\Sigma },\nonumber \\&\psi _\mathrm{KN}=\frac{q(\widetilde{r}^{2}+a^{2})\cos \theta }{\Sigma }. \end{aligned}$$
(6.6)

The Euler-Lagrange equations satisfied by this and any other harmonic map \(\Psi :\mathbb {R}^{3}\rightarrow \mathbb {H}^{2}_{\mathbb {C}}\) are given by

$$\begin{aligned} \Delta u-2e^{4u}|\nabla v+\chi \nabla \psi -\psi \nabla \chi |^{2} -e^{2u}(|\nabla \chi |^{2}+|\nabla \psi |^{2})&=0,\nonumber \\ {\text {div}}\left[ e^{4u}(\nabla v+\chi \nabla \psi -\psi \nabla \chi )\right]&=0,\nonumber \\ {\text {div}}(e^{2u}\nabla \chi )-2e^{4u}\nabla \chi \cdot (\nabla v+\chi \nabla \psi -\psi \nabla \chi )&=0,\nonumber \\ {\text {div}}(e^{2u}\nabla \psi )+2e^{4u}\nabla \psi \cdot (\nabla v+\chi \nabla \psi -\psi \nabla \chi )&=0. \end{aligned}$$
(6.7)

Consider now the Majumdar–Papapetrou spacetime \(\left( \mathbb {R}\times (\mathbb {R}^{3} {\setminus }\cup _{n=1}^{N}p_{n}),ds^2\right) \) with

$$\begin{aligned} ds^2=-f^{-2}dt^2+f^{2}\delta , \quad f=1+\sum _{n=1}^{N}\frac{m_{n}}{r_{n}}, \end{aligned}$$
(6.8)

where \(m_{n}=\sqrt{(q^{e}_{n})^{2}+(q^{b}_{n})^{2}}\) represents the mass and total electromagnetic charge of each black hole, \(\delta \) is the Euclidean metric, and \(r_{n}\) is the Euclidean distance to each puncture. Axisymmetry may be imposed by choosing the punctures \(p_{n}\) to lie on the z-axis. Cylindrical coordinates \((\rho ,z,\phi )\) in 3-space give rise to Brill coordinates with \(U_\mathrm{MP}=-\log f\), and the 4-potential is given by

$$\begin{aligned} \mathbf {A}=\kappa fdt+\sqrt{1-\kappa ^2}\sum _{n=1}^{N}\frac{m_{n}(z-z_{n})}{r_{n}}d\phi , \quad 0\le \kappa \le 1. \end{aligned}$$
(6.9)

The constant \(\kappa \) relates the electric and magnetic charges to the mass by \(q_{n}^{e}=\kappa m_{n}\) and \(q_{n}^{b}=\sqrt{1-\kappa ^{2}}m_{n}\). Typically the Majumdar–Papapetrou spacetime is stated without magnetic charges, however through a duality rotation

$$\begin{aligned} E=(\cos \vartheta )\tilde{E}-(\sin \vartheta )\tilde{B}, \quad B=(\sin \vartheta )\tilde{E}+(\cos \vartheta )\tilde{B}, \end{aligned}$$
(6.10)

magnetic charge may be introduced so that \(\kappa =\cos \vartheta \). Since \(E=\kappa \nabla \log f\) and \(B=\sqrt{1-\kappa ^{2}}\nabla \log f\), the electromagnetic potentials are obtained from (2.7)

$$\begin{aligned} d\chi _\mathrm{MP}=\kappa \rho (\partial _{z}f d\rho -\partial _{\rho }f dz), \quad d\psi _\mathrm{MP}=\sqrt{1-\kappa ^{2}}\rho (\partial _{z}f d\rho -\partial _{\rho }f dz), \end{aligned}$$
(6.11)

so that

$$\begin{aligned} \chi _\mathrm{MP}=\kappa \sum _{n=1}^{N}\frac{m_{n}(z-z_{n})}{r_{n}}, \quad \chi _\mathrm{MP}=\sqrt{1-\kappa ^{2}}\sum _{n=1}^{N}\frac{m_{n}(z-z_{n})}{r_{n}}. \end{aligned}$$
(6.12)

Lastly, since this spacetime is static there is no angular momentum, and hence \(v_\mathrm{MP}=0\). This, combined with the fact that \(\chi _\mathrm{MP}\) and \(\psi _\mathrm{MP}\) are proportional leads to a harmonic map with a 2-dimensional target that is isometric to hyperbolic space, namely \((u_\mathrm{MP},v_\mathrm{MP},\chi _\mathrm{MP},\psi _\mathrm{MP}):\mathbb {R}^{3}{\setminus }\Gamma \rightarrow \mathbb {H}^{2}\subset \mathbb {H}_{\mathbb {C}}^{2}\) where \(U_\mathrm{MP}=u_\mathrm{MP}+\log \rho \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Khuri, M., Weinstein, G. The positive mass theorem for multiple rotating charged black holes. Calc. Var. 55, 33 (2016). https://doi.org/10.1007/s00526-016-0969-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00526-016-0969-8

Mathematics Subject Classification

Navigation