Skip to main content

Advertisement

Log in

A Penrose-type inequality with angular momentum and charge for axisymmetric initial data

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

Abstract

A lower bound for the ADM mass is established in terms of angular momentum, charge, and horizon area in the context of maximal, axisymmetric initial data for the Einstein–Maxwell equations which satisfy the weak energy condition. If, on the horizon, the given data agree to a certain extent with the associated model Kerr–Newman data, then the inequality reduces to the conjectured Penrose inequality with angular momentum and charge. In addition, a rigidity statement is also proven whereby equality is achieved if and only if the data set arises from the canonical slice of a Kerr–Newman spacetime.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Anglada, P.: Penrose-like inequality with angular momentum for minimal surfaces. Class. Quantum Gravity 35, 045018 (2018). arXiv:1708.04646

  2. Anglada, P.: Penrose-like inequality with angular momentum for general horizons. (2018) (preprint). arXiv:1810.11321

  3. Bekenstein, J.: A universal upper bound on the entropy to energy ratio for bounded systems. Phys. Rev. D 23, 287 (1981)

    Article  ADS  MathSciNet  Google Scholar 

  4. Bray, H.: Proof of the Riemannian Penrose inequality using the positive mass theorem. J. Differ. Geom. 59, 177–267 (2001)

    Article  MathSciNet  Google Scholar 

  5. Brill, D.: On the positive definite mass of the Bondi–Weber–Wheeler time-symmetric gravitational waves. Ann. Phys. 7, 466–483 (1959)

    Article  ADS  MathSciNet  Google Scholar 

  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., 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

  9. Chruściel, P., Nguyen, L.: A lower bound for the mass of axisymmetric connected black hole data sets. Class. Quantum Gravity 28, 125001 (2011). arXiv:1102.1175

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

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

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

  13. Dain, S., Gabach-Clement, M.: Geometrical inequalities bounding angular momentum and charges in general relativity. Living Rev. Relativ. (2018). https://doi.org/10.1007/s41114-018-0014-7

    Article  Google Scholar 

  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. Gabach-Clement, M., Jaramillo, J., Reiris, M.: Proof of the area-angular momentum-charge inequality for axisymmetric black holes. Class. Quantum Gravity 30, 065017 (2012). arXiv:1207.6761

  16. Gibbons, G., Holzegel, G.: The positive mass and isoperimetric inequalities for axisymmetric black holes in four and five dimensions. Class. Quantum Gravity 23, 6459–6478 (2006). arXiv:gr-qc/0606116

  17. Huisken, G., Ilmanen, T.: The inverse mean curvature flow and the Riemannian Penrose inequality. J. Differ. Geom. 59, 353–437 (2001)

    Article  MathSciNet  Google Scholar 

  18. Jaracz, J., Khuri, M.: Bekenstein bounds, Penrose inequalities, and black hole formation. Phys. Rev. D 97, 124026 (2018). arXiv:1802.04438

  19. Khuri, M., Weinstein, G.: The positive mass theorem for multiple rotating charged black holes. Calc. Var. Partial Differ. Equ. 55(2), 1–29 (2016). arXiv.1502.06290v2

  20. Khuri, M., Weinstein, G., Yamada, S.: Extensions of the charged Riemannian Penrose inequality. Class. Quantum Gravity 32, 035019 (2015). arXiv:1410.5027

  21. Khuri, M., Weinstein, G., Yamada, S.: Proof of the Riemannian Penrose inequality with charge for multiple black holes. J. Differ. Geom. 106, 451–498 (2017). arXiv:1409.3271

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

  23. Penrose, R.: Naked singularities. Ann. N. Y. Acad. Sci. 224, 125–134 (1973)

    Article  ADS  Google Scholar 

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

  25. Stephani, H., Kramer, D., Maccallum, M., Hoenselaers, C., Herlt, E.: Exact Solutions to Einstein’s Field Equations, 2nd edn. Cambridge University Press, Cambridge (2003)

    Book  Google Scholar 

  26. Weinstein, G., Yamada, S.: On a Penrose inequality with charge. Commun. Math. Phys. 257(3), 703–723 (2005). arXiv:math/0405602

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marcus Khuri.

Additional information

Publisher's Note

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

M. Khuri acknowledges the support of NSF Grant DMS-1708798.

Appendices

Appendix A. Weyl coordinates

Here we prove Lemma 2.1. In [9] the existence of Weyl coordinates was established by first constructing so called pseudospherical coordinates \((\rho _{s},z_{s},\phi )\), in which the initial data boundary \(\partial M\) is represented by a semi-circle of radius \(\frac{m_0}{2}\) about the origin in the \(\rho _s z_s\)-plane. This contrasts with Weyl coordinates in which the boundary takes the form of an interval on the z-axis in the orbit space. Pseudospherical coordinates are valid on the planar region \({\mathbb {C}}_{+}\setminus D_{m_0/2}=\{\rho _s+iz_s\mid \rho _{s}>0, r_s>m_0/2\}\), where \(r_s^2=\rho _s^2+z_s^2\). In these coordinates the metric takes the standard ‘Brill’ form

$$\begin{aligned} g=e^{-2U_{s}+2\alpha _{s}}(d\rho _{s}^{2}+dz_{s}^{2})+\rho _s^{2}e^{-2U_{s}} (d\phi +A_{\rho _s} d\rho _{s}+A_{z_s} dz_{s})^{2}. \end{aligned}$$
(A.1)

This structure for the metric is preserved under any coordinate change of the plane which yields a conformal transformation, and Weyl coordinates are a particular example of this. The metric coefficients are axisymmetric, smooth up to the boundary in \({\mathbb {C}}_{+}\setminus D_{m_0/2}\) with \(\alpha _s=0\) on the \(z_s\)-axis, and satisfy the fall-off

$$\begin{aligned} U_{s}=O_{1}(r_{s}^{-1/2-\epsilon }),\quad \alpha _{s}=O_{1}(r_{s}^{-1/2-\epsilon }), \quad A_{\rho _s}=O_{1}(r_{s}^{-3/2-\epsilon }),\quad A_{z_s}=O_{1}(r_{s}^{-3/2-\epsilon }).\nonumber \\ \end{aligned}$$
(A.2)

Weyl coordinates \((\rho ,z,\phi )\) are constructed from pseudospherical coordinates as follows. Define complex coordinates \(\zeta _{s}=\rho _{s}+iz_{s}\) and \(\zeta =\rho +iz\) and consider the holomorphic diffeomorphism \(f:{\mathbb {C}}_+\setminus D_{m_0/2}\rightarrow {\mathbb {C}}_+\) given by

$$\begin{aligned} \zeta =f(\zeta _{s})=\zeta _{s}-\frac{m_0^{2}}{4\zeta _{s}}\quad \Rightarrow \quad \rho =\frac{\rho _{s}(r_{s}^{2}-\frac{m_0^{2}}{4})}{r_{s}^{2}},\quad z=\frac{z_{s}(r_{s}^{2}+\frac{m_0^{2}}{4})}{r_{s}^{2}}.\nonumber \\ \end{aligned}$$
(A.3)

Observe that

$$\begin{aligned} \frac{\partial {\zeta }}{\partial {\zeta _s}}=1+\frac{m_0^2}{4\zeta _s^2}, \end{aligned}$$
(A.4)

which is smooth up to the boundary of \({\mathbb {C}}_+\setminus D_{m_0/2}\) and is nonzero except at the points \(\zeta _s=\pm \frac{m_0}{2}i\). Thus by the inverse function theorem, the inverse transformation is holomorphic and has bounded derivatives away from the poles \(\zeta =\pm m_0i\) of the horizon. Near these points we have

$$\begin{aligned} \left| \frac{\partial {\zeta }}{\partial {\zeta _s}}\right| \ge C^{-1}|\zeta _s \mp \frac{m_0}{2}i|\quad \Rightarrow \quad \left| \frac{\partial {\zeta _s}}{\partial {\zeta }}\right| \le \frac{C}{|\zeta _s \mp \frac{m_0}{2}i|}. \end{aligned}$$
(A.5)

In particular, all first derivatives of the real and imaginary parts admit the bound

$$\begin{aligned} \left| \frac{\partial {\rho _s}}{\partial {\rho }}\right| + \left| \frac{\partial {\rho _s}}{\partial {z}}\right| + \left| \frac{\partial {z_s}}{\partial {\rho }}\right| + \left| \frac{\partial {z_s}}{\partial {z}}\right| \le \frac{C}{|\zeta _s \mp \frac{m_0}{2}i|} \end{aligned}$$
(A.6)

near the poles.

The relationship between U, \(\alpha \) of Weyl coordinates and \(U_s\), \(\alpha _s\) of pseudospherical coordinates is given by [9]

$$\begin{aligned} U(\rho ,z)=U_{s}(\rho _{s},z_{s})-\log \frac{\rho _{s}}{\rho },\quad \quad \alpha (\rho ,z)=\alpha _{s}(\rho _s,z_s) +\log \frac{|\zeta _{s}|^{2}-\frac{m_0^{2}}{4}}{|\zeta _{s}^{2}+\frac{m_0^{2}}{4}|}.\nonumber \\ \end{aligned}$$
(A.7)

Note that the second term on the right-hand side of both expressions depends only on the coordinate transformation. For the Schwarzschild solution

$$\begin{aligned} U_{s,0}(\rho _{s},z_{s})=-2\log \frac{2r_{s}+m_0}{2r_{s}},\quad \quad \alpha _{s,0}(\rho _s,z_s)=0, \end{aligned}$$
(A.8)

and the expressions for the Schwarzschild data \(U_0\) and \(\alpha _0\) in Weyl coordinates may then be obtained from the above formulas. We may then write \(U=U_{0}+{\overline{U}}\) and \(\alpha =\alpha _{0}+{\overline{\alpha }}\) where

$$\begin{aligned} {\overline{U}}(\rho ,z):=U(\rho ,z)-U_{0}(\rho ,z) =U_{s}(\rho _{s},z_{s})-U_{s,0}(\rho _{s},z_{s}), \end{aligned}$$
(A.9)

and

$$\begin{aligned} {\overline{\alpha }}(\rho ,z):=\alpha (\rho ,z)-\alpha _{0}(\rho ,z) =\alpha _{s}(\rho _{s},z_{s}). \end{aligned}$$
(A.10)

It immediately follows that \({\overline{U}}\) and \({\overline{\alpha }}\) are uniformly bounded and satisfy the desired decay at infinity. Furthermore since \(U_s\), \(U_{s,0}\), and \(\alpha _s\) are smooth, the regularity properties of \({\overline{U}}\) and \({\overline{\alpha }}\) depend on the coordinate transformation \(f^{-1}\), and the only possible issues arise at the poles.

Consider the partial derivative

$$\begin{aligned} \frac{\partial {\overline{U}}}{\partial \rho }=\left( \frac{\partial U_{s}}{\partial \rho _{s}}-\frac{\partial U_{s,0}}{\partial \rho _{s}}\right) \frac{\partial {\rho _{s}}}{\partial {\rho }} +\left( \frac{\partial U_{s}}{\partial {z_{s}}}-\frac{\partial U_{s,0}}{\partial {z_{s}}}\right) \frac{\partial {z_{s}}}{\partial {\rho }}. \end{aligned}$$
(A.11)

Since the horizon is a minimal surface

$$\begin{aligned} \frac{\partial }{\partial r_{s}}(U_{s}-\frac{1}{2}\alpha _{s}) =\frac{2}{m_0}=\frac{\partial U_{s,0}}{\partial r_{s}}\quad \quad \text { when }r_s=\frac{m_0}{2}. \end{aligned}$$
(A.12)

In particular this holds at \((\rho _{s},z_{s})=(0,\pm m_0/2)\). Moreover, since \(\alpha _{s}=0\) on the axis and \(\partial _{r_s}\) coincides with \(\pm \partial _{z_{s}}\) there, we have

$$\begin{aligned} \left( \frac{\partial U_{s}}{\partial z_{s}}-\frac{\partial U_{s,0}}{\partial z_{s}}\right) \left( 0,\pm \frac{m_0}{2}\right) =0. \end{aligned}$$
(A.13)

Next, use the fact that all functions are axisymmetric to find

$$\begin{aligned} \frac{\partial U_{s}}{\partial \rho _{s}}\left( 0,\pm \frac{m_0}{2}\right) =\frac{\partial U_{s,0}}{\partial \rho _{s}}\left( 0,\pm \frac{m_0}{2}\right) =0. \end{aligned}$$
(A.14)

Therefore the first derivatives of \(U_{s}-U_{s,0}\) vanish at the poles. This, combined with the smoothness of this function up to the boundary, shows that even though \(\partial _{\rho }\rho _s\) and \(\partial _{\rho }z_s\) may blow-up at these points in a manner controlled by (A.6), the full expression (A.11) remains bounded. Similar considerations may be used to treat the \(\partial _z {\overline{U}}\) and the derivatives of \({\overline{\alpha }}\).

Appendix B. Relation of \({\overline{\beta }}\) to surface gravity

Here we compute \({\overline{\beta }}={\overline{\alpha }}-2{\overline{U}}\) on the horizon rod for the Kerr black hole. Let us recall the constant time slice Kerr metric \(g_{kerr}\) in Weyl coordinates [25]. We will denote the mass and angular momentum of the Kerr metric by m and \({\mathcal {J}}=ma\), while the notation for half the horizon rod length will be \(m_0\). Then

$$\begin{aligned} g_{kerr}=e^{-2U_{kerr}+2\alpha _{kerr}}(d\rho ^{2}+dz^{2})+\rho ^{2}e^{-2U_{kerr}}d\phi ^{2}, \end{aligned}$$
(B.1)

where

$$\begin{aligned}&e^{-2U_{kerr}+2\alpha _{kerr}} =\frac{m_0^{2}\left( r_{+}+r_{-}+2m\right) ^{2}+a^{2}(r_{+}-r_{-})^{2}}{4m_0^{2}r_{+}r_{-}}, \end{aligned}$$
(B.2)
$$\begin{aligned}&\rho ^{2}e^{-2U_{kerr}} \nonumber \\&\quad = \frac{m_0^{2}\left( r_{+}+r_{-}+2m\right) ^{2}+a^{2}(r_{+}-r_{-})^{2}}{m_0^{2}\left( (r_{+}+r_{-})^{2}-4m^{2}\right) +a^{2}(r_{+}-r_{-})^{2}}\rho ^{2} \nonumber \\&\qquad -\frac{\left[ am(r_{+}+r_{-}+2m)(4m_0^{2}-(r_{+}-r_{-})^{2})\right] ^{2}}{\left[ m_0^{2}\left( (r_{+}+r_{-})^{2}-4m^{2}\right) +a^{2}(r_{+}-r_{-})^{2}\right] \left[ m_0^{2}\left( r_{+}+r_{-}+2m\right) ^{2}+a^{2}(r_{+}-r_{-})^{2}\right] },\nonumber \\ \end{aligned}$$
(B.3)

with \(r_{\pm }=\sqrt{\rho ^{2}+(z\pm m_0)^{2}}\). Write \(U_{kerr}=U_0+{\overline{U}}_{kerr}\) and \(\alpha _{kerr}=\alpha _0+{\overline{\alpha }}_{kerr}\), where \(U_0\) and \(\alpha _0\) are the corresponding Schwarzschild functions. It follows that for \(|z|<m_0\) we have

$$\begin{aligned} {\overline{U}}_{kerr}(0,z)=-\frac{1}{2} \log \left( \frac{m^{2}(m+m_{0})^{2}}{m_0^{2}(m+m_{0})^{2}+a^{2}z^{2}}\right) , \end{aligned}$$
(B.4)

and

$$\begin{aligned} {\overline{\alpha }}_{kerr}(0,z)=\frac{1}{2}\log \frac{\left[ m_0^{2}(m+m_{0})^{2}+a^{2}z^{2}\right] ^{2}}{4m_0^{4}m^{2}(m+m_{0})^{2}}. \end{aligned}$$
(B.5)

Notice that \({\mathcal {J}}=0\) implies that \({\overline{U}}_{kerr}(0,z)={\overline{\alpha }}_{kerr}(0,z)=0\) as expected, since half the horizon rod length is given by

$$\begin{aligned} m_0=\sqrt{m^{2}-a^{2}} =\sqrt{m^2-\frac{{\mathcal {J}}^{2}}{m^{2}}}. \end{aligned}$$
(B.6)

We now have that on the horizon rod

$$\begin{aligned} {\overline{\beta }}_{kerr}(0,z)={\overline{\alpha }}_{kerr}(0,z)-2{\overline{U}}_{kerr}(0,z) =\log \frac{m(m+m_{0})}{2m_0^{2}}\ge 0. \end{aligned}$$
(B.7)

Consider now the surface gravity of the Kerr black hole

$$\begin{aligned} \kappa =\frac{\sqrt{m^4-{\mathcal {J}}^2}}{2\left( m^3+m\sqrt{m^4-{\mathcal {J}}^2}\right) }. \end{aligned}$$
(B.8)

Comparing the two formulas produces

$$\begin{aligned} {\overline{\beta }}_{kerr}(0,z)=-\log (4m_0\kappa ). \end{aligned}$$
(B.9)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Khuri, M., Sokolowsky, B. & Weinstein, G. A Penrose-type inequality with angular momentum and charge for axisymmetric initial data. Gen Relativ Gravit 51, 118 (2019). https://doi.org/10.1007/s10714-019-2600-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10714-019-2600-8

Keywords

Navigation