Log in

Spherical orbits around a Kerr black hole

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

Abstract

A special class of orbits known to exist around a Kerr black hole are spherical orbits—orbits with constant coordinate radii that are not necessarily confined to the equatorial plane. Spherical time-like orbits were first studied by Wilkins almost 50 years ago. In the present paper, we perform a systematic and thorough study of these orbits, encompassing and extending previous works on them. We first present simplified forms for the parameters of these orbits. The parameter space of these orbits is then analysed in detail; in particular, we delineate the boundaries between stable and unstable orbits, bound and unbound orbits, and prograde and retrograde orbits. Finally, we provide analytic solutions of the geodesic equations, and illustrate a few representative examples of these orbits.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. Here, and subsequently, the first subscript refers to the upper sign, while the second subscript refers to the lower sign.

  2. The marginally bound case \(E^2=1\) will not be treated separately, as it can be obtained by taking the limit \(E^2\rightarrow 1^-\) of the bound case. Expressions for \(w_{1,2}\) in this limit can be found in Eq. (5.15) below.

  3. In the special case \(\Phi =0\), we have \(V(w\mathop {=}1)=0\) and so the single point \(w=1\) is an allowed orbit. It corresponds to a particle that moves along either of the axes \(\theta =0,\pi \). We do not consider such orbits in this paper.

  4. This is, of course, not a sufficient condition. The reality of the square root in (2.6) will also impose a limit to how negative Q can be. Nevertheless, checking the condition (2.7) is sufficient for us to rule out spherical orbits with negative Q.

  5. These two roots are where \(Q_\text {ms}\) vanishes, except in the extremal limit \(a=M\). In this limit, \(Q_\text {ms}\) remains positive at the smaller root, which coincides with the event horizon (c.f. Fig. 1 below).

  6. As in Footnote 5, these two roots are where \(Q_\text {mb}\) vanishes, except when \(a=M\). In this limit, \(Q_\text {mb}\) remains positive at the smaller root, which coincides with the event horizon (c.f. Fig. 1).

  7. This quintic equation cannot be solved in general. But when \(a=M\), it reduces to a cubic equation, and the relevant root is given by

    $$\begin{aligned} r_\text {mb}^*=\frac{M}{3}\bigg [5+2\sqrt{19}\cos \bigg (\frac{1}{3}\arccos \bigg (\frac{187}{722}\sqrt{19}\bigg )\bigg )\bigg ]\simeq 4.61M. \end{aligned}$$
  8. Our conventions for the elliptic integrals follow those of Gradshteyn and Ryzhik [26].

  9. It is also possible to rewrite the right-hand side of this equation in terms of \(u_1\) and \(u_2\), which will lead to some simplifications in the present case. However, we will not do so as \(u_2\) will be imaginary in the unbound case to be considered in Sect. 5.2.

  10. Obtaining this result actually requires expanding r in (A.5) to next order in \(\epsilon \).

References

  1. Kerr, R.P.: Gravitational field of a spinning mass as an example of algebraically special metrics. Phys. Rev. Lett. 11, 237 (1963). https://doi.org/10.1103/PhysRevLett.11.237

    Article  ADS  MathSciNet  MATH  Google Scholar 

  2. Abbott, B., et al.: Observation of gravitational waves from a binary black hole merger. Phys. Rev. Lett. 116, 061102 (2016). https://doi.org/10.1103/PhysRevLett.116.061102. ar**v:1602.03837 [gr-qc]

  3. Akiyama, K. et al.: First M87 event horizon telescope results. I. The shadow of the supermassive black hole. Astrophys. J. 875, L1, (2019). https://doi.org/10.3847/2041-8213/ab0ec7. ar**v:1906.11238 [astro-ph.GA]

  4. Rana, P., Mangalam, A.: Astrophysically relevant bound trajectories around a Kerr black hole. Class. Quant. Grav. 36, 045009 (2019). https://doi.org/10.1088/1361-6382/ab004c. ar**v:1901.02730 [gr-qc]

  5. Kapec, D., Lupsasca, A.: Particle motion near high-spin black holes. Class. Quant. Grav. 37, 015006 (2020). https://doi.org/10.1088/1361-6382/ab519e. ar**v:1905.11406 [hep-th]

  6. Gralla, S.E., Lupsasca, A.: Null geodesics of the Kerr exterior. Phys. Rev. D 101, 044032 (2020). https://doi.org/10.1103/PhysRevD.101.044032. ar**v:1910.12881 [gr-qc]

  7. Stein, L.C., Warburton, N.: Location of the last stable orbit in Kerr spacetime. Phys. Rev. D 101, 064007 (2020). https://doi.org/10.1103/PhysRevD.101.064007. ar**v:1912.07609 [gr-qc]

  8. Compère, G., Druart, A.: Near-horizon geodesics of high-spin blackholes. Phys. Rev. 101, 084042 (2020). https://doi.org/10.1103/PhysRevD.101.084042. Erratum-ibid 102, 029901 (2020). ar**v:2001.03478 [gr-qc]

  9. Rana, P., Mangalam, A.: A geometric origin for quasi-periodic oscillations in black hole X-ray binaries. Astrophys. J. 903, 121 (2020). https://doi.org/10.3847/1538-4357/abb707. ar**v:2009.01832 [astro-ph.HE]

  10. Carter, B.: Global structure of the Kerr family of gravitational fields. Phys. Rev. 174, 1559 (1968). https://doi.org/10.1103/PhysRev.174.1559

    Article  ADS  MATH  Google Scholar 

  11. Mino, Y.: Perturbative approach to an orbital evolution around a supermassive black hole. Phys. Rev. D 67, 084027 (2003). https://doi.org/10.1103/PhysRevD.67.084027. ar**v:gr-qc/0302075

  12. Fujita, R., Hikida, W.: Analytical solutions of bound timelike geodesic orbits in Kerr spacetime. Class. Quant. Grav. 26, 135002 (2009). https://doi.org/10.1088/0264-9381/26/13/135002. ar**v:0906.1420 [gr-qc]

  13. Lämmerzahl, C., Hackmann, E.: Analytical solutions for geodesic equation in black hole spacetimes. Springer Proc. Phys. 170, 43 (2016). https://doi.org/10.1007/978-3-319-20046-0_5. ar**v:1506.01572 [gr-qc]

  14. Wilkins, D.C.: Bound geodesics in the Kerr metric. Phys. Rev. D 5, 814 (1972). https://doi.org/10.1103/PhysRevD.5.814

    Article  ADS  Google Scholar 

  15. Johnston, M., Ruffini, R.: Generalized Wilkins effect and selected orbits in a Kerr-Newman geometry. Phys. Rev. D 10, 2324 (1974). https://doi.org/10.1103/PhysRevD.10.2324

    Article  ADS  Google Scholar 

  16. Stoghianidis, E., Tsoubelis, D.: Polar orbits in the Kerr space-time. Gen.Relativ.Gravit 19, 1235 (1987). https://doi.org/10.1007/BF00759103

  17. Hughes, S.A.: Evolution of circular, nonequatorial orbits of Kerr black holes due to gravitational wave emission. Phys. Rev. D 61, 084004 (2000). https://doi.org/10.1103/PhysRevD.61.084004. ar**v:gr-qc/9910091

  18. Hughes, S.A.: Evolution of circular, nonequatorial orbits of Kerr black holes due to gravitational wave emission. II. Inspiral trajectories and gravitational wave forms. Phys. Rev. D 64, 064004 (2001). https://doi.org/10.1103/PhysRevD.64.064004. ar**v:gr-qc/0104041

  19. Kraniotis, G.: Precise relativistic orbits in Kerr and Kerr-(anti) de Sitter spacetimes. Class. Quant. Grav. 21, 4743 (2004). https://doi.org/10.1088/0264-9381/21/19/016. ar**v:gr-qc/0405095

  20. Fayos, F., Teijón, C.: Geometrical locus of massive test particle orbits in the space of physical parameters in Kerr space-time. Gen.Relativ.Gravit 40, 2433 (2008). https://doi.org/10.1007/s10714-008-0629-1. ar**v:0706.1455 [gr-qc]

  21. Hackmann, E., Lämmerzahl, C., Kagramanova, V., Kunz, J.: Analytical solution of the geodesic equation in Kerr-(anti-) de Sitter space-times. Phys. Rev. D 81, 044020 (2010). https://doi.org/10.1103/PhysRevD.81.044020. ar**v:1009.6117 [gr-qc]

  22. Grossman, R., Levin, J., Perez-Giz, G.: The harmonic structure of generic Kerr orbits. Phys. Rev. D 85, 023012 (2012). https://doi.org/10.1103/PhysRevD.85.023012. ar**v:1105.5811 [gr-qc]

  23. Hod, S.: Marginally bound (critical) geodesics of rapidly rotating black holes. Phys. Rev. D 88, 087502 (2013). https://doi.org/10.1103/PhysRevD.88.087502. ar**v:1707.05680 [gr-qc]

  24. Teo, E.: Spherical photon orbits around a Kerr black hole. Gen.Relativ.Gravit 35, 1909 (2003). https://doi.org/10.1023/A:1026286607562

  25. Bardeen, J.M., Press, W.H., Teukolsky, S.A.: Rotating black holes: locally nonrotating frames, energy extraction, and scalar synchrotron radiation. Astrophys. J. 178, 347 (1972). https://doi.org/10.1086/151796

    Article  ADS  Google Scholar 

  26. Gradshteyn, I.S., Ryzhik, I.M., Jeffrey, A. (eds.): Table of Integrals, Series, and Products, 5th edn. Academic Press, London (1994)

    MATH  Google Scholar 

  27. Abramowitz, M., Stegun, I.A. (eds.): Handbook of Mathematical Functions. Dover, New York (1972)

    MATH  Google Scholar 

  28. Byrd, P.F., Friedman, M.D.: Handbook of Elliptic Integrals for Engineers and Scientists, 2nd edn. Springer, Berlin (1971)

    Book  Google Scholar 

  29. Prudnikov, A.P., Brychkov, Y.A., Marichev, O.I.: Integrals and Series, vol. 1. Gordon and Breach, New York (1986)

    MATH  Google Scholar 

  30. Goldstein, H.: Numerical calculation of bound geodesics in the Kerr metric. Z. Phys. 271, 275 (1974). https://doi.org/10.1007/BF01677935

    Article  ADS  Google Scholar 

  31. Ryan, F.D.: Effect of gravitational radiation reaction on circular orbits around a spinning black hole. Phys. Rev. D 52, 3159 (1995). https://doi.org/10.1103/PhysRevD.52.R3159. ar**v:gr-qc/9506023

  32. Kennefick, D., Ori, A.: Radiation reaction induced evolution of circular orbits of particles around Kerr black holes. Phys. Rev. D 53, 4319 (1996). https://doi.org/10.1103/PhysRevD.53.4319. ar**v:gr-qc/9512018

  33. Levin, J., Perez-Giz, G.: Homoclinic orbits around spinning black holes I Exact solution for the Kerr separatrix. Phys. Rev. D 79, 124013 (2009). https://doi.org/10.1103/PhysRevD.79.124013. ar**v:0811.3814 [gr-qc]

Download references

Acknowledgements

I would like to acknowledge all the past students of the NUS Physics Department, who have contributed in one way or another to this project. I also wish to thank the reviewers for suggestions that have helped improve the presentation of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Edward Teo.

Additional information

Publisher's Note

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

Appendices

A Horizon-skimming orbits

In [14], Wilkins pointed out the existence of a class of so-called horizon-skimming orbits, which appear to lie on the event horizon of the extremal Kerr black hole with \(a=M\). They arise by taking the \(r\rightarrow r_1\) limit of the solution \((E_\text {a},\Phi _\text {a})\), and have the energy and angular momentum

$$\begin{aligned} E=\frac{\sqrt{M^2+Q}}{\sqrt{3}M},\qquad \Phi =2ME, \end{aligned}$$
(A.1)

where Q takes the range \(0\le Q<\infty \). These orbits are represented by the black line on the left edge of the parameter space of Fig. 1. The limit \(Q\rightarrow \infty \) corresponds to taking the null limit of these orbits [24].

The fact that the radii of these orbits coincide with that of the event horizon, is due to the well-known fact that the extremal Kerr black hole has an infinite throat in this region of the space-time [25]. Points along this throat share the same coordinate radius \(r=M\), even though they might be (finitely or even infinitely) separated in space. Thus the horizon-skimming orbits remain above the event horizon; in fact, they also remain above the prograde circular photon orbit at \(r_1\).

To resolve the throat region, we introduce a new parameter \(\epsilon \) defined by

$$\begin{aligned} \epsilon =\sqrt{1-\frac{a^2}{M^2}}. \end{aligned}$$
(A.2)

The extremal limit is then taken as \(\epsilon \rightarrow 0\). We would like to understand the region of the parameter space near \(r_1\) as this limit is taken. In particular, we focus on the marginally stable and marginally bound orbits in this region. Our results are consistent with those recently obtained in [8].

Recall that marginally stable orbits are described by the curve \(Q=Q_\text {ms}\). Substituting an expansion of the form \(r=M+A\epsilon ^p+\cdots \) into the right-hand side of this equation, and requiring that the lowest-order term is zeroth order in \(\epsilon \), implies that \(p=2/3\). The coefficient A can then be determined in terms of Q, and we obtain [8]

$$\begin{aligned} r=M\Bigg [1+\bigg (\frac{M^2+Q}{M^2/2-Q}\bigg )^{1/3}\,\epsilon ^{2/3}\Bigg ]+O(\epsilon ^{4/3}). \end{aligned}$$
(A.3)

This parameterises the radii of these marginally stable orbits in terms of Q, which takes the range \(0\le Q<M^2/2\). The energy and angular momentum of these orbits are given by

$$\begin{aligned} E&=\frac{\sqrt{M^2+Q}}{\sqrt{3}M}\Bigg [1+\bigg (\frac{\sqrt{2}(M^2-2Q)\epsilon }{M^2+Q}\bigg )^{2/3}\Bigg ]+O(\epsilon ^{4/3}), \end{aligned}$$
(A.4a)
$$\begin{aligned} \Phi&=2E+O(\epsilon ^{4/3}). \end{aligned}$$
(A.4b)

On the other hand, recall that marginally bound orbits are described by the curve \(Q=Q_\text {mb}\). Again, substituting an expansion of the form \(r=M+A\epsilon ^p+\cdots \) into the right-hand side of this equation, and requiring that the lowest-order term is zeroth order in \(\epsilon \), implies that now \(p=1\). The coefficient A can then be determined in terms of Q, and we obtain [8]

$$\begin{aligned} r=M\bigg (1+\frac{2M\epsilon }{\sqrt{2M^2-Q}}\bigg )+O(\epsilon ^2). \end{aligned}$$
(A.5)

This parameterises the radii of these marginally bound orbits in terms of Q, which takes the range \(0\le Q<2M^2\). The angular momentum of these orbits is given by [8]Footnote 10

$$\begin{aligned} \Phi =2M+\sqrt{2M^2-Q}\,\epsilon +O(\epsilon ^2). \end{aligned}$$
(A.6)
Fig. 4
figure 4

The (rQ) parameter space near \(r_1\) when \(a=0.999995M\) (corresponding to \(\epsilon \simeq 0.003\)). The blue and red curves are, as in Fig. 1, the \(Q=Q_\text {ms}\) and \(Q_\text {mb}\) curves, respectively (colour figure online)

The (rQ) parameter space near \(r_1\) is shown in Fig. 4 for the case when \(a=0.999995M\), corresponding to \(\epsilon \simeq 0.003\). The blue and red curves are, as in Fig. 1, the \(Q=Q_\text {ms}\) and \(Q_\text {mb}\) curves, respectively. They terminate on the r-axis at \(r_\text {ms}\) and \(r_\text {mb}\), respectively, if we borrow the notation of [25] in the equatorial limit. We would now like to understand what happens to this part of the parameter space, and in particular the two curves, when we take \(\epsilon \rightarrow 0\).

We begin with the red curve corresponding to marginally bound orbits. We have seen that the part of this curve for which \(0\le Q<2M^2\) is approximated by (A.5) when \(\epsilon \) is small. In the limit \(\epsilon \rightarrow 0\), this part of the red curve gets “flattened” onto the black line on the left edge of Fig. 1, between \(Q=0\) and \(2M^2\). Thus, we see that although the red curve appears to terminate at the non-zero value \(Q=2M^2\) in Fig. 1, it actually continues down to \(Q=0\) along the black line.

A similar situation happens for the blue curve corresponding to marginally stable orbits. The part of this curve for which \(0\le Q<M^2/2\) is approximated by (A.3) when \(\epsilon \) is small. In the limit \(\epsilon \rightarrow 0\), this part of the curve gets “flattened” onto the same black line in Fig. 1, but now between \(Q=0\) and \(M^2/2\). Thus, the blue curve does not actually terminate at \(Q=M^2/2\) in Fig. 1, but continues down to \(Q=0\) along the black line.

It follows that the class of horizon-skimming orbits consists of at least a family of marginally bound orbits, and a family of marginally stable orbits, all sharing the same coordinate radius \(r=M\). However, as mentioned above, these orbits are separated in space along the throat of the extremal Kerr black hole. In fact, it can be shown that the distance between \(r_\text {mb}\) and \(r_\text {ms}\) becomes infinite in the limit \(\epsilon \rightarrow 0\) [25]. The distance between \(r_\text {ms}\) and the far regions of the space-time also becomes infinite in this limit. This is a manifestation of the fact that the throat is divided into distinct regions, as depicted in Fig. 2 of [25] (see also Fig. 1 of [5]). The marginally bound orbits belong to one region (together with the photon orbit at \(r_1\) and the event horizon itself), while the marginally stable orbits belong to another region. Other spherical orbits can also exist in these throat regions, and their locations relative to the marginally bound and marginally stable orbits are determined by the dependence of their radii on \(\epsilon \). Geodesic motion in these throat regions have been the focus of recent attention in [5, 8].

B Spherical photon orbits

In this appendix, we provide analytic solutions of the geodesic equations for the spherical photon orbits found in [24]. Recall that the null case corresponds to setting \(\mu =0\) in (2.4) and (2.5). This case can also be recovered from the time-like case \(\mu =1\), by taking the limit \(E\rightarrow \infty \) of the solution \((E_\text {a},\Phi _\text {a})\) when \(r_1<r<r_2\). As mentioned in Sect. 4.4, the ratios \(\Phi /E\) and \(Q/E^2\) remain finite in this limit. If we redefine \(\Phi /E\rightarrow \Phi \) and \(Q/E^2\rightarrow Q\), we arrive at the solution that was obtained in [24]:

$$\begin{aligned} \Phi&=-\,\frac{r\Delta -M(r^2-a^2)}{a(r-M)}, \end{aligned}$$
(B.1a)
$$\begin{aligned} Q&=-\,\frac{r^3\** }{a^2(r-M)^2}. \end{aligned}$$
(B.1b)

With these values of \(\Phi \) and Q, \(w_{1,2}\) can be calculated using

$$\begin{aligned} w_{1,2}=\frac{1}{2a^2}\bigg [a^2-Q-\Phi ^2\pm \sqrt{\big (a^2-Q-\Phi ^2\big )^2+4a^2Q}\bigg ]. \end{aligned}$$
(B.2)

The coordinates \((u,\phi ,t)\) of the geodesic can then be expressed in terms of the Mino parameter \(\lambda \) as

$$\begin{aligned} u&=\sqrt{-w_2}\,k{{\,\mathrm{sd}\,}}\big (a\sqrt{w_1-w_2}\,\lambda ,k\big ),\end{aligned}$$
(B.3a)
$$\begin{aligned} \phi&=\frac{\Phi }{a(1-w_2)\sqrt{w_1-w_2}}\big [F(\psi ,k)-w_2\Pi \big (\psi ,k^2(1-w_2),k\big )\big ]\nonumber \\&\quad +\frac{a}{\Delta }(2Mr-a\Phi )\lambda ,\end{aligned}$$
(B.3b)
$$\begin{aligned} t&=-\,\frac{a}{\sqrt{w_1-w_2}}\big [(1-w_2)F(\psi ,k)+w_2\Pi \big (\psi ,k^2,k\big )\big ]\nonumber \\&\quad +\frac{1}{\Delta }\big [(r^2+a^2)^2-2Mra\Phi \big ]\lambda , \end{aligned}$$
(B.3c)

where

$$\begin{aligned} \psi ={{\,\mathrm{am}\,}}\big (a\sqrt{w_1-w_2}\,\lambda ,k\big ), \end{aligned}$$
(B.4)

and k is given by (5.17b).

It follows that u is a periodic function of \(\lambda \), with period

$$\begin{aligned} \Delta \lambda =\frac{4K(k)}{a\sqrt{w_1-w_2}}. \end{aligned}$$
(B.5)

The change in \(\phi \) and t for one period \(\Delta \lambda \) are

$$\begin{aligned} \Delta \phi&=\frac{4}{\sqrt{w_1-w_2}}\bigg [\frac{\Phi }{a(1-w_1)}\,\Pi \Big (-\frac{w_1}{1-w_1},k\Big )+\frac{2Mr-a\Phi }{\Delta }\,K(k)\bigg ],\end{aligned}$$
(B.6a)
$$\begin{aligned} \Delta t&=\frac{4}{\sqrt{w_1-w_2}}\bigg [-a\big ((1-w_2)K(k)+w_2\Pi \big (k^2,k\big )\big )\nonumber \\&\quad +\frac{(r^2+a^2)^2-2Mra\Phi }{a\Delta }\,K(k)\bigg ]. \end{aligned}$$
(B.6b)

We note that the result for \(\Delta \phi \) agrees with that obtained in [24].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Teo, E. Spherical orbits around a Kerr black hole. Gen Relativ Gravit 53, 10 (2021). https://doi.org/10.1007/s10714-020-02782-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10714-020-02782-z

Keywords

Navigation