Next Article in Journal
Traveling Wave Solutions and Conservation Laws of a Generalized Chaffee–Infante Equation in (1+3) Dimensions
Next Article in Special Issue
Quasiperiodic Oscillations and Dynamics of Test Particles around Quasi- and Non-Schwarzschild Black Holes
Previous Article in Journal
Octonion Internal Space Algebra for the Standard Model
Previous Article in Special Issue
Effective Field Theory Description of Horizon-Fluid Determines the Scrambling Time
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Killing Horizons and Surface Gravities for a Well-Behaved Three-Function Generalization of the Kerr Spacetime

School of Mathematics and Statistics, Victoria University of Wellington, P.O. Box 600, Wellington 6140, New Zealand
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Universe 2023, 9(5), 223; https://doi.org/10.3390/universe9050223
Submission received: 6 April 2023 / Revised: 2 May 2023 / Accepted: 3 May 2023 / Published: 8 May 2023
(This article belongs to the Collection Open Questions in Black Hole Physics)

Abstract

:
Thanks to the recent advent of the event horizon telescope (EHT), we now have the opportunity to test the physical ramifications of the strong-field near-horizon regime for astrophysical black holes. Herein, emphasizing the trade-off between tractability and generality, the authors discuss a particularly powerful three-function distortion of the Kerr spacetime, depending on three arbitrary functions of the radial coordinate r, which on the one hand can be fit to future observational data, and on the other hand is sufficiently general so as to encompass an extremely wide class of theoretical models. In all of these spacetimes, both the timelike Hamilton–Jacobi (geodesic) and massive Klein–Gordon (wave) equations separate, and the spacetime geometry is asymptotically Kerr; hence, these spacetimes are well-suited to modeling real astrophysical black holes. The authors then prove the existence of Killing horizons for this entire class of spacetimes, and give tractable expressions for the angular velocities, areas, and surface gravities of these horizons. We emphasize the validity of rigidity results and zeroth laws for these horizons.

1. Introduction

In general relativity, the most astrophysically important exact vacuum solution to the Einstein equations is arguably the Kerr solution [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15] (see also the textbook presentations in [16,17,18,19,20,21,22,23]). This exact solution models a stationary, axially symmetric spacetime, and hence, models the vacuum, exterior spacetime of a rotating, axially symmetric black hole. For the nonrotating case, the spacetime is modeled by the Schwarzschild solution [24], which, due to Birkhoff’s theorem [25,26,27,28], is the unique spherically symmetric vacuum solution to Einstein’s equations. However, Birkhoff’s theorem does not hold for a (3+1)-dimensional axially symmetric vacuum spacetime [10,29,30]. While the Kerr solution models the exterior spacetime of a rotating general relativistic black hole to a high degree of accuracy, it is not guaranteed that the Kerr solution will model exterior spacetime geometry for all astrophysical rotating compact objects (especially in the strong field regime).
Such issues have become increasingly pressing now that we have been able to partially resolve the near horizon structure of some astrophysical black holes with the EHT—the event horizon telescope [31,32,33,34,35,36,37,38,39,40]. We are now in a position to begin to test the applicability of the Kerr solution in the strong field regime of astrophysical black holes.
When performing such probes of the Kerr spacetime, it is extremely useful to have a tractable template to work with, one that is not too general nor too restrictive. These distorted Kerr templates have typically been studied at the level of phenomenological models, where one is primarily guided by the key symmetries, and by requiring the ability to continuously connect one’s distorted Kerr spacetime back to the exact Kerr geometry. More recently there has been some effort devoted to studying the dynamics of such models, attempting to either find or manually construct a suitable Lagrangian model. For instance, many regular black holes, originally proposed as purely phenomenological models, now have dynamical interpretations in terms of nonlinear electrodynamics [41,42,43,44,45,46,47,48,49,50,51]. Similarly, extending the results of [52] to deal with energy condition violations, the phenomenologically inspired “black bounce” spacetimes now have a dynamical interpretation in terms of a combination of phantom scalars and nonlinear electrodynamics [53,54,55] (see also [56,57,58]). Research on these issues is ongoing.
Herein, we focus attention on a triply-infinite (three-function) class of stationary, axially symmetric spacetimes which can be tested against data of exterior spacetimes of real, astrophysical, rotating black holes. We prove the existence of Killing horizons in all of these spacetimes, and give compact and tractable expressions for the horizon area and surface gravity of these modified black holes, verifying that the zeroth law of black hole mechanics is satisfied. Ultimately, since we wish our candidate black hole spacetimes to model the gravitational fields of real astrophysical black holes, some observationally inspired physical constraints must be placed on these general black hole spacetimes. We shall impose:
  • Hamilton–Jacobi timelike separability.
  • Klein–Gordon separability.
  • Asymptotic flatness.
We shall not impose:
  • Dirac separability.
We shall explicitly demand that all of our distorted Kerr templates contain the exact Kerr spacetime as a special case. Stability analyses for these distorted Kerr templates are somewhat subtle. One has to carefully distinguish the notion of the stability of test fields perturbing these spacetimes, from the question of the internal stability of the spacetime itself. Test field stability, at the level of the Regge–Wheeler and Zerelli equations, has been addressed in references [59,60,61]. Once one specifies an appropriate class of gravitational sources, full dynamical stability can also be addressed—see for example [51,62,63,64].

1.1. Hamilton–Jacobi Timelike Separability

Firstly, we wish to impose that the Hamilton–Jacobi equation is timelike separable, since timelike separability of this equation implies the existence of a nontrivial Killing tensor, which then gives rise to a fourth constant of the motion, hence making the timelike geodesic equation integrable. We have good physical evidence to impose separability of the Hamilton–Jacobi equation. Separability of the Hamilton–Jacobi equation excludes the possibility for chaotic particle orbits; we wish to impose this since we certainly have observed long–lived accretion disks. These accretion disks would rapidly destabilize if we had chaotic particle orbits. As shown by Papadopoulos and Kokkotas [65,66], based on earlier research by Benenti and Francaviglia [67] (see also [68]), to guarantee the separability of the (timelike) Hamilton–Jacobi equation, for a general stationary axisymmetric spacetime, we have to impose the condition that the inverse metric can be written, in ( r , θ , ϕ , t ) coordinates, in the following form:
g a b ( r , θ ) = 1 A 1 + B 1 A 2 0 0 0 0 B 2 0 0 0 0 A 3 + B 3 A 4 + B 4 0 0 A 4 + B 4 A 5 + B 5 .
This contravariant metric contains ten arbitrary single-variable functions, where the five A i functions depend only on r and the five B i functions depend only on θ .
The metric (1) has the following Killing tensor:
K a b ( r , θ ) = 1 A 1 + B 1 A 2 B 1 0 0 0 0 B 2 A 1 0 0 0 0 B 1 A 3 A 1 B 3 B 1 A 4 A 1 B 4 0 0 B 1 A 4 A 1 B 4 B 1 A 5 A 1 B 5 .
Independently, Carson and Yagi have also generated a 10-function contravariant metric that separates the Hamilton–Jacobi Equation [69]. Their metric, ultimately, is a specific implementation of the 10-function metric proposed by Papadopoulos and Kokkotas, see Equation (1). We point out that Hamilton–Jacobi separability is known to have direct physical implications for photon orbits [70,71].
Approaching this from another direction, Johannsen was able to show that a certain 8-function metric also separates the Hamilton–Jacobi equation [72]. This 8-function metric is a restriction of the 10-function Papadopoulos–Kokkotas metric. Mathematically, the restriction can be phrased in terms of demanding that the metric determinant be a perfect square. Physically, this corresponds to demanding the existence of a relatively tractable tetrad. Johannsen then imposed further physical constraints [73], by analyzing the metric the context of the PPN framework and by examining the asymptotic behavior of the metric. By imposing these constraints, Johannsen was able to produce a four-function metric satisfying both Hamilton–Jacobi integrability and asymptotic flatness [73] (see also references [74,75]) for an approach based on the Newman–Janis ansatz.

1.2. Klein–Gordon Separability

Secondly, we wish to additionally impose the separability of the wave Equation (the Klein–Gordon equation). It can be shown that this condition is satisfied if the following commutator:
K , R a b K a c R c b R a c K c b ,
is a conserved T 1 1 tensor [68,76,77,78], see also [79,80,81]. That is, a necessary and sufficient condition for Klein–Gordon separability is as follows:
a K , R a b a K a c R c b R a c K c b = 0 ,
where K a b is the Killing tensor of the spacetime and R a b is the Ricci tensor. This condition can be manipulated into a constraint on the A i functions, see [65], so that only three of the five A i are functionally independent (see also [82,83]).
The physical reason we demand separability of the wave equation is to ensure the existence of quasi-normal modes which arise when considering ring-down effects which occur after black hole merger events [84,85] (see also reference [86] for an approach based on the Newman–Janis ansatz).

1.3. Asymptotic Flatness

Thirdly, we will explicitly demand asymptotic flatness—in particular that all of the B i functions are exactly those of the Kerr spacetime. If this were not the case, the angular dependence of the spacetime time would not asymptotically approach Kerr for large r. Even though the techniques used and the mathematical formulations employed are significantly different, there is nevertheless widespread agreement that the combination of Hamilton–Jacobi separability, Klein–Gordon separability, and asymptotic flatness ultimately leads to a distorted Kerr framework with exactly three free functions [65,82,83].
We shall now present a computationally efficient version of the three-function distorted Kerr metric, with a particular emphasis on verifying the existence of Killing horizons, verifying the existence of a rigidity result, explicitly calculating the surface gravity, and verifying that this entire class of three-function distorted Kerr spacetimes satisfies the zeroth law of black hole mechanics.

1.4. Dirac Separability—A Bridge Too Far

However, let us first say a word or two regarding a condition we will not be imposing—Dirac separability [87,88,89]. While separability of the Dirac equation would be nice to have, it is nowhere near as critical to observational astrophysics as Hamilton–Jacobi (geodesic) separability or Klein–Gordon (wave) separability. Because of the Pauli exclusion principle, Fermi fields do not have a macroscopic classical limit—while one might see individual fermions emitted by a black hole, there is no meaningful manner in which coherent Fermion emission might occur. Certainly, the Dirac equation partially separates, due to the timelike and azimuthal Killing vectors, but there is no pressing physical need to demand complete Dirac separability in the remaining ( r , θ ) plane. In contrast, complete separability of the geodesic and wave equations has direct observational significance.

2. Three-Function Generalization of the Kerr Spacetime

We now start our analysis of the three-function distorted Kerr spacetime by considering the following co-tetrad:
e 0 = Δ 0 Σ 0 ( d t a sin 2 ( θ ) d ϕ ) ; e 1 = Σ 0 Δ 0 d r ; e 2 = Σ 0 d θ ; e 3 = sin ( θ ) Σ 0 ( a d t + ( r 2 + a 2 ) d ϕ ) .
Here, we set Δ 0 = r 2 + a 2 and Σ 0 = r 2 + a 2 cos 2 ( θ ) . The corresponding metric tensor is g a b = e A a e B b η A B , which is just Minkowski spacetime in disguise (in oblate spheroidal coordinates, which reduce to spherical polar coordinates as a 0 ).
However, we if now introduce three arbitrary functions of r, namely Δ ( r ) , Φ ( r ) , and Ξ ( r ) , and furthermore, set Σ ( r , θ ) = Ξ ( r ) 2 + a 2 cos 2 ( θ ) , we can consider a modified (distorted) co-tetrad:
e 0 = exp ( Φ ( r ) ) Δ ( r ) Σ ( r , θ ) ( d t a sin 2 ( θ ) d ϕ ) ; e 1 = Σ ( r , θ ) Δ ( r ) d r ; e 2 = Σ ( r , θ ) d θ ; e 3 = sin ( θ ) Σ ( r , θ ) ( a d t + ( Ξ ( r ) 2 + a 2 ) d ϕ ) .
The resulting line element is now:
d s 2 = Δ ( r ) exp ( 2 Φ ( r ) ) a 2 sin 2 ( θ ) Ξ ( r ) 2 + a 2 cos 2 ( θ ) d t 2 + Ξ ( r ) 2 + a 2 cos 2 ( θ ) Δ ( r ) d r 2 + ( Ξ ( r ) 2 + a 2 cos 2 ( θ ) ) d θ 2 2 a sin 2 ( θ ) ( Ξ ( r ) 2 Δ ( r ) exp ( 2 Φ ( r ) ) + a 2 ) Ξ ( r ) 2 + a 2 cos 2 ( θ ) d t d ϕ + ( Ξ ( r ) 2 + a 2 ) 2 exp ( 2 Φ ( r ) ) Δ ( r ) a 2 sin 2 ( θ ) sin 2 ( θ ) Ξ ( r ) 2 + a 2 cos 2 ( θ ) d ϕ 2 .
Note that when we set Δ ( r ) = r 2 2 m r + a 2 , Φ ( r ) = 0 , and Ξ ( r ) = r , we recover the Kerr spacetime written in Boyer–Lindquist coordinates. In order to satisfy the condition that this spacetime is asymptotically Kerr, we impose the following conditions on our three arbitrary functions of r:
Δ ( r ) r 2 ; Φ = o 1 ; Ξ ( r ) r .
It is perhaps worthwhile to point out that the zero rotation a 0 limit is particularly pleasant. For the cotetrad:
e 0 exp ( Φ ( r ) ) Δ ( r ) Ξ ( r ) d t ; e 1 Ξ ( r ) Δ ( r ) d r ; e 2 Ξ ( r ) d θ ; e 3 Ξ ( r ) sin ( θ ) d ϕ ,
while for the line element:
d s 2 exp ( 2 Φ ( r ) ) Δ ( r ) Ξ ( r ) 2 d t 2 + Ξ ( r ) 2 Δ ( r ) d r 2 + Ξ ( r ) 2 [ d θ 2 + sin 2 ( θ ) d ϕ 2 ] .
This three-function class of spacetimes is generically no longer Ricci flat; however, it covers a vast class of physically interesting spacetimes, as outlined in Table 1 below.
Note that the Kerr black-bounce geometry [90,91] is a rotating version of the Simpson–Visser black bounce [92], which in turn generalizes the Morris–Thorne traversable wormhole [93,94,95]. The supergravity inspired STU “balanced” (equal charge) black holes of reference [96] are also a subset of the three-function class considered herein. However, the generic “unbalanced” STU black holes of reference [96], which have null separable geodesics, but not timelike separable geodesics, do not fall into our three-function class (see also reference [97].)
The Carter one-function off-shell spacetime (more on this below) is an asymptotically Kerr restriction of Carter’s canonical two-function off-shell spacetime [79,98,99]. In the other direction the Carter one-function off-shell spacetime can further be restricted to the “eye of storm” spacetime [100,101,102], which is particularly “close to Kerr”.
In counterpoint, Kiselev spacetimes [103,104,105,106] generalize the Schwarzschild and Reissner–Nordström spacetimes, whereas the static spherically symmetric spacetimes with arbitrary Ξ ( r ) are sufficiently general to enable one to deal with static spherically symmetric traversable wormholes [93,94,95,107,108,109,110,111,112,113,114,115,116,117,118,119,120]. (Note that restricting Ξ ( r ) r is more appropriate for stars and planets.)
Indeed, we note that the general three-function spacetime (7) is a specific example of the Hamilton–Jacobi separable spacetime (1), where we can identify:
A 1 = Ξ ( r ) 2 ; A 2 = Δ ( r ) ; A 3 = a 2 exp ( 2 Φ ( r ) ) Δ ( r ) ;
A 4 = a exp ( 2 Φ ( r ) ) ( Ξ ( r ) 2 + a 2 ) Δ ( r ) ; A 5 = exp ( 2 Φ ( r ) ) ( Ξ ( r ) 2 + a 2 ) 2 Δ ( r ) ,
and
B 1 = a 2 cos 2 ( θ ) ; B 2 = 1 ; B 3 = 1 sin 2 ( θ ) ; B 4 = a ; B 5 = a 2 sin 2 ( θ ) .
We emphasize that the B i functions are exactly those of Kerr, which is intentional. In particular, note that B 3 B 5 = B 4 2 . As advertised, only three of the A i functions are functionally independent; in particular, A 3 A 5 = A 4 2 .
We now calculate the nontrivial Killing tensor in ( t , r , θ , ϕ ) coordinates:
K a b ( r , θ ) = K t t 0 0 K t ϕ 0 Δ ( r ) a 2 cos 2 ( θ ) Ξ ( r ) 2 + a 2 cos 2 ( θ ) 0 0 0 0 Ξ ( r ) 2 Ξ ( r ) 2 + a 2 cos 2 ( θ ) 0 K t ϕ 0 0 K ϕ ϕ ,
where
K t t = a 2 exp ( 2 Φ ( r ) ) ( Ξ ( r ) 2 + a 2 ) 2 cos 2 ( θ ) + Ξ ( r ) 2 Δ ( r ) sin 2 ( θ ) Δ ( r ) ( Ξ ( r ) 2 + a 2 cos 2 ( θ ) ) ; K t ϕ = a a 2 exp ( 2 Φ ( r ) ) ( Ξ ( r ) 2 + a 2 ) cos 2 ( θ ) + Ξ ( r ) 2 Δ ( r ) Δ ( r ) ( Ξ ( r ) 2 + a 2 cos 2 ( θ ) ) ; K ϕ ϕ = a 4 exp ( 2 Φ ( r ) ) cos 2 ( θ ) sin 2 ( θ ) + Ξ ( r ) 2 Δ ( r ) Δ ( r ) sin 2 ( θ ) ( Ξ ( r ) 2 + a 2 cos 2 ( θ ) ) .
It is then not hard to check that ( c K a b ) = 0 . Hence, this object is indeed a Killing tensor, and its existence ensures that the Hamilton–Jacobi equation separates.
To check that the wave Equation (Klein–Gordon equation) separates, we need to check that Equation (4) is satisfied. While the commutator [ K , R ] a b is nonzero, it is relatively simple.
Axisymmetry (plus very mild circularity conditions on the Killing vectors) is enough to ensure that in ( t , r , θ , ϕ ) coordinates, both the metric and inverse metric are symmetric and of the form [20]:
0 0 0 0 0 0 0 0 0 0 .
Hamilton–Jacobi separability then guarantees that both K a b and K a b are of this same form. For any spacetime in our three-function distortion of Kerr, both R a b and R a b are also of this same form. (Thus, R a b and R a b do not hold if one merely assumes Hamilton–Jacobi separability.) Under the stated conditions, this is enough to imply that the commutator is then of the even more restrictive form:
[ K , R ] a b = 0 0 0 [ K , R ] t ϕ 0 0 0 0 0 0 0 0 [ K , R ] t ϕ 0 0 0 .
However, then:
a [ K , R ] a b = 1 g a ( g [ K , R ] a b ) = 1 g ϕ { g [ K , R ] t ϕ } , 0 , 0 , t { g [ K , R ] t ϕ } = 0 .
Thus, the Klein–Gordon (wave) equation is indeed separable as advertised.
For completeness we note that explicit calculation yields:
[ K , R ] t ϕ = a exp ( Φ ( r ) ) d d r exp ( Φ ( r ) ) Ξ ( r ) d Ξ ( r ) d r 1 Ξ ( r ) 2 + a 2 cos 2 ( θ ) .
Thus, the commutator is independent of Δ ( r ) and vanishes if both Φ ( r ) 0 and Ξ ( r ) ± r ; an observation that will subsequently be of use when discussing the possibility of having Killing–Yano tensors.

3. Killing Horizons for the Three-Function Generalization of Kerr

Let us now focus on near-horizon and on-horizon physics [121,122,123].
A plausible tentative location for the horizons of this spacetime is that they occur when the g r r component of the metric becomes singular; these locations are specified by:
{ r H i : Δ ( r H i ) = 0 } .
Since on these putative horizons:
g a b a r b r = g r r = Δ ( r ) Σ ( r , θ ) 0 ,
we see that these putative horizons are certainly null hypersurfaces.
For these horizons to be event horizons, they must at a minimum be invariant under both time translations and axial rotations. This is equivalent to the condition that geodesics on the horizon moving in either the t or ϕ direction should stay on the horizon. More generally, geodesics on the horizon moving in the t , θ , or ϕ direction should stay on the horizon. That is:
d 2 r d λ 2 r = r H i = Γ r i j d x i d λ d x j d λ r = r H i = 0 ,
where i = { t , θ , ϕ } , and j = { t , θ , ϕ } . That is, we wish that Γ r i j vanishes on the horizon. However:
Γ r i j = g r r g r ( i , j ) 1 2 g i j , r = 1 2 g r r g i j , r Δ ( r ) .
Thus, certainly Γ r i j 0 on the putative horizons. In fact, a brief explicit calculation yields:
Γ r t t Δ ( r ) ; Γ r t ϕ Δ ( r ) ; Γ r ϕ ϕ Δ ( r ) ; Γ r θ θ Δ ( r ) ;
and
Γ r t θ = 0 ; Γ r θ ϕ = 0 .
Thus, these putative horizons have the same physical effect as event horizons.
To prove that these horizons are indeed Killing horizons, we need to prove the existence of a Killing vector whose norm is null when evaluated on the horizons. There exist two Killing vectors in this spacetime. One associated with time translation symmetry:
ξ a = ( 1 , 0 , 0 , 0 ) a ,
and one associated with axial symmetry:
ψ a = ( 0 , 0 , 0 , 1 ) a .
One can construct a more general Killing vector by taking a linear combination of the two:
K a = ( 1 , 0 , 0 , Ω ) ,
where Ω is a constant. We wish to find a collection of constants Ω H i such that the norm of [ K H i ] a = ( 1 , 0 , 0 , Ω H i ) vanishes on the relevant horizon r H i . That is, we wish to impose:
g a b K a K b r = r H i = g t t + 2 Ω g t ϕ + Ω 2 g ϕ ϕ r = r H i = 0 .
A brief calculation yields:
g a b K a K b = ( Ξ ( r ) 2 + a 2 ) Ω a 2 sin 2 ( θ ) Ξ ( r ) 2 + a 2 cos 2 ( θ ) Δ ( r ) exp ( 2 Φ ( r ) ) [ 1 a Ω sin 2 ( θ ) ] 2 Ξ ( r ) 2 + a 2 cos 2 ( θ ) ,
so that on the horizons:
g a b K a K b r = r H i = ( Ξ ( r H i ) 2 + a 2 ) Ω a 2 sin 2 ( θ ) Ξ ( r H i ) 2 + a 2 cos 2 ( θ ) .
Hence, in order for the horizons to be Killing we must demand that:
Ω H i = a Ξ ( r H i ) 2 + a 2 .
Note that these constants are independent of θ ; hence, these horizons rotate as if they are a rigid body, as we should expect. This rigidity result would in standard general relativity subject to suitable energy conditions be a rigidity theorem; here, it is instead a rigidity observation based on our model for distorted Kerr.
Thus, we have shown that for these three-function models there exist distinct Killing vectors ( K H i ) a , whose norm vanishes on the various horizons; thus, the horizons specified by the condition in Equation (20) are indeed Killing horizons. This assertion is by no means vacuous, there are many other stationary axisymmetric models which can be constructed for which the horizons are not necessarily Killing horizons [124,125,126,127,128].
The area of surfaces of constant r in this spacetime are given by:
S ( r ) = 2 π 0 π g θ θ g ϕ ϕ sin ( θ ) d θ = 2 π 0 π ( Ξ ( r ) 2 + a 2 ) 2 + exp ( Φ ( r ) ) Δ ( r ) a 2 sin 2 ( θ ) sin ( θ ) d θ .
In general, this integral is somewhat messy. However, at the horizons, this equation simplifies greatly. The areas of the horizons are explicitly given by:
S ( r H i ) = 2 π ( Ξ ( r H i ) 2 + a 2 ) 0 π sin ( θ ) d θ = 4 π ( Ξ ( r H i ) 2 + a 2 ) .

4. Surface Gravities for These Killing Horizons

The surface gravity for any Killing horizon is given by [20]:
| κ | = 1 2 ( a K b ) ( a K b ) H ,
where K a is the Killing vector whose norm vanishes on the horizon. We have already given an expression for such a Killing vector above in Equation (28). In view of Killing’s equation, we can also write this as:
| κ | = 1 2 ( a K b ) ( b K a ) H = 1 2 ( a K b ) ( b K a ) H .
Explicit calculation then gives the remarkably simple, and quite straightforward, expression:
κ H i = exp ( Φ ( r H i ) ) Δ ( r H i ) 2 ( Ξ ( r H i ) 2 + a 2 ) .
where the prime denotes differentiation with respect to r, and the sign has been chosen such that the surface gravity of the outermost horizon is positive.
This expression is independent of θ , and hence, the zeroth law is satisfied. Furthermore, this expression can easily be checked against all of the well-known specific examples in the three-function class of spacetimes we are considering.
Note that for the product of surface gravity and horizon area, we have:
κ H i S ( r H i ) = 2 π exp ( Φ ( r H i ) ) Δ ( r H i ) .
It is convenient in some cases to write Δ ( r ) = r 2 2 r m ( r ) + a 2 ; then we have:
κ H i = exp ( Φ ( r H i ) ) r H i 2 ( 1 2 m ( r H i ) ) a 2 2 r H i ( Ξ ( r H i ) 2 + a 2 ) .
The zero-rotation limit, a 0 , then simplifies to:
κ H i exp ( Φ ( r H i ) ) r H i ( 1 2 m ( r H i ) ) 2 Ξ ( r H i ) 2 .

5. Two-Function Generalization of the Kerr Spacetime

When one is not analyzing physics on the throat (or anti-throat) of a wormhole, the areas of surfaces of constant r must be a monotone function of r (this is essentially the definition of a wormhole throat in the current stationary context). Without further loss of generality, we can then set Ξ ( r ) r , and our metric becomes:
d s 2 = Δ ( r ) exp ( 2 Φ ( r ) ) a 2 sin 2 ( θ ) r 2 + a 2 cos 2 ( θ ) d t 2 + r 2 + a 2 cos 2 ( θ ) Δ ( r ) d r 2 + ( r 2 + a 2 cos 2 ( θ ) ) d θ 2 + ( ( r 2 + a 2 ) 2 exp ( 2 Φ ( r ) ) Δ ( r ) a 2 sin 2 ( θ ) ) sin 2 ( θ ) r 2 + a 2 cos 2 ( θ ) d ϕ 2 2 a ( r 2 Δ ( r ) exp ( 2 Φ ( r ) ) + a 2 ) sin 2 ( θ ) r 2 + a 2 cos 2 ( θ ) d t d ϕ .
Our horizons are still given by the condition given in Equation (20), and they are still Killing horizons, but the constant which makes the norm of K a = ( 1 , 0 , 0 , Ω ) null on the horizons is now given by:
Ω r H i = a r H i 2 + a 2 .
The area of surfaces of constant r are now given by:
S ( r ) = 2 π 0 π ( r 2 + a 2 ) 2 + exp ( Φ ( r ) ) Δ ( r ) a 2 sin 2 ( θ ) sin ( θ ) d θ ,
and when evaluated on the horizons, give:
S ( r H i ) = 2 π r H i 2 + a 2 0 π sin ( θ ) d θ = 4 π r H i 2 + a 2 .
Our expression for the surface gravity is now given by:
κ H i = exp ( Φ ( r H i ) ) Δ ( r H i ) 2 r H i 2 + a 2 ,
where the prime again denotes differentiation with respect to r. The zeroth law is still satisfied. Note that we now have:
κ H i S ( r H i ) = 2 π exp ( Φ ( r H i ) ) Δ ( r H i ) .
It is convenient is some cases to write Δ ( r ) = r 2 + a 2 2 r m ( r ) , then we have:
κ H i = exp ( Φ ( r H i ) ) r H i 2 ( 1 2 m ( r H i ) ) a 2 2 r H i r H i 2 + a 2 .
If we let a 0 , then r H i 2 m ( r H i ) , and we have:
κ H i = exp ( Φ ( r H i ) ) 1 2 m ( r H i ) 2 r H i .
This is fully in agreement with known results in static spherical symmetry [129].

6. Would Be Killing–Yano Tensors

A Killing–Yano tensor is an antisymmetric tensor Y a b that satisfies:
( c Y a ) b = 0 .
This implies that a Killing–Yano tensor provides a `square root’ of a Killing tensor, in the sense that:
Y a c Y c b = K a b ,
is a Killing tensor. The existence of a nonzero Killing–Yano tensor in a spacetime is an extremely nontrivial constraint and has significant physical implications [98,99]. It can furthermore be shown that any Killing–Yano tensor commutes with the Ricci tensor [130,131,132]:
[ Y , R ] a b = Y a c R c b R a c Y c b = 0 .
Thus, the Killing tensor constructed from a Killing–Yano tensor Y a b will automatically satisfy [ K , R ] a b = 0 ; hence, Equation (4) is satisfied, and therefore, the wave equation will be separable. Conversely, if [ K , R ] a b 0 , then we cannot have [ Y , R ] a b = 0 , and thus, there cannot be a Killing–Yano tensor.
For our three-function distortion of Kerr, we start by first finding a `square root’ of the Killing tensor, Equation (14). Using our tetrad defined in Equation (6), we can re-write the Killing tensor in a tetrad basis where it takes on a particularly simple (diagonal) form:
K A B ( r , θ ) = a 2 cos 2 ( θ ) 0 0 0 0 a 2 cos 2 ( θ ) 0 0 0 0 Ξ ( r ) 2 0 0 0 0 Ξ ( r ) 2 .
where indices in capital letters denote tetrad indices. We can then make an ansatz for our potential Killing–Yano tensor; in fact, there are two distinct potential Killing–Yano tensors:
( Y 1 ) A B = 0 a cos ( θ ) 0 0 a cos ( θ ) 0 0 0 0 0 0 Ξ ( r ) 0 0 Ξ ( r ) 0 ,
and
( Y 2 ) A B = 0 a cos ( θ ) 0 0 a cos ( θ ) 0 0 0 0 0 0 Ξ ( r ) 0 0 Ξ ( r ) 0 .
One can easily check that both ( Y 1 ) A C ( Y 1 ) C B = K A B and ( Y 2 ) A C ( Y 2 ) C B = K A B , so these are two distinct “square roots” of the Killing tensor. In our coordinate basis, we can write:
( Y 1 ) a b = 0 Y t r a Ξ ( r ) 0 Y t r 0 Y r θ 0 a Ξ ( r ) 0 0 Ξ ( r ) ( Ξ ( r ) 2 + a 2 ) 0 Y r θ Ξ ( r ) ( Ξ ( r ) 2 + a 2 ) 0 ,
and
( Y 2 ) a b = 0 Y t r a Ξ ( r ) 0 Y t r 0 Y r θ 0 a Ξ ( r ) 0 0 Ξ ( r ) ( Ξ ( r ) 2 + a 2 ) 0 Y r θ Ξ ( r ) ( Ξ ( r ) 2 + a 2 ) 0 ,
where Y t r = a exp ( Φ ( r ) ) cos ( θ ) and Y r θ = a 2 exp ( Φ ( r ) ) cos ( θ ) sin 2 ( θ ) .
Unfortunately, these two objects are not (in general) Killing–Yano tensors, since one can check that (in general) ( c ( Y 1 ) a ) b 0 and ( c ( Y 2 ) a ) b 0 . These objects are merely “would be” Killing–Yano tensors. However, there are two special subcases of interest:
  • When Φ ( r ) = 0 and Ξ ( r ) = r , as per our previous discussion, [ K , R ] a b 0 , and furthermore, one has ( c ( Y 1 ) a ) b = 0 , so ( Y 1 ) a b , then this becomes a true Killing–Yano tensor.
  • Similarly, when Φ ( r ) = 0 and Ξ ( r ) = r , as per our previous discussion, we also have [ K , R ] a b 0 , and furthermore, ( c ( Y 2 ) a ) b = 0 , so ( Y 2 ) a b then becomes a true Killing–Yano tensor.
However, in general, there does not exist a Killing–Yano tensor for the full three-function generalization of the Kerr spacetime given by Equation (7).

7. Carter Canonical Off-Shell Metric

7.1. Two-Function Version

From another direction, considerable research has been undertaken into understanding the geometry of the Carter canonical two-function off-shell metric [5,79,98,99]. Carter’s two-function metric overlaps with the three-function metric considered herein, but is not a subset thereof. There is instead a one-function overlap.
Carter’s two-function metric depends on one radial function Δ ( r ) and one angular function Υ ( θ ) . Unfortunately, the angular function Υ ( θ ) will generically destroy the desired asymptotically Kerr behavior, so while the Carter two-function metric is certainly mathematically interesting, it is not of direct astrophysical importance.
The co-tetrad for Carter’s canonical two-function spacetime is as follows:
e 0 = Δ ( r ) r 2 + a 2 cos 2 ( θ ) ( d t a sin 2 ( θ ) d ϕ ) ; e 1 = r 2 + a 2 cos 2 ( θ ) Δ ( r ) d r ; e 2 = r 2 + a 2 cos 2 ( θ ) Υ ( θ ) d θ ; e 3 = Υ ( θ ) r 2 + a 2 cos 2 ( θ ) sin ( θ ) ( a d t + ( r 2 + a 2 ) d ϕ ) .
The Kerr geometry corresponds to Δ ( r ) = r 2 2 m r + a 2 and Υ ( θ ) = 1 .
However, in general for the θ θ metric component, we have:
g θ θ = { r 2 + a 2 cos 2 ( θ ) } Υ ( θ ) .
To have an appropriate asymptotically flat limit g θ θ = r 2 + O ( 1 ) , or even an appropriate asymptotically Kerr limit g θ θ = r 2 + a 2 cos 2 ( θ ) + o ( 1 ) , one is forced to take Υ ( θ ) = 1 . Similar issues arise with the spacetime curvature. For instance, the Ricci scalar is:
R = r 2 Δ ( r ) + θ 2 Υ ( θ ) + 3 cot ( θ ) θ Υ ( θ ) 2 Υ ( θ ) r 2 + a 2 cos 2 ( θ ) .
Trying to enforce suitable asymptotic falloff o ( r 2 ) for the Ricci scalar, given we already know Δ ( r ) r 2 , again forces: Υ ( θ ) = 1 . Overall, we see that whereas the Carter two-function metric is certainly of quite significant mathematical interest, it does not seem to be of direct astrophysical importance.

7.2. One-Function Version

If we set Υ ( θ ) = 1 , then the resulting one-function Carter canonical metric is a subset of the three-function metrics considered above, corresponding to Φ ( r ) = 0 and Ξ ( r ) = r . Explicitly, the co-tetrad for Carter’s canonical one-function spacetime is:
e 0 = Δ ( r ) r 2 + a 2 cos 2 ( θ ) ( d t a sin 2 ( θ ) d ϕ ) ; e 1 = r 2 + a 2 cos 2 ( θ ) Δ ( r ) d r ; e 2 = r 2 + a 2 cos 2 ( θ ) d θ ; e 3 = sin ( θ ) r 2 + a 2 cos 2 ( θ ) ( a d t + ( r 2 + a 2 ) d ϕ ) .
This restricted one-function Carter canonical metric certainly enjoys many nice mathematical features, and is compatible with asymptotic flatness, but might reasonably be viewed as perhaps being a little too restrictive—the three-function metric that was the main theme of this article is somewhat more relaxed; that metric still satisfies the most central of the physically motivated restrictions, while being general enough to observationally interesting.

8. Conclusions

From the discussion above, we have learned that the entire triply infinite class of three-function spacetimes parameterized by { Δ ( r ) , Φ ( r ) , Ξ ( r ) } and given by the following line element:
d s 2 = Δ ( r ) exp ( 2 Φ ( r ) ) a 2 sin 2 ( θ ) Ξ ( r ) 2 + a 2 cos 2 ( θ ) d t 2 + Ξ ( r ) 2 + a 2 cos 2 ( θ ) Δ ( r ) d r 2 + ( Ξ ( r ) 2 + a 2 cos 2 ( θ ) ) d θ 2 2 a ( Ξ ( r ) 2 Δ ( r ) exp ( 2 Φ ( r ) ) + a 2 ) sin 2 ( θ ) Ξ ( r ) 2 + a 2 cos 2 ( θ ) d t d ϕ + ( Ξ ( r ) 2 + a 2 ) 2 exp ( 2 Φ ( r ) ) Δ ( r ) a 2 sin 2 ( θ ) sin 2 ( θ ) Ξ ( r ) 2 + a 2 cos 2 ( θ ) d ϕ 2 .
which has a nontrivial Killing tensor given by Equations (14), (15) and (52); implying the existence of a Carter constant.
Furthermore, in these three-function spacetimes, one has the “conservation law” a [ K , R ] a b = 0 . These conditions imply that both the Hamilton–Jacobi (geodesic) equation and the Klein–Gordon (wave) equation are separable in this entire class of spacetimes. This class of spacetimes can also be arranged to be asymptotically Kerr by imposing:
Δ ( r ) r 2 ; Φ = o 1 ; Ξ ( r ) r .
Hence, these spacetimes are particularly well-suited to modeling real astrophysical black holes. These spacetimes also admit Killing horizons, which are located at the zeroes of Δ ( r ) . That is, at { r H : Δ ( r H ) = 0 } . The angular velocities of these horizons are:
Ω r H i = a Ξ ( r H i ) 2 + a 2 .
This expression is independent of θ , and hence, the horizon rotates is if it were a solid body, thereby yielding a “rigidity” result. The area of the horizons are given by:
S ( r H i ) = 4 π ( Ξ ( r H i ) 2 + a 2 ) ,
and the surface gravity on the horizons is given by the remarkably simple and robust expression:
κ = exp ( Φ ( r H i ) ) Δ ( r H i ) 2 ( Ξ ( r H i ) 2 + a 2 ) .
This expression is independent of θ , and hence, the horizons fulfil the zeroth law of black hole mechanics.
The triply infinite class of three-function spacetimes studied in this article exhibit an interesting (possibly optimal) trade-off between generality and tractability. They cover a vast class of physically interesting spacetimes, as outlined in Table 1 above, and also give toy models that can model the exterior spacetime of real astrophysical black holes. Such models will become particularly useful when probing the near-horizon strong-field dynamics of black holes, as we have already started doing with projects such as the event horizon telescope.

Author Contributions

Conceptualization, M.V.; methodology, J.B. and M.V.; software, J.B. and M.V.; validation, J.B. and M.V.; formal analysis, J.B. and M.V.; investigation, J.B. and M.V.; resources, M.V.; writing—original draft preparation, M.V.; writing—review and editing, J.B. and M.V.; supervision, M.V.; project administration, M.V.; funding acquisition, M.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Marsden Fund, administered by the Royal Society of New Zealand. J.B. was additionally supported by a Doctoral Fellowship provided by Victoria University of Wellington.

Data Availability Statement

All relevant data are included in the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kerr, R. Gravitational field of a spinning mass as an example of algebraically special metrics. Phys. Rev. Lett. 1963, 11, 237–238. [Google Scholar] [CrossRef]
  2. Kerr, R. Gravitational collapse and rotation. In Quasi-Stellar Sources and Gravitational Collapse: Including the Proceedings of the First Texas Symposium on Relativistic Astrophysics, Austin, TX, USA, 16–18 December 1963; Robinson, I., Schild, A., Schücking, E.L., Eds.; University of Chicago Press: Chicago, IL, USA, 1965; pp. 99–102. [Google Scholar]
  3. Newman, E.T.; Janis, A.I. Note on the Kerr spinning particle metric. J. Math. Phys. 1965, 6, 915–917. [Google Scholar] [CrossRef]
  4. Boyer, R.H.; Lindquist, R.W. Maximal analytic extension of the Kerr metric. J. Math. Phys. 1967, 8, 265. [Google Scholar] [CrossRef]
  5. Carter, B. Global structure of the Kerr family of gravitational fields. Phys. Rev. 1968, 174, 1559–1571. [Google Scholar] [CrossRef]
  6. Robinson, D.C. Uniqueness of the Kerr black hole. Phys. Rev. Lett. 1975, 34, 905–906. [Google Scholar] [CrossRef]
  7. Doran, C. A New form of the Kerr solution. Phys. Rev. D 2000, 61, 067503. [Google Scholar] [CrossRef]
  8. O’Neill, B. The Geometry of Kerr Black Holes; Dover: Mineloa, NY, USA, 2014. [Google Scholar]
  9. Wiltshire, D.L.; Visser, M.; Scott, S.M. The Kerr Spacetime: Rotating Black Holes in General Relativity; Cambridge University Press: Cambridge, UK, 2009. [Google Scholar]
  10. Visser, M. The Kerr spacetime: A brief introduction. arxiv 2007, arXiv:0706.0622. Published in Wiltshire, D.L., 2009. [Google Scholar]
  11. Kerr, R.P. Discovering the Kerr and Kerr-Schild metrics. arXiv 2007, arXiv:0706.1109. Published in Wiltshire, D.L., 2009. [Google Scholar]
  12. Teukolsky, S.A. The Kerr Metric. Class. Quantum Gravity 2015, 32, 124006. [Google Scholar] [CrossRef]
  13. Adamo, T.; Newman, E.T. The Kerr–Newman metric: A Review. Scholarpedia 2014, 9, 31791. [Google Scholar] [CrossRef]
  14. Baines, J.; Visser, M. Physically motivated ansatz for the Kerr spacetime. Class. Quantum Gravity 2022, 39, 235004. [Google Scholar] [CrossRef]
  15. Baines, J.; Berry, T.; Simpson, A.; Visser, M. Darboux diagonalization of the spatial 3-metric in Kerr spacetime. Gen. Rel. Grav. 2021, 53, 3. [Google Scholar] [CrossRef]
  16. Adler, R.J.; Bazin, M.; Schiffer, M. Introduction to General Relativity, 2nd ed.; McGraw-Hill: New York, NY, USA, 1975; [It is important to acquire the 1975 second edition, the 1965 first edition does not contain any discussion of the Kerr spacetime.]. [Google Scholar]
  17. D’Inverno, R. Introducing Einstein’s Relativity; Oxford University Press: Oxford, UK, 1992. [Google Scholar]
  18. Hartle, J. Gravity: An introduction to Einstein’s General Relativity; Addison Wesley: San Francisco, CA, USA, 2003. [Google Scholar]
  19. Carroll, S. An Introduction to General Relativity: Spacetime and Geometry; Addison Wesley: San Francisco, CA, USA, 2004. [Google Scholar]
  20. Wald, R. General Relativity; University of Chicago Press: Chicago, IL, USA, 1984. [Google Scholar]
  21. Weinberg, S. Gravitation and Cosmology: Principles and Applications of the General Theory of Relativity; Wiley: Hoboken, NJ, USA, 1972. [Google Scholar]
  22. Hobson, M.P.; Estathiou, G.P.; Lasenby, A.N. General Relativity: An Introduction for Physicists; Cambridge University Press: Cambridge, UK, 2006. [Google Scholar]
  23. Misner, C.; Thorne, K.; Wheeler, J.A. Gravitation; Freeman: San Francisco, CA, USA, 1973. [Google Scholar]
  24. Schwarzschild, K. On the gravitational field of a mass point according to Einstein’s theory. Sitzungsber. Preuss. Akad. Wiss. Berlin Math. Phys. 1916, 1916, 189–196. [Google Scholar]
  25. Birkhoff, G.D. Relativity and Modern Physics; Harvard University Press: Cambridge, MA, USA, 1923. [Google Scholar]
  26. Jebsen, J.T. Über die allgemeinen kugelsymmetrischen Lösungen der Einsteinschen Gravitationsgleichungen im Vakuum. Ark. FÖr Mat. Astron. Och Fys. 1921, 15, 1–9, Translated and reprinted as: On the general symmetric solutions of Einstein’s gravitational equations in vacuo. Gen. Relativ. Cosmol. 2005, 37, 2253–2259.. [Google Scholar]
  27. Deser, S.; Franklin, J. Schwarzschild and Birkhoff a la Weyl. Am. J. Phys. 2005, 73, 261–264. [Google Scholar] [CrossRef]
  28. Johansen, N.V.; Ravndal, F. On the discovery of Birkhoff’s theorem. Gen. Rel. Grav. 2006, 38, 537–540. [Google Scholar] [CrossRef]
  29. Skakala, J.; Visser, M. Birkhoff-like theorem for rotating stars in (2+1) dimensions. arXiv 2009, arXiv:0903.2128. [Google Scholar]
  30. Baines, J.; Berry, T.; Simpson, A.; Visser, M. Painlevé–Gullstrand form of the Lense–Thirring Spacetime. Universe 2021, 7, 105. [Google Scholar] [CrossRef]
  31. Akiyama, K. et al. [Event Horizon Telescope]. First M87 Event Horizon Telescope Results. I. The Shadow of the Supermassive Black Hole. Astrophys. J. Lett. 2019, 875, L1. [Google Scholar] [CrossRef]
  32. Akiyama, K. et al. [Event Horizon Telescope]. First M87 Event Horizon Telescope Results. IV. Imaging the Central Supermassive Black Hole. Astrophys. J. Lett. 2019, 875, L4. [Google Scholar] [CrossRef]
  33. Akiyama, K. et al. [Event Horizon Telescope]. First M87 Event Horizon Telescope Results. VI. The Shadow and Mass of the Central Black Hole. Astrophys. J. Lett. 2019, 875, L6. [Google Scholar] [CrossRef]
  34. Akiyama, K. et al. [Event Horizon Telescope]. First Sagittarius A* Event Horizon Telescope Results. I. The Shadow of the Supermassive Black Hole in the Center of the Milky Way. Astrophys. J. Lett. 2022, 930, L12. [Google Scholar] [CrossRef]
  35. Akiyama, K. et al. [Event Horizon Telescope]. First Sagittarius A* Event Horizon Telescope Results. VI. Testing the Black Hole Metric. Astrophys. J. Lett. 2022, 930, L17. [Google Scholar] [CrossRef]
  36. Psaltis, D.; Ozel, F.; Chan, C.K.; Marrone, D.P. A General Relativistic Null Hypothesis Test with Event Horizon Telescope Observations of the black-hole shadow in Sgr A*. Astrophys. J. 2015, 814, 115. [Google Scholar] [CrossRef]
  37. Broderick, A.E.; Johannsen, T.; Loeb, A.; Psaltis, D. Testing the No-Hair Theorem with Event Horizon Telescope Observations of Sagittarius A*. Astrophys. J. 2014, 784, 7. [Google Scholar] [CrossRef]
  38. Cardoso, V.; Gualtieri, L. Testing the black hole ‘no-hair’ hypothesis. Class. Quantum Gravity 2016, 33, 174001. [Google Scholar] [CrossRef]
  39. Carballo-Rubio, R.; Filippo, F.D.; Liberati, S.; Visser, M. Phenomenological aspects of black holes beyond general relativity. Phys. Rev. D 2018, 98, 124009. [Google Scholar] [CrossRef]
  40. Carballo-Rubio, R.; Filippo, F.D.; Liberati, S.; Pacilio, C.; Visser, M. On the viability of regular black holes. JHEP 2018, 07, 023. [Google Scholar] [CrossRef]
  41. Toshmatov, B.; Stuchlík, Z.; Ahmedov, B. Generic rotating regular black holes in general relativity coupled to nonlinear electrodynamics. Phys. Rev. D 2017, 95, 084037. [Google Scholar] [CrossRef]
  42. Toshmatov, B.; Stuchlík, Z.; Ahmedov, B. Note on the character of the generic rotating charged regular black holes in general relativity coupled to nonlinear electrodynamics. arXiv 2017, arXiv:1712.04763. [Google Scholar]
  43. Toshmatov, B.; Stuchlík, Z.; Ahmedov, B. Comment on “Construction of regular black holes in general relativity”. Phys. Rev. D 2018, 98, 028501. [Google Scholar] [CrossRef]
  44. Rodrigues, M.E.; Junior, E.L.B. Comment on “Generic rotating regular black holes in general relativity coupled to non-linear electrodynamics”. Phys. Rev. D 2017, 96, 128502. [Google Scholar] [CrossRef]
  45. Yu, S.; Gao, C. Exact black hole solutions with nonlinear electrodynamic field. Int. J. Mod. Phys. D 2020, 29, 2050032. [Google Scholar] [CrossRef]
  46. Bretón, N.; Lämmerzahl, C.; Macías, A. Rotating black holes in the Einstein–Euler—Heisenberg theory. Class. Quantum Gravity 2019, 36, 235022. [Google Scholar] [CrossRef]
  47. Cañate, P. Black bounces as magnetically charged phantom regular black holes in Einstein-nonlinear electrodynamics gravity coupled to a self-interacting scalar field. Phys. Rev. D 2022, 106, 024031. [Google Scholar] [CrossRef]
  48. Cañate, P.; Bergliaffa, S.E.P. Transforming singular black holes into regular black holes sourced by nonlinear electrodynamics. arXiv 2022, arXiv:2203.03088. [Google Scholar]
  49. Bronnikov, K.A. Regular black holes sourced by nonlinear electrodynamics. arXiv 2022, arXiv:2211.00743. [Google Scholar]
  50. Kubiznak, D.; Tahamtan, T.; Svitek, O. Slowly rotating black holes in nonlinear electrodynamics. Phys. Rev. D 2022, 105, 104064. [Google Scholar] [CrossRef]
  51. Franzin, E.; Liberati, S.; Mazza, J.; Vellucci, V. Stable rotating regular black holes. Phys. Rev. D 2022, 106, 104060. [Google Scholar] [CrossRef]
  52. Boonserm, P.; Ngampitipan, T.; Visser, M. Mimicking static anisotropic fluid spheres in general relativity. Int. J. Mod. Phys. D 2015, 25, 1650019. [Google Scholar] [CrossRef]
  53. Bronnikov, K.A.; Walia, R.K. Field sources for Simpson-Visser spacetimes. Phys. Rev. D 2022, 105, 044039. [Google Scholar] [CrossRef]
  54. Bronnikov, K.A. Black bounces, wormholes, and partly phantom scalar fields. Phys. Rev. D 2022, 106, 064029. [Google Scholar] [CrossRef]
  55. Rodrigues, M.E.; Silva, M.V.d.S. Source of black bounces in general relativity. Phys. Rev. D 2023, 107, 044064. [Google Scholar] [CrossRef]
  56. Huang, H.; Yang, J. Charged Ellis Wormhole and Black Bounce. Phys. Rev. D 2019, 100, 124063. [Google Scholar] [CrossRef]
  57. Huang, H.; Lü, H.; Yang, J. Bronnikov-like wormholes in Einstein-scalar gravity. Class. Quantum Gravity 2022, 39, 185009. [Google Scholar] [CrossRef]
  58. Bronnikov, K.A.; Badalov, K.; Ibadov, R. Arbitrary Static, Spherically Symmetric Space-Times as Solutions of Scalar-Tensor Gravity. Grav. Cosmol. 2023, 29, 43–49. [Google Scholar] [CrossRef]
  59. Boonserm, P.; Ngampitipan, T.; Visser, M. Regge-Wheeler equation, linear stability, and greybody factors for dirty black holes. Phys. Rev. D 2013, 88, 041502. [Google Scholar] [CrossRef]
  60. Saleh, M.; Thomas, B.B.; Kofane, T.C. Quasinormal modes of gravitational perturbation around regular Bardeen black hole surrounded by quintessence. Eur. Phys. J. C 2018, 78, 325. [Google Scholar] [CrossRef]
  61. Villani, M. Stability of the Hayward black hole under electromagnetic perturbations. Class. Quantum Gravity 2021, 38, 075028. [Google Scholar] [CrossRef]
  62. Toshmatov, B.; Stuchlík, Z.; Ahmedov, B.; Malafarina, D. Relaxations of perturbations of spacetimes in general relativity coupled to nonlinear electrodynamics. Phys. Rev. D 2019, 99, 064043. [Google Scholar] [CrossRef]
  63. Toshmatov, B.; Stuchlík, Z.; Schee, J.; Ahmedov, B. Electromagnetic perturbations of black holes in general relativity coupled to nonlinear electrodynamics. Phys. Rev. D 2018, 97, 084058. [Google Scholar] [CrossRef]
  64. Nomura, K.; Yoshida, D.; Soda, J. Stability of magnetic black holes in general nonlinear electrodynamics. Phys. Rev. D 2020, 101, 124026. [Google Scholar] [CrossRef]
  65. Papadopoulos, G.O.; Kokkotas, K.D. On Kerr black hole deformations admitting a Carter constant and an invariant criterion for the separability of the wave equation. arXiv 2020, arXiv:2007.12125. [Google Scholar] [CrossRef]
  66. Papadopoulos, G.O.; Kokkotas, K.D. Preserving Kerr symmetries in deformed spacetimes. Class. Quantum Gravity 2018, 35, 185014. [Google Scholar] [CrossRef]
  67. Benenti, S.; Francaviglia, M. Remarks on certain separability structures and their applications to General Relativity. Gen. Relativ. Gravit. 1979, 10, 79–92. [Google Scholar] [CrossRef]
  68. Baines, J.; Berry, T.; Simpson, A.; Visser, M. Killing Tensor and Carter constant for Painlevé–Gullstrand form of Lense–Thirring Spacetime. Universe 2021, 7, 473. [Google Scholar] [CrossRef]
  69. Carson, Z.; Yagi, K. Asymptotically flat, parameterized black hole metric preserving Kerr symmetries. Phys. Rev. D 2020, 101, 084030. [Google Scholar] [CrossRef]
  70. Pappas, G.; Glampedakis, K. On the connection of spacetime separability and spherical photon orbits. arXiv 2018, arXiv:1806.04091. [Google Scholar]
  71. Glampedakis, K.; Pappas, G. Modification of photon trapping orbits as a diagnostic of non-Kerr spacetimes. Phys. Rev. D 2019, 99, 124041. [Google Scholar] [CrossRef]
  72. Johannsen, T. Regular Black Hole Metric with Three Constants of Motion. Phys. Rev. D 2013, 88, 044002. [Google Scholar] [CrossRef]
  73. Johannsen, T. Sgr A* and General Relativity. Class. Quantum Gravity 2016, 33, 113001. [Google Scholar] [CrossRef]
  74. Lima, H.C.D., Jr.; Crispino, L.C.B.; Cunha, P.V.P.; Herdeiro, C.A.R. Spinning black holes with a separable Hamilton–Jacobi equation from a modified Newman–Janis algorithm. Eur. Phys. J. C 2020, 80, 1036. [Google Scholar] [CrossRef]
  75. Shaikh, R.; Pal, K.; Pal, K.; Sarkar, T. Constraining alternatives to the Kerr black hole. Mon. Not. Roy. Astron. Soc. 2021, 506, 1229–1236. [Google Scholar] [CrossRef]
  76. Benenti, S.; Chanu, C.; Rastelli, G. Remarks on the connection between the additive separation of the Hamilton–Jacobi equation and the multiplicative separation of the Schrödinger equation. II. First integrals and symmetry operators. J. Math. Phys. 2002, 43, 5223. [Google Scholar] [CrossRef]
  77. Giorgi, E. The Carter tensor and the physical-space analysis in perturbations of Kerr–Newman spacetime. arXiv 2021, arXiv:2105.14379. [Google Scholar]
  78. Franzin, E.; Liberati, S.; Mazza, J.; Simpson, A.; Visser, M. Charged black-bounce spacetimes. JCAP 2021, 7, 036. [Google Scholar] [CrossRef]
  79. Carter, B. Hamilton–Jacobi and Schrodinger Separable Solutions of Einstein’s Equations. Commun. Math. Phys. 1968, 10, 280–310. [Google Scholar] [CrossRef]
  80. Teukolsky, S.A. Rotating Black Holes: Separable Wave Equations for Gravitational and Electromagnetic Perturbations. Phys. Rev. Lett. 1972, 29, 1114. [Google Scholar] [CrossRef]
  81. Kalnins, E.G.; Miller, W., Jr. Killing tensors and variable separation for Hamilton–Jacobi and Helmholtz equations. SIAM J. Math. Anal. 1980, 11, 1011–1026. [Google Scholar] [CrossRef]
  82. Konoplya, R.A.; Stuchlk, Z.; Zhidenko, A. Axisymmetric black holes allowing for separation of variables in the Klein-Gordon and Hamilton–Jacobi equations. Phys. Rev. D 2018, 97, 084044. [Google Scholar] [CrossRef]
  83. Konoplya, R.; Rezzolla, L.; Zhidenko, A. General parametrization of axisymmetric black holes in metric theories of gravity. Phys. Rev. D 2016, 93, 064015. [Google Scholar] [CrossRef]
  84. Kokkotas, K.D.; Schmidt, B.G. Quasi-Normal Modes of Stars and Black Holes. Living Rev. Rel. 1999, 2, 2. [Google Scholar] [CrossRef] [PubMed]
  85. Konoplya, R.A.; Zhidenko, A. Quasinormal modes of black holes: From astrophysics to string theory. Rev. Mod. Phys. 2011, 83, 793. [Google Scholar] [CrossRef]
  86. Chen, C.Y.; Chen, P. Separability of the Klein-Gordon equation for rotating spacetimes obtained from Newman-Janis algorithm. Phys. Rev. D 2019, 100, 104054. [Google Scholar] [CrossRef]
  87. Iyer, B.R.; Vishveshwara, C.V. Separability of the Dirac equation in a class of perfect fluid space-times with local rotational symmetry. J. Math. Phys. 1985, 26, 1034. [Google Scholar] [CrossRef]
  88. Rüdiger, R. Separable systems for the Dirac equation in curved space-times. J. Math. Phys. 1984, 25, 649. [Google Scholar] [CrossRef]
  89. McLenaghan, R. Symmetry operators and separation of variables for the Dirac equation on curved space. In Proceedings of the S4 Conference, Minneapolis, MN, USA, 17–19 September 2010. [Google Scholar]
  90. Mazza, J.; Franzin, E.; Liberati, S. A novel family of rotating black hole mimickers. JCAP 2021, 04, 082. [Google Scholar] [CrossRef]
  91. Islam, S.U.; Kumar, J.; Ghosh, S.G. Strong gravitational lensing by rotating Simpson-Visser black holes. JCAP 2021, 10, 013. [Google Scholar] [CrossRef]
  92. Simpson, A.; Visser, M. Black-bounce to traversable wormhole. JCAP 2019, 02, 042. [Google Scholar] [CrossRef]
  93. Morris, M.S.; Thorne, K.S. Wormholes in space-time and their use for interstellar travel: A tool for teaching general relativity. Am. J. Phys. 1988, 56, 395–412. [Google Scholar] [CrossRef]
  94. Morris, M.S.; Thorne, K.S.; Yurtsever, U. Wormholes, Time Machines, and the Weak Energy Condition. Phys. Rev. Lett. 1988, 61, 1446–1449. [Google Scholar] [CrossRef] [PubMed]
  95. Visser, M. Lorentzian Wormholes: From Einstein to Hawking; Springer: New York, NY, USA, 1995. [Google Scholar]
  96. Cvetic, M.; Gibbons, G.W.; Pope, C.N. STU Black Holes and SgrA*. JCAP 2017, 08, 016. [Google Scholar] [CrossRef]
  97. Keeler, C.; Larsen, F. Separability of Black Holes in String Theory. JHEP 2012, 10, 152. [Google Scholar] [CrossRef]
  98. Frolov, V.; Krtous, P.; Kubiznak, D. Black holes, hidden symmetries, and complete integrability. Living Rev. Rel. 2017, 20, 6. [Google Scholar] [CrossRef]
  99. Frolov, V.P.; Krtous, P.; Kubiznak, D. Separability of Hamilton–Jacobi and Klein-Gordon Equations in General Kerr-NUT-AdS Spacetimes. JHEP 2007, 02, 005. [Google Scholar] [CrossRef]
  100. Simpson, A.; Visser, M. The eye of the storm: A regular Kerr black hole. JCAP 2022, 03, 011. [Google Scholar] [CrossRef]
  101. Simpson, A.; Visser, M. Astrophysically viable Kerr-like spacetime. Phys. Rev. D 2022, 105, 064065. [Google Scholar] [CrossRef]
  102. Ghosh, S.G. A nonsingular rotating black hole. Eur. Phys. J. C 2015, 75, 532. [Google Scholar] [CrossRef]
  103. Kiselev, V.V. Quintessence and black holes. Class. Quantum Gravity 2003, 20, 1187–1198. [Google Scholar] [CrossRef]
  104. Visser, M. The Kiselev black hole is neither perfect fluid, nor is it quintessence. Class. Quantum Gravity 2020, 37, 045001. [Google Scholar] [CrossRef]
  105. Rodrigues, M.E.; Silva, M.V.d.; Vieira, H.A. Bardeen-Kiselev black hole with a cosmological constant. Phys. Rev. D 2022, 105, 084043. [Google Scholar] [CrossRef]
  106. Ṡemiz, I. On the (non)genericity of the Kiselev spacetime. IOP SciNotes 2020, 1, 025206. [Google Scholar] [CrossRef]
  107. Visser, M. Traversable wormholes: Some simple examples. Phys. Rev. D 1989, 39, 3182–3184. [Google Scholar] [CrossRef] [PubMed]
  108. Visser, M. Traversable wormholes from surgically modified Schwarzschild space-times. Nucl. Phys. B 1989, 328, 203–212. [Google Scholar] [CrossRef]
  109. Lemos, J.P.S.; Lobo, F.S.N.; Quinet de Oliveira, S. Morris-Thorne wormholes with a cosmological constant. Phys. Rev. D 2003, 68, 064004. [Google Scholar] [CrossRef]
  110. Lobo, F.S.N. Phantom energy traversable wormholes. Phys. Rev. D 2005, 71, 084011. [Google Scholar] [CrossRef]
  111. Lobo, F.S.N. Stability of phantom wormholes. Phys. Rev. D 2005, 71, 124022. [Google Scholar] [CrossRef]
  112. Teo, E. Rotating traversable wormholes. Phys. Rev. D 1998, 58, 024014. [Google Scholar] [CrossRef]
  113. Bueno, P.; Cano, P.A.; Goelen, F.; Hertog, T.; Vercnocke, B. Echoes of Kerr-like wormholes. Phys. Rev. D 2018, 97, 024040. [Google Scholar] [CrossRef]
  114. Simpson, A.; Visser, M. Regular black holes with asymptotically Minkowski cores. Universe 2019, 6, 8. [Google Scholar] [CrossRef]
  115. Berry, T.; Simpson, A.; Visser, M. Photon spheres, ISCOs, and OSCOs: Astrophysical observables for regular black holes with asymptotically Minkowski cores. Universe 2020, 7, 2. [Google Scholar] [CrossRef]
  116. Lobo, F.S.N.; Rodrigues, M.E.; Silva, M.V.d.; Simpson, A.; Visser, M. Novel black-bounce spacetimes: Wormholes, regularity, energy conditions, and causal structure. Phys. Rev. D 2021, 103, 084052. [Google Scholar] [CrossRef]
  117. Bambi, C.; Stojkovic, D. Astrophysical Wormholes. Universe 2021, 7, 136. [Google Scholar] [CrossRef]
  118. Maeda, H. Simple traversable wormholes violating energy conditions only near the Planck scale. Class. Quantum Gravity 2022, 39, 075027. [Google Scholar] [CrossRef]
  119. Clément, G.; Gal’tsov, D. Rotating traversable wormholes in Einstein–Maxwell theory. Phys. Lett. B 2023, 838, 137677. [Google Scholar] [CrossRef]
  120. Boonserm, P.; Ngampitipan, T.; Simpson, A.; Visser, M. Exponential metric represents a traversable wormhole. Phys. Rev. D 2018, 98, 084048. [Google Scholar] [CrossRef]
  121. Hawking, S.W. Information Preservation and Weather Forecasting for Black Holes. arXiv 2014, arXiv:1401.5761. [Google Scholar]
  122. Visser, M. Physical observability of horizons. Phys. Rev. D 2014, 90, 127502. [Google Scholar] [CrossRef]
  123. Medved, A.J.M.; Martin, D.; Visser, M. Dirty black holes: Symmetries at stationary nonstatic horizons. Phys. Rev. D 2004, 70, 024009. [Google Scholar] [CrossRef]
  124. Cropp, B.; Liberati, S.; Visser, M. Surface gravities for non-Killing horizons. Class. Quantum Gravity 2013, 30, 125001. [Google Scholar] [CrossRef]
  125. Barceló, C.; Liberati, S.; Visser, M. Analogue gravity. Living Rev. Rel. 2005, 8, 12. [Google Scholar] [CrossRef] [PubMed]
  126. Visser, M. Acoustic black holes: Horizons, ergospheres, and Hawking radiation. Class. Quantum Gravity 1998, 15, 1767–1791. [Google Scholar] [CrossRef]
  127. Fischetti, S.; Way, B. Towards Fluid Instabilities of Stationary Non-Killing Horizons. Class. Quantum Gravity 2016, 33, 245009. [Google Scholar] [CrossRef]
  128. Li, R.; Wang, J. Hawking radiation, local temperatures, and nonequilibrium thermodynamics of the black holes with non-Killing horizon. Phys. Rev. D 2021, 104, 026011. [Google Scholar] [CrossRef]
  129. Visser, M. Dirty black holes: Thermodynamics and horizon structure. Phys. Rev. D 1992, 46, 2445–2451. [Google Scholar] [CrossRef]
  130. Lindström, U.; Sarıoğlu, Ö. New currents with Killing–Yano tensors. Class. Quantum Gravity 2021, 38, 195011. [Google Scholar] [CrossRef]
  131. Lindström, U.; Sarıoğlu, Ö. Uses of Killing-Yano Tensors. PoS CORFU 2022, 2021, 149. [Google Scholar] [CrossRef]
  132. Lindström, U.; Sarıoğlu, Ö. Geometry, conformal Killing-Yano tensors and conserved currents. arXiv 2022, arXiv:2206.08037. [Google Scholar]
Table 1. Various spacetimes that fall into the three-function class discussed herein.
Table 1. Various spacetimes that fall into the three-function class discussed herein.
Spacetime Δ ( r ) Φ ( r ) Ξ ( r ) a
Minkowski r 2 + a 2 0rarbitrary
Kerr r 2 + a 2 2 m r 0rnonzero
Kerr–Newman r 2 + a 2 2 m r + Q 2 0rnonzero
Eye of storm r 2 2 e l / r m r + a 2 0rnonzero
Carter one-function off-shellarbitrary0rnonzero
Sugra STU “balanced” r 2 2 m r + a 2 0 r 2 2 m r nonzero
Kerr black bounce r 2 + l 2 2 m r 2 + l 2 + a 2 0 r 2 + l 2 nonzero
Schwarzschild r 2 2 m r 0r0
Reissner–Nordström r 2 2 m r + Q 2 0r0
Kiselevarbitrary0r0
Static spherical symmetryarbitraryarbitraryarbitrary0
Morris–Thorne wormhole r 2 + l 2 0 r 2 + l 2 0
Simpson–Visser black bounce r 2 + l 2 2 m r 2 + l 2 0 r 2 + l 2 0
exponential wormhole r 2 0 r e m / r 0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Baines, J.; Visser, M. Killing Horizons and Surface Gravities for a Well-Behaved Three-Function Generalization of the Kerr Spacetime. Universe 2023, 9, 223. https://doi.org/10.3390/universe9050223

AMA Style

Baines J, Visser M. Killing Horizons and Surface Gravities for a Well-Behaved Three-Function Generalization of the Kerr Spacetime. Universe. 2023; 9(5):223. https://doi.org/10.3390/universe9050223

Chicago/Turabian Style

Baines, Joshua, and Matt Visser. 2023. "Killing Horizons and Surface Gravities for a Well-Behaved Three-Function Generalization of the Kerr Spacetime" Universe 9, no. 5: 223. https://doi.org/10.3390/universe9050223

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop