This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A Methodological Framework for Solving Einstein’s Equations in Axially Symmetric Spacetimes

J. Ospino and J.L. Hernández-Pastora [email protected] Departamento de Matemática Aplicada and Instituto Universitario de Física Fundamental y Matemáticas, Universidad de Salamanca, Salamanca 37007, Spain    A.V. Araujo-Salcedo [email protected] Centro de Estudios, Gimnasio Campestre, Bogotá, Colombia    L.A. Núñez [email protected] Escuela de Física, Universidad Industrial de Santander, Bucaramanga, Colombia
Departamento de Física, Universidad de Los Andes, Mérida, Venezuela
Abstract

This work presents a novel methodology for deriving stationary and axially symmetric solutions to Einstein’s field equations using the 1+3 tetrad formalism. This approach reformulates the Einstein equations into first-order scalar equations, enabling systematic resolution in vacuum scenarios. We derive two distinct solutions in polar and hyperbolic geometries by assuming the separability of a key metric function. Our method reproduces well-known solutions such as Schwarzschild and Kerr metrics and extends the case of rotating spacetimes to hyperbolic configurations. Additionally, we explore the role of Killing tensors in enabling separable metric components, simplifying analyses of geodesic motion and physical phenomena. This framework demonstrates robustness and adaptability for addressing the complexities of axially symmetric spacetimes, paving the way for further applications to Kerr-like solutions in General Relativity.

Nonspherical sources, Exterior solutions.
pacs:
04.40.-b, 04.40.Nr, 04.40.Dg

I Introduction

The Event Horizon Telescope (EHT) has revolutionised our understanding of axially symmetric spacetimes by providing groundbreaking observational evidence. Its iconic images of the supermassive black holes in M87 and Sagittarius A* unveiled shadow-like features that align with theoretical predictions based on the Kerr metric AkiyamaEtal2019A ; AkiyamaEtal2019B . These observations represent the first direct confirmation of an event horizon’s existence, offering crucial constraints on the spin and mass of black holes AkiyamaEtal2022A . A key coming goal of the EHT is to measure the properties of emission rings, apparent shadows, and other observables that are directly influenced by the spacetime characteristics of M87* and Sgr A* AkiyamaEtal2024A .

These groundbreaking images have sparked significant scientific discussions, particularly regarding the implications of two key astrophysical vacuum solutions to Einstein’s field equations: the Schwarzschild and Kerr metrics. Notably, some observations from the object in SgrASgrA* appear to conflict with the predictions of the Kerr solution AkiyamaEtal2022A . Researchers have explored these discrepancies, with several studies attempting to explain the geodesic motion of the SS-stars using the classical rotating Kerr solution PerlickEtal2022 . Alternatively, YounsiEtal2023 proposed two distinct spacetime geometries: one derived from a modified gravity theory and another that deviates parametrically from the Kerr solution, reducing to Kerr spacetime when the deviation parameters vanish.

Decades before this observational boom, a considerable effort has been dedicated to developing methods for deriving axisymmetric and stationary solutions to Einstein’s vacuum field equations. Comprehensive reviews of these techniques and their results can be found in Quevedo1990 ; StephaniEtal2006 and the references therein. Early in the history of General Relativity, in 1917, Weyl Weyl1917 derived the general family of axisymmetric static solutions. It was just one year after Schwarzschild’s pioneering spherically symmetric solution Schwarzschild1999 was published. Decades later, in 1985, Gutsunayev and Manko GutsunayevManko1985 ; DenisovaKhakimovManko1994 reformulated this family of solutions using a different representation of Weyl’s work. However, it was Roy Kerr who, in 1963, discovered the first physically realistic stationary axisymmetric solution Kerr1963 , marking a pivotal moment, leading to the development of various techniques and encouraging significant efforts in the relativistic community to derive exact solutions to Einstein’s vacuum equations.

Early solution-generating methods incorporate the Kerr-Schild ansatz Kerr1963 , complex transformations NewmanJanis1965 , and approaches based on Hamilton-Jacobi separability Carter1968 . Other notable contributions include techniques by Newman, Tamburino, and Unti NewmanTamburinoUnti1963 , Kinnersley Kinnersley1969 , and Plebanski and Demianski PlebanskiDemianski1976 .

Much of the progress in generating axially symmetric solutions is credited to Ernst’s scheme Ernst1968A ; Ernst1968B , which reformulated the Einstein system and introduced two powerful generating techniques. These methods were later expanded and unified by Kinnersley Kinnersley1973 ; Kinnersley1977 . Building on Ernst’s compact field equation framework, Tomimatsu and Sato TomimatsuSato1972 ; TomimatsuSato1973 , along with Yamazaki and collaborators YamazakiHori1977 ; Yamazaki1977A ; Yamazaki1977B ; YamazakiHori1978 , developed new solutions by specifying particular functional forms for the transformation of the Ernst potential.

Modern generating techniques often involve Lie group or Bäcklund transformations. For example, Cosgrove utilised these methods to derive generalised Tomimatsu-Sato solutions Cosgrove1977A ; Cosgrove1977B , while Maison Maison1979 studied an infinite-dimensional group and the Geroch group Geroch1971 ; Geroch1972 . Although the Geroch group does not yield asymptotically flat solutions, Kinnersley and Chitre KinnersleyChitre1977 ; KinnersleyChitre1978A ; KinnersleyChitre1978B ; KinnersleyChitre1978C identified infinitesimal subgroups that preserve asymptotic flatness. Using the Zipoy-Vorhees metric Zipoy1966 , they generated solutions that include the NUT class and the Tomimatsu-Sato metric as particular cases. Later, Belinskii and Zakharov BelinskiiZakharov1978 formulated a linear eigenvalue problem equivalent to the nonlinear field equations, solvable via the inverse scattering method. This approach was further developed by researchers such as Sibgatullin Sibgatullin1984 and Harrison Harrison1978 ; Harrison1980 . Neugebauer also made significant contributions to these methods HoenselaersKinnersley1979A ; HoenselaersKinnersley1979B ; Neugebauer1980A ; Neugebauer1980B .

Sibgatullin’s method Sibgatullin1984 has become widely used for generating exact solutions, with applications by authors such as V.S. Manko MankoMartinRuiz1994 ; MankoEtal1994 , L. Herrera HerreraManko1992 , and E. Ruiz RuizMankoMartin1995 . Ruiz further refined the method to provide general expressions applicable in a standardised and straightforward manner MankoEtal1994 . This technique constructs solutions to the Ernst equation based on the form of the Ernst potential along the symmetry axis. Static and axisymmetric vacuum solutions have been derived using relativistic multipole moments, as demonstrated by Hernández-Pastora and Martín HernandezpastoraMartin1993 ; HernandezpastoraMartin1994 . These solutions are part of the Weyl family, showcasing the versatility of modern methods in exploring complex spacetime geometries.

In this work, we develop a new method to solve Einstein’s equations for stationary, axially symmetric sources by assuming the separability of a metric function, Φ(r,θ)=ΦR(r)Φ(θ)\Phi(r,\theta)~{}=~{}\Phi_{R}(r)\Phi(\theta) and that the spacetime admitting a Killing Tensor field. The separability yields two distinct physically viable solutions: one expressed in polar coordinates and the other involving hyperbolic functions. To ensure consistency, we required a Killing tensor field which generalises the Carter constantOspinoHernandezpastoraNunez2022 and allows for the separability of additional metric function. By applying the method in polar coordinates, we recover the Kerr solution. This scheme naturally leads to developing a rotating exterior solution for hyperbolic spacetimes.

Hyperbolic geometry in General Relativity enables alternative black hole interior descriptions, preserving staticity and avoiding singularities. It models exotic matter with negative energy density and vacuum cavities, supports thermodynamic insights, and reveals repulsive gravitational effects. Its role extends to dynamic systems, anisotropic compact objects, and non-standard cosmological structures HerreraDiPriscoOspino2021B ; HerreraDiPriscoOspino2021 .

This work is organized as follows. Section II introduces the tetrad used to describe the stationary, axially symmetric solution and the scalars derived from the metric. The kinematical variables are then expressed in terms of these scalars. Next, in section III, the vacuum field equations regarding the scalars derived from the Ricci identities are explicitly presented. Additionally, specific combinations of these equations are performed to isolate the independent equations that need to be solved. The following section, IV, focuses on obtaining the general solution to the equations and describes the overall method. It includes resolving static and stationary solutions for polar and hyperbolic geometries. The simplifications introduced by requiring the existence of a Killing tensor are explored in Section 81. Finally, section VI concludes with a summary of the objectives achieved and key findings.

II Tetrad & kinematical variables

We shall consider stationary and axially symmetric sources with the line element written as

ds2=A2dt2+B2dr2+C2dθ2+R2dϕ2+2ω3dtdϕ,{\rm d}s^{2}=-A^{2}{\rm d}t^{2}+B^{2}{\rm d}r^{2}+C^{2}{\rm d}\theta^{2}+R^{2}{\rm d}\phi^{2}+2\omega_{3}{\rm d}t\,{\rm d}\phi\,, (1)

with A=A(r,θ)A=A(r,\theta), B=B(r,θ)B=B(r,\theta), C=C(r,θ)C=C(r,\theta), R=R(r,θ)R=R(r,\theta), and ω3=ω3(r,θ)\omega_{3}~{}=~{}\omega_{3}(r,\theta).

In this case the components of the orthonormal tetrad V, K, L and S are:

Vα\displaystyle V^{\alpha} =\displaystyle= (1A,0,0,0),Kα=(0,1B,0,0),\displaystyle\left(\frac{1}{A},0,0,0\right),\quad K^{\alpha}=\left(0,\frac{1}{B},0,0\right),
Lα\displaystyle\quad L^{\alpha} =\displaystyle= (0,0,1C,0),andSα=1Φ(ω3A,0,0,A),\displaystyle\left(0,0,\frac{1}{C},0\right),\quad{\rm and}\quad S^{\alpha}=\frac{1}{\Phi}\left(\frac{\omega_{3}}{A},0,0,A\right)\,,

with Φ=A2R2+ω32\Phi=\sqrt{A^{2}R^{2}+\omega^{2}_{3}}.

With the above tetrad (II), we shall also define the corresponding directional derivative operators

f=Vααf,f=Kααf,f=Lααfandf=Sααf.f^{\bullet}=V^{\alpha}\partial_{\alpha}f,\;f^{{\dagger}}=K^{\alpha}\partial_{\alpha}f,\;f^{\ast}=L^{\alpha}\partial_{\alpha}f\;\mathrm{and}\;f^{\circ}=S^{\alpha}\partial_{\alpha}f. (2)

Additionally, the covariant derivatives of the orthonormal tetrad are

Vα;β=aαVβ+Ωαβ,V_{\alpha;\beta}=-a_{\alpha}V_{\beta}+\Omega_{\alpha\beta},
Kα;β=a1VαVβ+2Ω2V(αSβ)+(J1Kβ+J2Lβ)Lα+J6SαSβ,K_{\alpha;\beta}=-a_{1}V_{\alpha}V_{\beta}+2\Omega_{2}V_{(\alpha}S_{\beta)}+(J_{1}K_{\beta}+J_{2}L_{\beta})L_{\alpha}+J_{6}S_{\alpha}S_{\beta},
Lα;β=a2VαVβ+2Ω3V(αSβ)(J1Kβ+J2Lβ)Kα+J9SαSβ,L_{\alpha;\beta}=-a_{2}V_{\alpha}V_{\beta}+2\Omega_{3}V_{(\alpha}S_{\beta)}-(J_{1}K_{\beta}+J_{2}L_{\beta})K_{\alpha}+J_{9}S_{\alpha}S_{\beta},

and

Sα;β=2Ω2V(αKβ)2Ω3V(αLβ)(J6Kα+J9Lα)Sβ.S_{\alpha;\beta}=-2\Omega_{2}V_{(\alpha}K_{\beta)}-2\Omega_{3}V_{(\alpha}L_{\beta)}-(J_{6}K_{\alpha}+J_{9}L_{\alpha})S_{\beta}\,.

The scalars for the axisymmetric metric (1) are:

a1\displaystyle a_{1} =\displaystyle= A,rAB,a2=A,θAC,J1=B,θBC,J2=C,rBC,\displaystyle\frac{A_{,r}}{AB}\,,\,a_{2}=\frac{A_{,\theta}}{AC}\,,\quad J_{1}=-\frac{B_{,\theta}}{BC}\,,\,J_{2}=\frac{C_{,r}}{BC}\,, (3)
J6\displaystyle J_{6} =\displaystyle= a1+Φ,rBΦ,J9=a2+Φ,θCΦ,\displaystyle-a_{1}+\frac{\Phi_{,\,r}}{B\Phi}\,,\quad J_{9}=-a_{2}+\frac{\Phi_{,\theta}}{C\Phi}\,, (4)
Ω2\displaystyle\Omega_{2} =\displaystyle= A2ψ,r2BΦandΩ3=A2ψ,θ2CΦ,\displaystyle\frac{A^{2}\psi_{,r}}{2B\Phi}\quad{\rm and}\quad\Omega_{3}=\frac{A^{2}\psi_{,\theta}}{2C\Phi}\,, (5)

where ψ=ψ(r,θ)\psi=\psi(r,\theta) is a general function which allows us to define ω3=A2ψ\omega_{3}=A^{2}\psi.

The kinematical variables aαa_{\alpha} and Ωαβ\Omega_{\alpha\beta} can be written, in terms of the tetrad, as follows

aα\displaystyle a_{\alpha} =\displaystyle= a1Kα+a2Lαand,\displaystyle a_{1}K_{\alpha}+a_{2}L_{\alpha}\quad{\rm and}\,, (6)
Ωαβ\displaystyle\Omega_{\alpha\beta} =\displaystyle= Ω2(KαSβKβSα)+Ω3(LαSβLβSα),\displaystyle\Omega_{2}(K_{\alpha}S_{\beta}-K_{\beta}S_{\alpha})+\Omega_{3}(L_{\alpha}S_{\beta}-L_{\beta}S_{\alpha})\,, (7)

III Vacuum equations

Now, from the Ricci identities

e;β;ααe;α;βα=Rδβeδ,e(𝐕,𝐊,𝐋,𝐒)e^{\alpha}_{;\beta;\alpha}-e^{\alpha}_{;\alpha;\beta}=R_{\delta\beta}e^{\delta}\,,\qquad e\in(\mathbf{V},\mathbf{K},\mathbf{L},\mathbf{S}) (8)

we get

a1+a2+a1(a1+J2+J6)+a2(a2+J9J1)+2(Ω22+Ω32)=0,a_{1}^{\dagger}+a_{2}^{\ast}+a_{1}(a_{1}+J_{2}+J_{6})+a_{2}(a_{2}+J_{9}-J_{1})+2(\Omega_{2}^{2}+\Omega_{3}^{2})=0, (9)
Ω2+Ω2(2a1+J2)+Ω3+Ω3(2a2J1)=0,\Omega^{\dagger}_{2}+\Omega_{2}(2a_{1}+J_{2})+\Omega^{\ast}_{3}+\Omega_{3}(2a_{2}-J_{1})=0, (10)
J1(a1+J2+J6)+J1(a2+J9J1)a12J22J62+2Ω22=0,J_{1}^{\ast}-(a_{1}+J_{2}+J_{6})^{\dagger}+J_{1}(a_{2}+J_{9}-J_{1})-a^{2}_{1}-J_{2}^{2}-J_{6}^{2}+2\Omega_{2}^{2}=0, (11)
(a1+J6)+a2(a1J2)+J9(J6J2)2Ω2Ω3=0,(a_{1}+J_{6})^{\ast}+a_{2}(a_{1}-J_{2})+J_{9}(J_{6}-J_{2})-2\Omega_{2}\Omega_{3}=0, (12)
(a2+J9)+a1(a2+J1)+J1J6+J6J92Ω2Ω3=0,(a_{2}+J_{9})^{\dagger}+a_{1}(a_{2}+J_{1})+J_{1}J_{6}+J_{6}J_{9}-2\Omega_{2}\Omega_{3}=0, (13)
J2+(a2J1+J9)+a1J2+J2J6+a22+J12+J22+J922Ω32=0,J_{2}^{\dagger}+(a_{2}-J_{1}+J_{9})^{\ast}+a_{1}J_{2}+J_{2}J_{6}+a_{2}^{2}+J_{1}^{2}+J_{2}^{2}+J_{9}^{2}-2\Omega_{3}^{2}=0, (14)
J6+J9+J6(a1+J2+J6)+J9(a2J1+J9)2(Ω22+Ω32)=0,J_{6}^{\dagger}+J_{9}^{\ast}+J_{6}(a_{1}+J_{2}+J_{6})+J_{9}(a_{2}-J_{1}+J_{9})-2(\Omega_{2}^{2}+\Omega_{3}^{2})=0, (15)

The equation (9) can be written as

(a1CΦ),r+(a2BΦ),θ+2BCΦ(Ω22+Ω32)=0,(a_{1}C\Phi)_{,\,r}+(a_{2}B\Phi)_{,\theta}+2BC\Phi(\Omega^{2}_{2}+\Omega_{3}^{2})=0\,, (16)

likewise, the equation (15) turns

(J6CΦ),r+(J9BΦ),θ2BCΦ(Ω22+Ω32)=0,(J_{6}C\Phi)_{,\,r}+(J_{9}B\Phi)_{,\theta}-2BC\Phi(\Omega^{2}_{2}+\Omega_{3}^{2})=0\,, (17)

where the corresponding derivatives are denoted by

(),r()rand(),θ()θ.(\bullet)_{,\,r}\equiv\frac{\partial(\bullet)}{\partial r}\quad{\rm and}\quad(\bullet)_{,\,\theta}\equiv\frac{\partial(\bullet)}{\partial\theta}\,. (18)

Next, combining equations (16) and (17) we find the following equation for the function Φ\Phi

Φ,rr+f,rfΦ,r+1f2Φ,θθ=0.\Phi_{,\,r\,r}+\frac{f_{,\,r}}{f}\Phi_{,\,r}+\frac{1}{f^{2}}\Phi_{,\theta\theta}=0. (19)

Here, we have set, without loss of generality, that

C=Bf(r).C=Bf(r)\,. (20)

In the same way, the equation turns (12)

Φ,rBΦJ1Φ,θCΦJ2+a2a1+J9J6=2Ω2Ω3.\frac{\Phi_{,\,r}}{B\Phi}J_{1}-\frac{\Phi_{,\theta}}{C\Phi}J_{2}+a_{2}a_{1}+J_{9}J_{6}=2\Omega_{2}\Omega_{3}\,. (21)

Finally equation (10) can be re-written as follows

(Ω2A2C),r+(Ω3A2B),θ=0.(\Omega_{2}A^{2}C)_{,\,r}+(\Omega_{3}A^{2}B)_{,\theta}=0\,. (22)

IV General Solution

Let’s start integrating equation (19) assuming Φ=ΦR(r)ΦΘ(θ)\Phi~{}=~{}\Phi_{R}(r)\Phi_{\Theta}(\theta) separable, Thus we find three general solutions

Φ1=(c1cosμθ+c2sinμθ)1(r),\displaystyle\Phi_{1}=(c_{1}\cos{\mu\theta}+c_{2}\sin\mu\theta)\mathcal{F}_{1}(r)\,, (23)
Φ2=(c~1coshμθ+c~2sinhμθ)2(r)and\displaystyle\Phi_{2}=(\tilde{c}_{1}\cosh\mu\theta+\tilde{c}_{2}\sinh\mu\theta)\mathcal{F}_{2}(r)\quad{\rm and} (24)
Φ3=(c¯1θ+c¯2)3(r),\displaystyle\Phi_{3}=(\bar{c}_{1}\theta+\bar{c}_{2})\mathcal{F}_{3}(r)\,, (25)

where μ\mu is the separability constant with cic_{i}, c~i\tilde{c}_{i}, and c¯i\bar{c}_{i}, with i=1,2i=1,2, are integration constants. If ΦR(r)=f(r)\Phi_{R}(r)=f(r), the above three solutions for the equation (19) become particular solutions of the form

Φ1=(c1cosμθ+c2sinμθ)2μ2r2+c3r+c4,\displaystyle\Phi_{1}=(c_{1}\cos{\mu\theta}+c_{2}\sin\mu\theta)\sqrt{2\mu^{2}r^{2}+c_{3}r+c_{4}}\,\,, (26)
Φ2=(c~1coshμθ+c~2sinhμθ)2μ2r2+c~3r+c~4,\displaystyle\Phi_{2}=(\tilde{c}_{1}\cosh\mu\theta+\tilde{c}_{2}\sinh\mu\theta)\sqrt{-2\mu^{2}r^{2}+\tilde{c}_{3}r+\tilde{c}_{4}}\,\,, (27)
andΦ3=(c¯1θ+c¯2)(c¯3r+c¯4).\displaystyle{\rm and}\quad\Phi_{3}=(\bar{c}_{1}\theta+\bar{c}_{2})(\sqrt{\bar{c}_{3}r+\bar{c}_{4}})\,. (28)

again cic_{i}, c~i\tilde{c}_{i}, and c¯i\bar{c}_{i}, with i=1,2,3,4i=1,2,3,4, are integration constant.

The asymptotic behaviour of the metric rules out expressions (27) and (28), but when we do not have this physical restriction, Φ2\Phi_{2} type of solution is also possible. This is the case for the recently discussed examples of hyperbolically symmetric spacetimes HerreraWitten2018 ; HerreraDiPriscoOspino2021 , which will considered below.

Next, we observe that the general solution of (22) can be written as

Ω2=Ω,θ2A2CandΩ3=Ω,r2A2B,\Omega_{2}=\frac{\Omega_{,\theta}}{2A^{2}C}\qquad{\rm and}\qquad\Omega_{3}=-\frac{\Omega_{,\,r}}{2A^{2}B}, (29)

where Ω=Ω(r,θ)\Omega=\Omega(r,\theta) is an arbitrary function. Now, combining the definition (5) with (29) we get

ψ,rΩ,r+1f2ψ,θΩ,θ=0.\psi_{,\,r}\Omega_{,\,r}+\frac{1}{f^{2}}\psi_{,\theta}\Omega_{,\theta}=0\,. (30)

Introducing

Y=ACΦ,Y=\frac{AC}{\sqrt{\Phi}}\,, (31)

equation (21) becomes

Φ,rΦY,θY+Φ,θΦY,rY=2A,rA,θA2A4ψ,rψ,θ2Φ2.\frac{\Phi_{,\,r}}{\Phi}\frac{Y_{,\theta}}{Y}+\frac{\Phi_{,\theta}}{\Phi}\frac{Y_{,\,r}}{Y}=\frac{2A_{,\,r}A_{,\theta}}{A^{2}}-\frac{A^{4}\psi_{,\,r}\psi_{,\theta}}{2\Phi^{2}}\,. (32)

Taking into account the expression of Ω2\Omega_{2} and Ω3\Omega_{3} from (29) and (5), we can write the equation (16) as

(a1CΦ),r+(a2BΦ),θ+12(ψ,rΩ,θψ,θΩ,r)=0,(a_{1}C\Phi)_{,\,r}+(a_{2}B\Phi)_{,\theta}+\frac{1}{2}(\psi_{,\,r}\Omega_{,\theta}-\psi_{,\theta}\Omega_{,\,r})=0, (33)

or

(A,rAfΦ),r+(A,θAΦf),θ+12(ψ,rΩ,θψ,θΩ,r)=0,\left(\frac{A_{,r}}{A}f\Phi\right)_{,r}+\left(\frac{A_{,\theta}}{A}\frac{\Phi}{f}\right)_{,\theta}+\frac{1}{2}(\psi_{,r}\Omega_{,\theta}-\psi_{,\theta}\Omega_{,r})=0, (34)

The integration of (34) leads us to

A,rAfΦ+12ψΩ,θ\displaystyle\frac{A_{,r}}{A}f\Phi+\frac{1}{2}\psi~{}\Omega_{,\theta} =\displaystyle= χ~,θand\displaystyle\tilde{\chi}_{,\theta}\quad{\rm and} (35)
A,θAΦf12ψΩ,r\displaystyle\frac{A_{,\theta}}{A}\frac{\Phi}{f}-\frac{1}{2}\psi~{}\Omega_{,r} =\displaystyle= χ~,r\displaystyle-\tilde{\chi}_{,r} (36)

where χ~\tilde{\chi} is an arbitrary function.

Next, the integrability condition of the metric function AA, leads X~\tilde{X} to satisfy

(fχ~,rΦ),r+(χ~,θfΦ),θ=(fψΩ,r2Φ),r+(ψΩ,θ2fΦ),θ.\left(\frac{f\tilde{\chi}_{,r}}{\Phi}\right)_{,r}+\left(\frac{\tilde{\chi}_{,\theta}}{f\Phi}\right)_{,\theta}=\left(\frac{f\psi\Omega_{,r}}{2\Phi}\right)_{,r}+\left(\frac{\psi\Omega_{,\theta}}{2f\Phi}\right)_{,\theta}\,. (37)

On the other hand, combining the equations (5) and (29) we get a relation between the two arbitrary functions

ψ,r\displaystyle\psi_{,r} =\displaystyle= Ω,θΦA4f,\displaystyle\frac{\Omega_{,\theta}\Phi}{A^{4}f}, (38)
ψ,θ\displaystyle\psi_{,\theta} =\displaystyle= Ω,rfΦA4.\displaystyle-\frac{\Omega_{,r}f\Phi}{A^{4}}. (39)

Finally, taking into account equations (35)-(36), the integrability condition gives

(Ω,rfΦ),r+(Ω,θΦf),θ=4(χ~,θΩ,rχ~,rΩ,θ).(\Omega_{,r}f\Phi)_{,r}+\left(\Omega_{,\theta}\frac{\Phi}{f}\right)_{,\theta}=4(\tilde{\chi}_{,\theta}\Omega_{,r}-\tilde{\chi}_{,r}\Omega_{,\theta}). (40)

At this point, we can see, from equations (30)-(40), that the general solution of this system of equations depends on the choice of an arbitrary function Ω\Omega or χ~\tilde{\chi}.

IV.1 The method

The method we shall follow can be stated as

  1. 1.

    Integrate equation (19). If Φ\Phi is separable, then it is possible to obtain particular solutions of Φ\Phi in the form of (26)-28;

  2. 2.

    provide Ω\Omega, and from (30) obtain ψ\psi;

  3. 3.

    having Ω\Omega and ψ\psi we get the first metric function AA from equations (38) and (39);

  4. 4.

    with the definition ω3=A2ψ\omega_{3}=A^{2}\psi, we obtain the second metric coefficient, ω3\omega_{3};

  5. 5.

    calculate the third metric function R=Φ2ω32A2R=\sqrt{\frac{\Phi^{2}-\omega_{3}^{2}}{A^{2}}}, from the Φ\Phi-definition;

  6. 6.

    from equation (32) solve the forth metric coefficient CC;

  7. 7.

    finally, equation (20) leads to the last metric function BB.

In the next section, we shall illustrate the method with two particular selections of Ω\Omega. One leads to a general solution for the spherical static case, and the second recovers the Kerr metric.

IV.2 The Static Polar Solutions

For the static case, we choose

Ω=0,ψ=0\Omega=0,\quad\Rightarrow\quad\psi=0 (41)

and a particular solution of (26) with the form Φ=fsinθ=r22mrsinθ\Phi~{}=~{}f\sin\theta~{}=~{}\sqrt{r^{2}-2mr}\sin\theta. Thus, for X=lnAX=\ln A, equations (32) and (34) become

(X,rf2),rsinθ+(X,θsinθ),θ=0,(X_{,\,r}f^{2})_{,\,r}\sin\theta+(X_{,\theta}\sin\theta)_{,\theta}=0, (42)

and

f,rfY,θY+cosθsinθY,rY=2X,rX,θ,\frac{f_{,\,r}}{f}\frac{Y_{,\theta}}{Y}+\frac{\cos\theta}{\sin\theta}\frac{Y_{,\,r}}{Y}=2X_{,\,r}X_{,\theta}\,, (43)

respectively.

Additionally, it is easy to check that in the new coordinates

ρ=r22mrsinθandz=(rm)cosθ,\rho=\sqrt{r^{2}-2mr}\sin\theta\quad{\rm and}\quad z=(r-m)\cos\theta\,, (44)

equation (42) transforms into the well-known Weyl equation

Xρρ+1ρXρ+Xzz=0,X_{\rho\rho}+\frac{1}{\rho}X_{\rho}+X_{zz}=0\,, (45)

having a general solution written as follows

X=lnASP=χ~,θf2sinθ𝑑r+A~SP(θ).X=\ln A_{SP}=\int\frac{\tilde{\chi}_{,\theta}}{f^{2}\sin\theta}dr+\tilde{A}_{SP}(\theta)\,. (46)

Here, A~SP(θ)\tilde{A}_{SP}(\theta) is an arbitrary function while χ~\tilde{\chi} is any solution of the equation

χ~,rr+1f2(χ~,θθcotθχ~,θ)=0.\tilde{\chi}_{,\,rr}+\frac{1}{f^{2}}(\tilde{\chi}_{,\theta\theta}-\cot\theta\tilde{\chi}_{,\theta})=0\,. (47)

On the other hand, changing variables

η=lnf,andξ=lnsinθ,\eta=\ln f\,,\quad{\rm and}\quad\xi=\ln\sin\theta\,, (48)

the equation (43) can be written as

YηY+YξY=2XηXξ,\frac{Y_{\eta}}{Y}+\frac{Y_{\xi}}{Y}=2X_{\eta}X_{\xi}\,, (49)

having a general solution

Y=Exp[η0ηf~(χ~,ξη+χ~)𝑑χ~]Δ~(ξη),Y=Exp\left[\int_{\eta_{0}}^{\eta}\tilde{f}(\tilde{\chi},\xi-\eta+\tilde{\chi})d\tilde{\chi}\right]\tilde{\Delta}(\xi-\eta)\,, (50)

with f~=2XηXξ\tilde{f}=2X_{\eta}X_{\xi}, and Δ~\tilde{\Delta} an arbitrary integration function.

IV.3 The Kerr solution

For this case, we set

ΩK=2macosθr2+a2cos2θ.\Omega_{K}=\frac{2ma\cos\theta}{r^{2}+a^{2}\cos^{2}\theta}\,. (51)

From equation (40), we find

χ~K=m(a2+r2)cosθr2+a2cos2θ,\tilde{\chi}_{K}=-\frac{m(a^{2}+r^{2})\cos\theta}{r^{2}+a^{2}\cos^{2}\theta}\,, (52)

as a particular solution and from (30), we get

ψK=2marsin2θr22mr+a2cos2θ,\psi_{K}=\frac{2mar\sin^{2}\theta}{r^{2}-2mr+a^{2}\cos^{2}\theta}, (53)

Thus, the first metric coefficient emerges from equations (38) and (39) as

AK2=ΩK,θsinθψK,r=12mrr2+a2cos2θ.A^{2}_{K}=\sqrt{\frac{\Omega_{K,\theta}\sin\theta}{\psi_{K,r}}}=1-\frac{2mr}{r^{2}+a^{2}\cos^{2}\theta}\,. (54)

Next, the metric functions ω3K\omega_{3K} and RKR_{K} are calculated as follows

ω3K=AK2ψK=2marsin2θr2+a2cos2θ\omega_{3K}=A^{2}_{K}\psi_{K}=\frac{2mar\sin^{2}\theta}{r^{2}+a^{2}\cos^{2}\theta} (55)

and

RK=ΦK2ω3K2AK2=r2+a2+2a2mrsin2θr2+a2cos2θsinθ.R_{K}=\sqrt{\frac{\Phi^{2}_{K}-\omega_{3K}^{2}}{A^{2}_{K}}}=\sqrt{r^{2}+a^{2}+\frac{2a^{2}mr\sin^{2}\theta}{r^{2}+a^{2}\cos^{2}\theta}}\sin\theta\,. (56)

On the other hand, in this case, the right-hand side of equation (32) vanishes and its general solution can be written as

Φ,rΦY,θY+Φ,θΦY,rY=0,Y=F1(sinθf)\frac{\Phi_{,\,r}}{\Phi}\frac{Y_{,\theta}}{Y}+\frac{\Phi_{,\theta}}{\Phi}\frac{Y_{,\,r}}{Y}=0,\quad\Rightarrow Y=F_{1}\Big{(}\frac{\sin\theta}{f}\Big{)} (57)

where F1F_{1} is an arbitrary function of its argument.

Under the choice of

F1=fsinθa2sinθf,F_{1}=\sqrt{\frac{f}{\sin{\theta}}-\frac{a^{2}\sin\theta}{f}}\,, (58)

CKC_{K} is obtained from (31) as

CK=r2+a2cos2θ,C_{K}=\sqrt{r^{2}+a^{2}\cos^{2}\theta}\,, (59)

and in turn

BK=CKf=r2+a2cos2θr22mr+a2.B_{K}=\frac{C_{K}}{f}=\sqrt{\frac{r^{2}+a^{2}\cos^{2}\theta}{r^{2}-2mr+a^{2}}}\,. (60)

Now, as was expected, with equations 54-60 we obtain the Kerr metric In the case of the Kerr metric, it is

ds2=(12mrΛ)dt24marsin2θΛdtdϕ+Λr22mr+a2dr2+Λdθ2+sin2θ(r2+a2+2ma2rsin2θΛ)dϕ2,{\rm ds}^{2}=-\left(1-\frac{2mr}{\Lambda}\right){\rm d}t^{2}-\frac{4mar\sin^{2}\theta}{\Lambda}{\rm d}t{\rm d}\phi+\frac{\Lambda}{r^{2}-2mr+a^{2}}{\rm d}r^{2}+\Lambda{\rm d}\theta^{2}+\sin^{2}\theta\left(r^{2}+a^{2}+\frac{2ma^{2}r\sin^{2}\theta}{\Lambda}\right){\rm d}\phi^{2}\,, (61)

with Λ=r2+a2cos2θ\Lambda=r^{2}+a^{2}\cos^{2}\theta

IV.4 The Static hyperbolic solutions

Motivated by recent works in hyperbolic coordinates in the inner region r<2mr<2m HerreraWitten2018 ; HerreraDiPriscoOspino2021 ; HerreraDiPriscoOspino2021B ; HerreraEtal2023 ; HerreraEtal2020 , we shall apply the method to recover the hyperbolic qq-metric QuevedoToktarbayYerlan2013 .

Again, to study hyperbolic static solutions, we set

ΩSH=0,ψSH=0.\Omega_{SH}=0,\quad\Rightarrow\psi_{SH}=0\,. (62)

As a particular solution from equation (27) we choose Φ=f(r)sinhθ=r2+2mra2sinhθ\Phi=f(r)\sinh\theta=\sqrt{-r^{2}+2mr-a^{2}}\sinh\theta, and clearly X=ln(ASH)X=\ln(A_{SH}). Thus, considering ρ=r22mrsinhθ\rho=\sqrt{r^{2}-2mr}\sinh\theta, and z=(rm)coshθz=(r-m)\cosh\theta , equation (34) turns again into the hyperbolic Weyl equation

Xρρ+1ρXρ+Xzz=0.X_{\rho\rho}+\frac{1}{\rho}X_{\rho}+X_{zz}=0\,. (63)

Now, for the case of the qq- metric we have that by setting a2=0a_{2}=0 in (33), we obtain

ASH2(r)=(12mr)1+q,A_{SH}^{2}(r)=\Big{(}1-\frac{2m}{r}\Big{)}^{1+q}, (64)

where we have defined 1+q=α/2m1+q=-\alpha/2m.

To find CSHC_{SH}, the solution of equation (32) is given by

Y=F2(sinhθf)F1(η)Y=F_{2}\Big{(}\frac{\sinh\theta}{f}\Big{)}\equiv F_{1}(-\eta) (65)

where F1F_{1} is an arbitrary function given by

F2=1η(η2m2+1)q(q+2).F_{2}=\frac{1}{\sqrt{\eta\left(\eta^{2}m^{2}+1\right)^{q(q+2)}}}\,. (66)

Likewise,

CSH2=r2(12mr)q(m2sinh2(θ)2mrr2+1)q(q+2),C_{SH}^{2}=r^{2}\left(1-\frac{2m}{r}\right)^{-q}\left(\frac{m^{2}\sinh^{2}(\theta)}{2mr-r^{2}}+1\right)^{-q(q+2)}, (67)
BSH2=(12mr)q1(m2sinh2(θ)2mrr2+1)q(q+2),B_{SH}^{2}=\left(1-\frac{2m}{r}\right)^{-q-1}\left(\frac{m^{2}\sinh^{2}(\theta)}{2mr-r^{2}}+1\right)^{-q(q+2)}\,, (68)

and,

RSH2=r2(12mr)qsinh2(θ).R_{SH}^{2}=r^{2}\left(1-\frac{2m}{r}\right)^{-q}\sinh^{2}(\theta)\,. (69)

Considering equations (64) through (69) finally, the hyperbolic qq-metric is given by

ds2=(12mr)1+qdt2(12mr)q[(1+m2sinh2(θ)2mrr2)q(q+2)×((12mr)1dr2+r2dθ2)+r2sinh2θdϕ2],ds^{2}=-\left(1-\frac{2m}{r}\right)^{1+q}dt^{2}-\left(1-\frac{2m}{r}\right)^{-q}\left[\left(1+\frac{m^{2}\sinh^{2}(\theta)}{2mr-r^{2}}\right)^{-q(q+2)}\times\left(\left(1-\frac{2m}{r}\right)^{-1}dr^{2}+r^{2}d\theta^{2}\right)+r^{2}\sinh^{2}\theta d\phi^{2}\right], (70)

IV.5 Kerr solution in hyperbolic coordinates

To implement the Kerr solution in hyperbolic coordinates, we choose

ΩKH=2amcoshθr2+a2cosh2θ.\Omega_{KH}=\frac{2am\cosh\theta}{r^{2}+a^{2}\cosh^{2}\theta}. (71)

From equation 30 we get

ψKH=2marsinh2θ2mrr2a2cosh2θ.\psi_{KH}=\frac{2mar\sinh^{2}\theta}{2mr-r^{2}-a^{2}\cosh^{2}\theta}. (72)

Next, combining the equations (5) and (29) we obtain that the metric function AKH2A_{KH}^{2} is written as

AKH2=ΩKH,θsinhθψKH=2mrr2a2cosh2θr2+a2cosh2θ,A_{KH}^{2}=\sqrt{\frac{\Omega_{KH,\theta}\sinh\theta}{\psi^{\prime}_{KH}}}=\frac{2mr-r^{2}-a^{2}\cosh^{2}\theta}{r^{2}+a^{2}\cosh^{2}\theta}, (73)

and the metric functions ω3\omega_{3} and RR as follows

ω3KH=AKH2ψKH=2amrsinh2θr2+a2cosh2θ,\omega_{3KH}=A_{KH}^{2}\psi_{KH}=\frac{2amr\sinh^{2}\theta}{r^{2}+a^{2}\cosh^{2}\theta}, (74)

and

RKH=(r2+a22ma2rsinh2θr2+a2cosh2θ)sinhθR_{KH}=\sqrt{\left(r^{2}+a^{2}-\frac{2ma^{2}r\sinh^{2}\theta}{r^{2}+a^{2}\cosh^{2}\theta}\right)}\sinh\theta (75)

On the other hand, as in the case of the Kerr solution, the right side of (32) vanishes and its solution can be written as follows

YKH=fsinhθa2sinhθf,Y_{KH}=\sqrt{\frac{f}{\sinh{\theta}}-\frac{a^{2}\sinh\theta}{f}}, (76)

then

CKH=r2+a2cosh2θC_{KH}=\sqrt{r^{2}+a^{2}\cosh^{2}\theta} (77)

Finally,

BKH=r2+a2cosh2θ2mrr2a2B_{KH}=\sqrt{\frac{r^{2}+a^{2}\cosh^{2}\theta}{2mr-r^{2}-a^{2}}} (78)
ds2=(12mrΛKH)dt24marsinh2θΛKHdtdϕ+ΛKH2mrr2a2dr2+ΛKHdθ2+sinh2θ(r2+a22ma2rsinh2θΛKH)dϕ2,{\rm ds}^{2}=-\left(1-\frac{2mr}{\Lambda_{KH}}\right){\rm d}t^{2}-\frac{4mar\sinh^{2}\theta}{\Lambda_{KH}}{\rm d}t{\rm d}\phi+\frac{\Lambda_{KH}}{2mr-r^{2}-a^{2}}{\rm d}r^{2}+\Lambda_{KH}{\rm d}\theta^{2}+\sinh^{2}\theta\left(r^{2}+a^{2}-\frac{2ma^{2}r\sinh^{2}\theta}{\Lambda_{KH}}\right){\rm d}\phi^{2}\,, (79)

with ΛKH=r2+a2cosh2θ\Lambda_{KH}=r^{2}+a^{2}\cosh^{2}\theta

V Axially symmetric Killing tensor

In this section, we shall show that those axially symmetric space-times having a Killing tensor should have a more straightforward separable form for the metric component C2=C2(r,θ)C^{2}=C^{2}(r,\theta).

For the stationary axially symmetric space-time that admits a Killing tensor OspinoHernandezpastoraNunez2022 ξαβ\xi_{\alpha\beta} satisfying the Killing equation

ξαβ;μ+ξμα;β+ξβμ;α=0\xi_{\alpha\beta;\mu}+\xi_{\mu\alpha;\beta}+\xi_{\beta\mu;\alpha}=0 (80)

and written in terms of the tetrad vector is given by

ξαβ\displaystyle\xi_{\alpha\beta} =\displaystyle= ξ00VαVβ+ξ11KαKβ+ξ22LαLβ\displaystyle\xi_{00}V_{\alpha}V_{\beta}+\xi_{11}K_{\alpha}K_{\beta}+\xi_{22}L_{\alpha}L_{\beta} (81)
+\displaystyle+ ξ33SαSβ+ξ03(VαSβ+VβSα).\displaystyle\xi_{33}S_{\alpha}S_{\beta}+\xi_{03}(V_{\alpha}S_{\beta}+V_{\beta}S_{\alpha}).

Now, from the integration of the Killing tensor, we have the following

ξ11=F1(θ),ξ22=F2(r)ξ11ξ22=C2\xi_{11}=F_{1}(\theta),\quad\xi_{22}=F_{2}(r)\quad\Rightarrow\quad\xi_{11}-\xi_{22}=C^{2}

obtaining

C2=F1(θ)F2(r).C^{2}=F_{1}(\theta)-F_{2}(r). (82)

Also,

ξ33ξ11=Φ2A2g1(θ)ξ33ξ22=Φ2A2g2(r)}ξ22ξ11=Φ2A2(g1(θ)g2(r)).\left.\begin{array}[]{ccc}\xi_{33}-\xi_{11}&=&\frac{\Phi^{2}}{A^{2}}g_{1}(\theta)\\ \xi_{33}-\xi_{22}&=&\frac{\Phi^{2}}{A^{2}}g_{2}(r)\end{array}\right\}\,\,\xi_{22}-\xi_{11}=\frac{\Phi^{2}}{A^{2}}(g_{1}(\theta)-g_{2}(r)).

from where we have

A2C2=Φ2(g2(r)g1(θ))A^{2}C^{2}=\Phi^{2}(g_{2}(r)-g_{1}(\theta)) (83)

On the other hand,

ξ03=Φ2(g1(θ)ψ+C1(θ))ξ03=Φ2(g2(r)ψ+C2(r))}ψ=C2(r)C1(θ)g1(θ)g2(r),\left.\begin{array}[]{ccc}\xi_{03}&=&\Phi^{2}\left(g_{1}(\theta)\psi+C_{1}(\theta)\right)\\ \xi_{03}&=&\Phi^{2}\left(g_{2}(r)\psi+C_{2}(r)\right)\end{array}\right\}\,\,\psi=\frac{C_{2}(r)-C_{1}(\theta)}{g_{1}(\theta)-g_{2}(r)},

obtaining

ψ=C2(r)C1(θ)g1(θ)g2(r).\psi=\frac{C_{2}(r)-C_{1}(\theta)}{g_{1}(\theta)-g_{2}(r)}. (84)

Likewise, ξ00\xi_{00}, is given by

ξ00+F1(θ)=A2[g1(θ)(ψ+C1(θ)g1(θ))2+h1(θ)]ξ00+F2(r)=A2[g2(r)(ψ+C2(r)g2(r))2+h2(r)]\begin{array}[]{ccc}\xi_{00}+F_{1}(\theta)&=&A^{2}\left[g_{1}(\theta)(\psi+\frac{C_{1}(\theta)}{g_{1}(\theta)})^{2}+h_{1}(\theta)\right]\\ \xi_{00}+F_{2}(r)&=&A^{2}\left[g_{2}(r)(\psi+\frac{C_{2}(r)}{g_{2}(r)})^{2}+h_{2}(r)\right]\end{array}

from we found

C2A2\displaystyle\frac{C^{2}}{A^{2}} =\displaystyle= g1(θ)(ψ+C1(θ)g1(θ))2+h1(θ)\displaystyle g_{1}(\theta)\Big{(}\psi+\frac{C_{1}(\theta)}{g_{1}(\theta)}\Big{)}^{2}+h_{1}(\theta) (85)
\displaystyle- g2(r)(ψ+C2(r)g2(r))2h2(r)\displaystyle g_{2}(r)\Big{(}\psi+\frac{C_{2}(r)}{g_{2}(r)}\Big{)}^{2}-h_{2}(r)

In the following, we will examine the definitions for each integration function for the spherical and hyperbolic cases.

Now, for the spherical case and considering the asymptotic flatness boundary condition, from equations 83-85, we have that

g1(θ)=1sin2(θ),C1(θ)=0,g_{1}(\theta)=-\frac{1}{\sin^{2}(\theta)},\quad C_{1}(\theta)=0, (86)

also, without loss of generality we choose

F1(θ)=h1(θ)=b2sin(θ)2.F_{1}(\theta)=h_{1}(\theta)=-b^{2}\sin(\theta)^{2}. (87)

Under this, the expressions 82-85 turns

C2=b2sin(θ)2F2(r),C^{2}=-b^{2}\sin(\theta)^{2}-F_{2}(r), (88)
A2C2=f2(g2(r)sin(θ)2+1),A^{2}C^{2}=f^{2}(g_{2}(r)\sin(\theta)^{2}+1), (89)
ψ=C2(r)sin2(θ)g2(r)sin2(θ)+1,\psi=-\frac{C_{2}(r)\sin^{2}(\theta)}{g_{2}(r)\sin^{2}(\theta)+1}, (90)
C2A2\displaystyle\frac{C^{2}}{A^{2}} =\displaystyle= C22(r)sin2(θ)(g2(r)sin2(θ)+1)2b2sin2(θ)\displaystyle-\frac{C^{2}_{2}(r)\sin^{2}(\theta)}{(g_{2}(r)\sin^{2}(\theta)+1)^{2}}-b^{2}\sin^{2}(\theta) (91)
\displaystyle- C22(r)g2(r)(g2(r)sin(θ)2+1)2h2(r).\displaystyle\frac{C^{2}_{2}(r)}{g_{2}(r)(g_{2}(r)\sin(\theta)^{2}+1)^{2}}-h_{2}(r).

Next, combining equations 88,89,and 91, we get

(b4+f2b2g2(r))sin4(θ)+(2b2F2(r)+b2f2+h2(r)f2g2(r))sin2(θ)+F2(r)2+f2(C2(r)2g2(r)+h2(r))=0,(b^{4}+f^{2}b^{2}g_{2}(r))\sin^{4}(\theta)\\ +(2b^{2}F_{2}(r)+b^{2}f^{2}+h_{2}(r)f^{2}g_{2}(r))\sin^{2}(\theta)\\ +F_{2}(r)^{2}+f^{2}\Big{(}\frac{C_{2}(r)^{2}}{g_{2}(r)}+h_{2}(r)\Big{)}=0,

where it follows that

g2(r)=b2f2,g_{2}(r)=-\frac{b^{2}}{f^{2}}, (92)
h2(r)=2F2(r)+f2,h_{2}(r)=2F_{2}(r)+f^{2}, (93)
C2(r)=±b(F2(r)+f2)f2.C_{2}(r)=\pm\frac{b(F_{2}(r)+f^{2})}{f^{2}}. (94)

Now, as we can see each one of the last expressions depends on the function F2(r)F_{2}(r), and with this, 88-90 turns

A2=b2sin2(θ)f2b2sin2(θ)+F2(r),A^{2}=\frac{b^{2}\sin^{2}(\theta)-f^{2}}{b^{2}\sin^{2}(\theta)+F_{2}(r)}, (95)

and

ψ=±b(F2(r)+f2)sin2(θ)(b2sin2(θ)+f2),\psi=\pm\frac{b(F_{2}(r)+f^{2})\sin^{2}(\theta)}{(-b^{2}\sin^{2}(\theta)+f^{2})}, (96)

and

C2=b2sin2(θ)F2(r).C^{2}=-b^{2}\sin^{2}(\theta)-F_{2}(r). (97)

Now, for the Kerr metric, the F2(r)F_{2}(r) is defined by

F2(r)=a2r2,whereb2=a2.F_{2}(r)=-a^{2}-r^{2},\ \text{where}\ b^{2}=a^{2}. (98)

Likewise, motivated by the approach presented HerreraWitten2018 and the version of the Schwarzschild solution inside the horizon, we must consider this last solution as a limit case for the metric 79, when a=0a=0, and also well behaved when rr\rightarrow\infty.

g1H(θ)=1sinh2(θ),C1H(θ)=0,g_{1H}(\theta)=\frac{1}{\sinh^{2}(\theta)},\quad C_{1H}(\theta)=0, (99)
F1H(θ)=h1H(θ)=b2sinh(θ)2.F_{1H}(\theta)=h_{1H}(\theta)=b^{2}\sinh(\theta)^{2}. (100)

For this case, the expressions 83-85 turns

C2=b2sinh(θ)2F2H(r),C^{2}=b^{2}\sinh(\theta)^{2}-F_{2H}(r), (101)
A2C2=fH2(g2H(r)sinh(θ)21),A^{2}C^{2}=f_{H}^{2}(g_{2H}(r)\sinh(\theta)^{2}-1), (102)
ψH=C2H(r)sinh(θ)2(g2H(r)sinh(θ)21),\psi_{H}=-\frac{C_{2H}(r)\sinh(\theta)^{2}}{(g_{2H}(r)\sinh(\theta)^{2}-1)}, (103)
C2A2\displaystyle\frac{C^{2}}{A^{2}} =\displaystyle= C2H2(r)sinh(θ)2(g2H(r)sinh(θ)21)2+b2sinh(θ)2\displaystyle\frac{C^{2}_{2H}(r)\sinh(\theta)^{2}}{(g_{2H}(r)\sinh(\theta)^{2}-1)^{2}}+b^{2}\sinh(\theta)^{2} (104)
\displaystyle- C2H2(r)g2H(r)(g2H(r)sinh(θ)21)2h2H(r).\displaystyle\frac{C^{2}_{2H}(r)}{g_{2H}(r)(g_{2H}(r)\sinh(\theta)^{2}-1)^{2}}-h_{2H}(r).

Next, combining equations 101,102,and 104, we get

(b4fH2b2g2H)sinh(θ)4+(2b2F2H+b2fH2+h2HfH2g2H)sinh(θ)2+F2H2fH2(C2H(r)2g2H+h2H(r))=0,(b^{4}-f_{H}^{2}b^{2}g_{2H})\sinh(\theta)^{4}\\ +(-2b^{2}F_{2H}+b^{2}f_{H}^{2}+h_{2H}f_{H}^{2}g_{2H})\sinh(\theta)^{2}\\ +F_{2H}^{2}-f_{H}^{2}\Big{(}\frac{C_{2H(r)}^{2}}{g_{2H}}+h_{2H}(r)\Big{)}=0,

where it follows that

g2H(r)=b2fH2,g_{2H}(r)=\frac{b^{2}}{f_{H}^{2}}, (105)
h2H=2F2H(r)fH2,h_{2H}=2F_{2H}(r)-f_{H}^{2}, (106)
C2H(r)=±b(fH2F2H)fH2.C_{2H}(r)=\pm\frac{b(f_{H}^{2}-F_{2H})}{f_{H}^{2}}. (107)

Again, the expression depends on the function F2H(r)F_{2H}(r), and with this, 101-103 turns

A2=b2sinh(θ)2fH2b2sinh(θ)2F2H(r),A^{2}=\frac{b^{2}\sinh(\theta)^{2}-f_{H}^{2}}{b^{2}\sinh(\theta)^{2}-F_{2H}(r)}, (108)
C2=b2sinh(θ)2F2(r),C^{2}=b^{2}\sinh(\theta)^{2}-F_{2}(r), (109)

and

ψ=±b(fH2F2H(r))sinh(θ)2(b2sinh(θ)2fH2).\psi=\pm\frac{b(f_{H}^{2}-F_{2H}(r))\sinh(\theta)^{2}}{(b^{2}\sinh(\theta)^{2}-f_{H}^{2})}. (110)

Finally, for the case of the metric in 79

F2H=b2r2,whereb2=a2F_{2H}=-b^{2}-r^{2},\ \text{where}\ b^{2}=a^{2} (111)

On the other hand, for non-rotating configuration, we must recover the metric function related to the spherically static Schwarzschild solution.So, considering b2=a2=0b^{2}=a^{2}=0, we have the following

F1(θ)=h1(θ)=C2(r)=g2(r)=0,F_{1}(\theta)=h_{1}(\theta)=C_{2}(r)=g_{2}(r)=0, (112)

also,

C2=F2(r)=r2,C^{2}=-F_{2}(r)=r^{2}, (113)

and

A2=f2F2(r)=(12mr).A^{2}=\frac{-f^{2}}{F_{2}(r)}=\Big{(}1-\frac{2m}{r}\Big{)}. (114)

The remaining metric functions are obtained by the relationship between BB and CC and Φ\Phi with A2A^{2} and R2R^{2}.

VI Final Remarks

This work presents a method for solving the Einstein equations using the 1+3 tetrad formalism. This approach employs the orthogonal splitting of the Riemann tensor and the covariant derivatives of the tetrad vectors. It transforms the Einstein equations into a set of first-order scalar equations that can be systematically solved, particularly in vacuum, stationary, and axially symmetric systems.

We design a method to solve Einstein Equations for stationary axially symmetric sources. Assuming that the metric function Φ(r,θ)=A2R2+ω32\Phi(r,\theta)~{}=\sqrt{A^{2}R^{2}+\omega^{2}_{3}} is separable Φ(r,θ)=ΦR(r)Φ\Phi(r,\theta)=\Phi_{R}(r)\Phi, we integrated equation (19), leading to three distinct solutions: one in polar coordinates, one involving hyperbolic functions, and a linear solution. Next, specifying an arbitrary function Ω(r,θ)\Omega(r,\theta), we integrated the system (9)-(15), recovering the well-known Schwarzschild and Kerr solutions. We applied this methodology to the hyperbolic case, reobtaining the static Schwarzschild solution within the horizon in hyperbolic coordinates HerreraWitten2018 . Extending this approach, we obtained the stationary Kerr solution for analogous physical scenarios HerreraEtal2020 ; HerreraDiPriscoOspino2021 ; HerreraDiPriscoOspino2021B .

We also discuss the critical role of Killing tensors in identifying and understanding the symmetries of these spacetimes. We show that axially symmetric space-times with a Killing tensor should have a separable metric component C2=C2(r,θ)C^{2}=C^{2}(r,\theta). Killing tensors enable the determination of conserved quantities and simplify the integration of geodesic equations. In rotating and stationary spacetimes, these tensors are particularly valuable, facilitating the separation of variables in the Hamilton-Jacobi and geodesic equations OspinoHernandezpastoraNunez2022 . This capability results in tractable solutions for particle motion, which are crucial for analyzing phenomena such as gravitational lensing, accretion disk dynamics, and the trajectories of stars and compact objects.

This 1+3 methodology exhibits its robustness and versatility in addressing the complexities of Einstein’s field equations for axially symmetric spacetimes. The approach presented here introduces the gauge function, Ω=Ω(r,θ)\Omega=\Omega(r,\theta), enabling the derivation of axial stationary exterior solutions for hyperbolic spacetimes. This gauge freedom could also be applied to obtain other Kerr-like solutions where the nonrotating limits would not yield to the standard Schwarzschild scenario. This is ongoing work and will be reported in the future.

Acknowledgements.
L.A.N. acknowledges the financial sponsorship of the Vicerrectoría de Investigación y Extensión de la Universidad Industrial de Santander and Universidad de Salamanca through the research mobility programs. L.A.N. also thanks the hospitality of the Departamento de Matemáticas Aplicadas, Universidad de Salamanca. J.O. and J.L.H.P. express gratitude for financial support from Spain Ministerio de Ciencia, Innovación, (Programa Estatal de Generación de Conocimiento y Fortalecimiento Científico y Tecnológico del Sistema de I+D+i Grant number: PID2021-122938NB-I00, and Junta de Castilla y León, (Fondos Feder al 50%50\% y en línea con objetivos RIS3). Grant number: SA097P24.

References

  • [1] K. Akiyama, A. Alberdi, and EHT Collaboration. First m87 event horizon telescope results. iv. imaging the central supermassive black hole. The Astrophysical Journal Letters, 875(1):L4, 2019.
  • [2] K. Akiyama, A. Alberdi, and EHT Collaboration. First m87 event horizon telescope results. i. the shadow of the supermassive black hole. The Astrophysical Journal Letters, 875(1):L4, 2019.
  • [3] K. Akiyama, A. Alberdi, and EHT Collaboration. First sagittarius a* event horizon telescope results. vi. testing the black hole metric. The Astrophysical Journal Letters, 930(2):L17, 2022.
  • [4] The Event Horizon Telescope Collaboration. Mid-range science objectives for the event horizon telescope. arXiv preprint arXiv:2410.02986, 2024.
  • [5] V. Perlick and O.Y. Tsupko. Calculating black hole shadows: review of analytical studies. Physics Reports, 947:1–39, 2022.
  • [6] Z. Younsi, D. Psaltis, and F. Özel. Black hole images as tests of general relativity: effects of spacetime geometry. The Astrophysical Journal, 942(1):47, 2023.
  • [7] H. Quevedo. Multipole moments in general relativity—static and stationary vacuum solutions—. Fortschritte der Physik/Progress of Physics, 38(10):733–840, 1990.
  • [8] H. Stephani, D. Kramer, M. MacCallum, C. Hoenselaers, and E. Herlt. Exact solutions to einstein equations. Cambridge Monographs on Mathematical Physics, page 732, Jul 2006.
  • [9] H. Weyl. The theory of gravitation. Annalen Phys, 54:117, 1917.
  • [10] K. Schwarzschild. On the gravitational field of a sphere of incompressible fluid according to einstein’s theory. arXiv preprint physics/9912033, 1999.
  • [11] Ts. I. Gutsunayev and V.S. Manko. On the gravitational field of a mass possessing a multipole moment. General relativity and gravitation, 17:1025–1027, 1985.
  • [12] T.E. Denisova, Sh. A. Khakimov, and V.S. Manko. The gutsunaev-manko static vacuum solution. General relativity and gravitation, 26:119–123, 1994.
  • [13] R. P. Kerr. Gravitational field of a spinning mass as an example of algebraically special metrics. Physical review letters, 11(5):237, 1963.
  • [14] E.T. Newman and A.I. Janis. Note on the kerr spinning-particle metric. Journal of Mathematical Physics, 6(6):915–917, 1965.
  • [15] B. Carter. Hamilton-jacobi and schrodinger separable solutions of einstein’s equations. Commun. Math. Phys., 10:280, 1968.
  • [16] E. Newman, L. Tamburino, and T. Unti. Empty-space generalization of the schwarzschild metric. Journal of Mathematical Physics, 4(7):915–923, 1963.
  • [17] W. Kinnersley. Type d vacuum metrics. Journal of Mathematical Physics, 10(7):1195–1203, 1969.
  • [18] J.F. Plebanski and M. Demianski. Rotating, charged, and uniformly accelerating mass in general relativity. Annals of Physics, 98(1):98–127, 1976.
  • [19] F.J. Ernst. New formulation of the axially symmetric gravitational field problem. Physical Review, 167(5):1175, 1968.
  • [20] F.J. Ernst. New formulation of the axially symmetric gravitational field problem. ii. Physical Review, 168(5):1415, 1968.
  • [21] W. Kinnersley. Generation of stationary einstein-maxwell fields. Journal of Mathematical Physics, 14(5):651–653, 1973.
  • [22] W. Kinnersley. Symmetries of the stationary einstein–maxwell field equations. i. Journal of Mathematical Physics, 18(8):1529–1537, 1977.
  • [23] A. Tomimatsu and H. Sato. New exact solution for the gravitational field of a spinning mass. Physical Review Letters, 29(19):1344, 1972.
  • [24] A. Tomimatsu and H. Sato. New series of exact solutions for gravitational fields of spinning masses. Progress of Theoretical Physics, 50(1):95–110, 1973.
  • [25] M. Yamazaki and S. Hori. Generalization of the tomimatsu-sato solutions. Progress of Theoretical Physics, 57(2):696–697, 1977.
  • [26] M. Yamazaki. On the kerr and the tomimatsu-sato spinning mass solutions. Progress of Theoretical Physics, 57(6):1951–1957, 1977.
  • [27] M. Yamazaki. On the Kerr-Tomimatsu-Sato family of spinning mass solutions. Journal of Mathematical Physics, 18(12):2502–2508, December 1977.
  • [28] M. Yamazaki and S. Hori. Generalization of the tomimatsu sato solutions. Progress of Theoretical Physics, 60(4):1248–1248, 1978.
  • [29] C.M. Cosgrove. New family of exact stationary axisymmetric gravitational fields generalising the tomimatsu-sato solutions. Journal of Physics A: Mathematical and General, 10(9):1481, 1977.
  • [30] C.M. Cosgrove. Limits of the generalised tomimatsu-sato gravitational fields. Journal of Physics A: Mathematical and General, 10(12):2093, 1977.
  • [31] D. Maison. On the complete integrability of the stationary, axially symmetric einstein equations. Journal of Mathematical Physics, 20(5):871–877, 1979.
  • [32] R. Geroch. A method for generating solutions of einstein’s equations. Journal of Mathematical Physics, 12(6):918–924, 1971.
  • [33] R. Geroch. A method for generating new solutions of einstein’s equation. ii. Journal of Mathematical Physics, 13(3):394–404, 1972.
  • [34] W. Kinnersley and D. M. Chitre. Symmetries of the stationary Einstein-Maxwell field equations. II. Journal of Mathematical Physics, 18(8):1538–1542, August 1977.
  • [35] W. Kinnersley and D.M. Chitre. Symmetries of the stationary einstein–maxwell field equations. iii. Journal of Mathematical Physics, 19(9):1926–1931, 1978.
  • [36] W. Kinnersley and D. M. Chitre. Symmetries of the stationary Einstein-Maxwell equations. IV. Transformations which preserve asymptotic flatness. Journal of Mathematical Physics, 19(10):2037–2042, October 1978.
  • [37] W. Kinnersley and D. M. Chitre. Group Transformation That Generates the Kerr and Tomimatsu-Sato Metrics. Physical Review Letters, 40(25):1608–1610, June 1978.
  • [38] D.M. Zipoy. Topology of Some Spheroidal Metrics. Journal of Mathematical Physics, 7(6):1137–1143, June 1966.
  • [39] V.A. Belinskii and V.E. Zakharov. Integration of the einstein equations by means of the inverse scattering problem technique and construction of exact soliton solutions. Sov. Phys.-JETP, 48(6), 1978.
  • [40] N. R. Sibgatullin. Oscillations and waves in intense gravitational and electromagnetic fields. Moscow Izdatel Nauka, January 1984.
  • [41] B.K. Harrison. Bäcklund transformation for the ernst equation of general relativity. Physical Review Letters, 41(18):1197, 1978.
  • [42] B.K. Harrison. New large family of vacuum solutions of the equations of general relativity. Physical Review D, 21(6):1695–1697, March 1980.
  • [43] C. Hoenselaers, W. Kinnersley, and B.C. Xanthopoulos. Generation of asymptotically flat, stationary space-times with any number of parameters. Physical Review Letters, 42(8):481, 1979.
  • [44] C. Hoenselaers, W. Kinnersley, and B.C. Xanthopoulos. Symmetries of the stationary einstein–maxwell equations. vi. transformations which generate asymptotically flat spacetimes with arbitrary multipole moments. Journal of Mathematical Physics, 20(12):2530–2536, 1979.
  • [45] G. Neugebauer. Recursive calculation of axially symmetric stationary Einstein fields. Journal of Physics A Mathematical General, 13:1737–1740, May 1980.
  • [46] G. Neugebauer. A general integral of the axially symmetric stationary Einstein equations. Journal of Physics A Mathematical General, 13(2):L19–L21, February 1980.
  • [47] V. S. Manko, J. Martin, and E. Ruiz. On the simplest binary system of stationary black holes. Physics Letters A, 196(1-2):23–28, December 1994.
  • [48] V. S. Manko, J. Martín, E. Ruiz, N. R. Sibgatullin, and M. N. Zaripov. Metric of a rotating, charged, magnetized, deformed mass. Physical Review D, 49(10):5144–5149, May 1994.
  • [49] L. Herrera and V. S. Manko. Stationary solution of the Einstein equations possessing zero total angular momentum. Physics Letters A, 167(3):238–242, July 1992.
  • [50] E. Ruiz, V. S. Manko, and J. Martín. Extended N-soliton solution of the Einstein-Maxwell equations. Physical Review D, 51(8):4192–4197, April 1995.
  • [51] J. L. Hernandez-Pastora and J. Martin. New static axisymmetric solution of the Einstein field equations. Classical and Quantum Gravity, 10(12):2581–2585, December 1993.
  • [52] J. L. Hernández-Pastora and J. Martín. Monopole-quadrupole static axisymmetric solutions of Einstein field equations. General Relativity and Gravitation, 26(9):877–907, September 1994.
  • [53] J. Ospino, J.L. Hernandez-Pastora, and L.A. Nunez. All analytic solutions for geodesic motion in axially symmetric space-times. Eur. Phys. J. C, 2022.
  • [54] L. Herrera, A. Di Prisco, and J. Ospino. Dynamics of hyperbolically symmetric fluids. Symmetry, 13(9):1568, 2021.
  • [55] L Herrera, A Di Prisco, and J Ospino. Hyperbolically symmetric static fluids: A general study. Physical Review D, 103(2):024037, 2021.
  • [56] L. Herrera and L. Witten. An Alternative Approach to the Static Spherically Symmetric, Vacuum Global Solution to the Einstein Equations. Advances in High Energy Physics, 2018:1–5, 2018.
  • [57] L. Herrera, A. Di Prisco, J. Ospino, and J. Carot. Quasi-hyperbolically symmetric γ\gamma-metric. Entropy, 25(9):1338, 2023.
  • [58] L. Herrera, A. Di Prisco, J. Ospino, and L. Witten. Geodesics of the hyperbolically symmetric black hole. Physical Review D, 101(6):064071, 2020.
  • [59] H. Quevedo, S. Toktarbay, and A. Yerlan. Quadrupolar gravitational fields described by the qq- metric. arXiv preprint arXiv:1310.5339, 2013.