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

Static spherical vacuum solutions in the bumblebee gravity model

Rui Xu [email protected] Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, China    Dicong Liang Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, China    Lijing Shao [email protected] Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, China National Astronomical Observatories, Chinese Academy of Sciences, Beijing 100012, China
Abstract

The bumblebee gravity model is a vector-tensor theory of gravitation where the vector field nonminimally couples to the Ricci tensor. By investigating the vacuum field equations with spherical symmetry, we find two families of black-hole (BH) solutions in this model: one has a vanishing radial component of the vector field and the other has a vanishing radial component of the Ricci tensor. When the coupling between the vector field and the Ricci tensor is set to zero, the first family becomes the Reissner-Nordström solution while the second family degenerates to the Schwarzschild solution with the vector field being zero. General numerical solutions in both families are obtained for nonzero coupling between the vector field and the Ricci tensor. Besides BH solutions, we also reveal the existence of solutions that have a nonvanishing tttt-component of the metric on the supposed event horizon where the rrrr-component of the metric diverges while the curvature scalars are finite. These solutions are not supported by existing observations but present certain properties that are of academic interests. We conclude the study by putting the BH solutions into tests against the Solar-System observations and the images of supermassive BHs.

I Introduction

Einstein’s theory of general relativity (GR) and Maxwell’s theory of electromagnetism (EM) are the two representative classical field theories in physics. A straightforward combination of them, i.e., the Einstein-Maxwell theory, provides the simplest unification of the gravitational interaction and the EM interaction at the classical level, and serves as the prototype of vector-tensor theories in which vector fields are employed to complement the gravitational interaction described conventionally by the metric tensor. The action of the Einstein-Maxwell theory reads

S=d4xg(12κR14FμνFμν)+Sm,\displaystyle S=\int d^{4}x\sqrt{-g}\left(\frac{1}{2\kappa}R-\frac{1}{4}F^{\mu\nu}F_{\mu\nu}\right)+S_{\rm m}, (1)

where gg is the determinant of the metric gμνg_{\mu\nu}, the constant κ\kappa is 8πG8\pi G with GG being the gravitational constant, RR is the Ricci scalar, Fμν:=DμAνDνAμF_{\mu\nu}:=D_{\mu}A_{\nu}-D_{\nu}A_{\mu} is the field strength of the vector field AμA_{\mu} with DμD_{\mu} being the covariant derivative, and SmS_{\rm m} represents the action of matter. Without identifying the vector field as the EM vector potential, the action in Eq. (1) was first used by Will and Nordtvedt to illustrate possible preferred-frame effects in gravity Will and Nordtvedt (1972). The point is that if the vector field AμA_{\mu} possesses a nonzero background configuration, then the spacetime is anisotropic and hence has preferred frames.

Preferred frames violate Lorentz symmetry. The motivation of such a violation from string theory and the implications of it in low-energy effective field theories were first studied by Kostelecký and Samuel Kostelecký and Samuel (1989a, b). A generic framework, called the Standard-Model Extension (SME), has been developed to systematically incorporate all possible Lorentz-violating couplings into the actions of the Standard Model of particle physics and GR Colladay and Kostelecký (1997, 1998); Kostelecký (2004); Kostelecký and Mewes (2009, 2012, 2013); Kostelecký and Li (2019), and to calculate a number of predictions that can be tested in modern high-precision experiments Kostelecký and Russell (2011); Tasson (2016). The primary Lorentz-violating coupling in the gravitational sector of the SME takes the form sμνRμνs^{\mu\nu}R_{\mu\nu}, where sμνs^{\mu\nu} is a tensor field possessing a nonzero background configuration and RμνR_{\mu\nu} is the Ricci tensor. Because the background configuration of sμνs^{\mu\nu} defines preferred frames in general, this coupling violates Lorentz symmetry. The SME focuses on general properties and consequences caused by the coupling term, but asks little about the dynamics of the field sμνs^{\mu\nu}. To fill the gap between the SME and specific gravitational theories having the Lorentz-violating coupling, the action in Eq. (1) was generalized to Kostelecký (2004); Bailey and Kostelecký (2006)

S\displaystyle S =\displaystyle= d4xg(12κR+ξ2κBμBνRμν14BμνBμνV)\displaystyle\int d^{4}x\sqrt{-g}\left(\frac{1}{2\kappa}R+\frac{\xi}{2\kappa}B^{\mu}B^{\nu}R_{\mu\nu}-\frac{1}{4}B^{\mu\nu}B_{\mu\nu}-V\right) (2)
+Sm,\displaystyle+S_{\rm m},

where BμB^{\mu} is a dynamical vector field sometimes called the bumblebee field, and the generalized vector-tensor theory (2) is called the bumblebee gravity model Kostelecký (2004).

Apart from replacing the vector field AμA^{\mu} by BμB^{\mu} (correspondingly FμνF_{\mu\nu} by Bμν:=DμBνDνBμB_{\mu\nu}:=D_{\mu}B_{\nu}-D_{\nu}B_{\mu}), the important changes are adding (i) the coupling term BμBνRμνB^{\mu}B^{\nu}R_{\mu\nu} controlled by the constant ξ\xi to resemble the SME term sμνRμνs^{\mu\nu}R_{\mu\nu}, and (ii) the potential VV that takes its extremum when the bumblebee vector field BμB^{\mu} acquires a certain background configuration. By considering the linearized, perturbative bumblebee model where the deviation of the metric from the Minkowski metric ημν\eta_{\mu\nu} and the deviation of the bumblebee vector field from its background configuration are small, conclusions made in the general framework of the SME are verified by and contrasted with results in the bumblebee model Kostelecký (2004); Bluhm and Kostelecký (2005); Bluhm (2006); Bailey and Kostelecký (2006); Bluhm et al. (2008a, b); Liang et al. (2022).

Simply speaking, the bumblebee model is an essential example theory for studying Lorentz violation in gravity as its specified action provides information on the dynamics of the auxiliary field that breaks Lorentz symmetry, which is otherwise not contained in the general framework of the SME. But it is more than that. In fact, the action in Eq. (2) without the potential VV has been studied by Hellings and Nordtvedt (1973) as an alternative to GR in the context of the parametrized post-Newtonian (PPN) formalism as well as the cosmological solutions. At the end of their work, they brought up the proposal of identifying the vector field as the EM vector potential. A similar possibility was suggested by Bluhm and Kostelecký (2005) in the framework of the SME, where EM waves are the Nambu-Goldstone modes of the bumblebee vector field when Lorentz symmetry is spontaneously broken due to a background bumblebee field resulting from the potential VV. The idea of replacing the Einstein-Maxwell theory with the bumblebee model for unifying gravity and the EM theory is attractive.

Compared with the Einstein-Maxwell theory, the bumblebee model is expected to produce new and maybe even eccentric phenomena when the vector field interacts with gravity through the coupling term BμBνRμνB^{\mu}B^{\nu}R_{\mu\nu}. The gravitational field is optimally to be strong for such phenomena to be detectable. Therefore, strong-field solutions without applying the perturbation approach, as was done in the SME or the PPN formalism, should be considered. Black holes (BHs) are ideal strong-field systems to study, for not only the absence of matter simplifies the field equations, but also they have been detected with both gravitational waves (GWs) Abbott et al. (2019, 2021a, 2021b) and EM waves Akiyama et al. (2019, 2022a) so that unprecedented tests can be performed using the rapidly developing technology of multimessenger astronomy Akiyama et al. (2022b).

In the bumblebee model, an analytical solution that is very close to the Schwarzschild spacetime with the bumblebee field having only a nonvanishing radial component has been reported by Casana et al. (2018). In our work, we substantially extend the study and find spherical solutions with a nonvanishing temporal component of the bumblebee field. In fact, as suggested by the EM-like kinetic term in the action, the radial component of the bumblebee field is nondynamic, and can be eliminated from the final set of equations. It turns out there are two families of solutions: one solved from a group of two second-order ordinary differential equations (ODEs), and one solved from a group of three second-order ODEs. The presence of the temporal component of the bumblebee field is vital for both families of solutions as it alters the behavior of the metric near the BH event horizon radically. Most amazingly, it makes solutions with a nonvanishing gttg_{tt} at the event horizon possible.

We find that it is necessary to distinguish the BH solutions where gttg_{tt} vanishes at the event horizon from the solutions having a nonvanishing gttg_{tt} at the event horizon, for geodesics in the spacetime of the latter solutions remarkably bounce back on the event horizon. For this reason, those solutions with a nonvanishing gttg_{tt} at the event horizon are called compact hills (CHs) in our work. The bouncing-back behavior of geodesics has no experimental or observational evidence in gravity phenomena yet. Nevertheless, academically interesting features of the CH solutions are discussed, including the GW echoes Cardoso and Pani (2017).

For the BH solutions, we put them into tests against Solar-System observations and the shadow images of supermassive BHs achieved by the Event Horizon Telescope (EHT) Collaboration. As the two families of solutions have different numbers of free parameters, constraints on their parameter spaces are different. In one family, the BHs are characterized by two parameters, namely the mass and the bumblebee charge, so bounds on the bumblebee charge of the Sun and of the supermassive BHs are obtained. In the other family, the BHs have five free parameters where three of them are in the metric functions and two of them are in the bumblebee field. The Solar-System observations happen to only depend on the metric parameters for the solutions in this family, so bounds on the metric parameters are obtained while leaving the two bumblebee parameters unconstrained. The parameter space excluded in our work is limited, especially for the second family of solutions. Future study on the GWs in the spacetime of the BH solutions might yield tighter or complementary constraints.

The organization of this paper is as follows. We start with setting up the equations in Sec. II.1. Then an analytical solution is presented in Sec. II.2. General numerical solutions are discussed in Sec. II.3, and a detailed discussion about the CH solutions is made in Sec. II.4. In Sec. III.1, the BH solutions are tested by considering Solar-System observations, while in Sec. III.2, the test is done by considering the results from the shadow images of supermassive BHs. Finally, we summarize our findings and give a brief outlook for the directions worthy of further study in Sec. IV. The appendices list equations to supplement the main text.

In the remaining of the paper, equations are written in the geometrized unit system where G=c=1G=c=1. The sign convention of the metric is (,+,+,+)(-,+,+,+).

II Static spherical solutions in vacuum

II.1 Field equations

The field equations are obtained by taking variations with respect to the metric and the bumblebee field in Eq. (2). Under the assumption that the bumblebee field does not couple to conventional matter, the field equations can be written as

Gμν=κ(Tm)μν+κ(TB)μν,\displaystyle G_{\mu\nu}=\kappa\left(T_{\rm m}\right)_{\mu\nu}+\kappa\left(T_{B}\right)_{\mu\nu},
DμBμν2BνdVd(BλBλ)+ξκBμRμν=0,\displaystyle D^{\mu}B_{{\mu\nu}}-2B_{\nu}\frac{dV}{d(B^{\lambda}B_{\lambda})}+\frac{\xi}{\kappa}B^{\mu}R_{\mu\nu}=0, (3)

where (Tm)μν\left(T_{\rm m}\right)_{\mu\nu} is the energy-momentum tensor for conventional matter, and the contribution of the bumblebee field to the Einstein equations is

(TB)μν\displaystyle\left(T_{B}\right)_{\mu\nu} =\displaystyle= ξ2κ[gμνBαBβRαβ2BμBλRνλ2BνBλRμλg(BμBν)\displaystyle\frac{\xi}{2\kappa}\Big{[}g_{\mu\nu}B^{\alpha}B^{\beta}R_{\alpha\beta}-2B_{\mu}B_{\lambda}R_{\nu}^{\phantom{\nu}\lambda}-2B_{\nu}B_{\lambda}R_{\mu}^{\phantom{\mu}\lambda}-\Box_{g}(B_{\mu}B_{\nu}) (4)
gμνDαDβ(BαBβ)+DκDμ(BκBν)+DκDν(BμBκ)]\displaystyle-g_{{\mu\nu}}D_{\alpha}D_{\beta}(B^{\alpha}B^{\beta})+D_{\kappa}D_{\mu}\left(B^{\kappa}B_{\nu}\right)+D_{\kappa}D_{\nu}(B_{\mu}B^{\kappa})\Big{]}
+BμλBνλgμν(14BαβBαβ+V)+2BμBνdVd(BλBλ).\displaystyle+B_{\mu\lambda}B_{\nu}^{\phantom{\nu}\lambda}-g_{\mu\nu}\left(\frac{1}{4}B^{\alpha\beta}B_{\alpha\beta}+V\right)+2B_{\mu}B_{\nu}\frac{dV}{d(B^{\lambda}B_{\lambda})}.

The d’Alembertian in the curved spacetime is g:=gαβDαDβ\Box_{g}:=g^{\alpha\beta}D_{\alpha}D_{\beta}. Note that the potential VV has been assumed to be a function of the scalar product BμBμB^{\mu}B_{\mu}.

We are interested in nontrivial background configurations of the bumblebee field that are compatible with vacuum, namely (Tm)μν=0\left(T_{\rm m}\right)_{\mu\nu}=0. The potential VV is supposed to take extrema at these background configurations. Denoting Bμ=bμB_{\mu}=b_{\mu} as a background bumblebee field, we have

dVd(BλBλ)|Bμ=bμ=0.\displaystyle\frac{dV}{d(B^{\lambda}B_{\lambda})}\Big{|}_{B_{\mu}=b_{\mu}}=0. (5)

In addition, the value of VV at Bμ=bμB_{\mu}=b_{\mu} is equivalent to a cosmological constant. We drop it in the current study of BH solutions. Therefore, the vacuum field equations are simplified to

Gμν=κ(Tb)μν,\displaystyle G_{\mu\nu}=\kappa\left(T_{b}\right)_{\mu\nu},
Dμbμν+ξκbμRμν=0,\displaystyle D^{\mu}b_{{\mu\nu}}+\frac{\xi}{\kappa}b^{\mu}R_{\mu\nu}=0, (6)

where bμν=DμbνDνbμb_{\mu\nu}=D_{\mu}b_{\nu}-D_{\nu}b_{\mu}, and

(Tb)μν\displaystyle\left(T_{b}\right)_{\mu\nu} =\displaystyle= ξ2κ[gμνbαbβRαβ2bμbλRνλ2bνbλRμλg(bμbν)\displaystyle\frac{\xi}{2\kappa}\Big{[}g_{\mu\nu}b^{\alpha}b^{\beta}R_{\alpha\beta}-2b_{\mu}b_{\lambda}R_{\nu}^{\phantom{\nu}\lambda}-2b_{\nu}b_{\lambda}R_{\mu}^{\phantom{\mu}\lambda}-\Box_{g}(b_{\mu}b_{\nu}) (7)
gμνDαDβ(bαbβ)+DκDμ(bκbν)+DκDν(bμbκ)]\displaystyle-g_{{\mu\nu}}D_{\alpha}D_{\beta}(b^{\alpha}b^{\beta})+D_{\kappa}D_{\mu}\left(b^{\kappa}b_{\nu}\right)+D_{\kappa}D_{\nu}(b_{\mu}b^{\kappa})\Big{]}
+bμλbνλ14gμνbαβbαβ.\displaystyle+b_{\mu\lambda}b_{\nu}^{\phantom{\nu}\lambda}-\frac{1}{4}g_{\mu\nu}b^{\alpha\beta}b_{\alpha\beta}.

Focusing on static spherical solutions, we use the metric ansatz

ds2=e2νdt2+e2μdr2+r2(dθ2+sin2θdϕ2),\displaystyle ds^{2}=-e^{2\nu}dt^{2}+e^{2\mu}dr^{2}+r^{2}\left(d\theta^{2}+\sin^{2}\theta\,d\phi^{2}\right), (8)

and assume the background bumblebee field to be bα=(bt,br, 0, 0)b_{\alpha}=\left(b_{t},\,b_{r},\,0,\,0\right), where the unknowns, μ,ν,bt\mu\,,\nu,\,b_{t} and brb_{r}, are functions of rr. Then the Einstein field equations are nonvanishing for the diagonal components as well as the trtr-component, and the vector field equation has nonvanishing tt-component and rr-component. It turns out that the trtr-component of the Einstein field equations is guaranteed by the rr-component of the vector field equation, leaving 5 equations for the 4 unknowns, μ,ν,bt\mu\,,\nu,\,b_{t} and brb_{r}. So one of the equations depends on the others. These equations are displayed in Appendix A.

II.2 An analytical solution

Inspired by the solution obtained in Ref. Casana et al. (2018), we checked

ν=12ln(12Mr),\displaystyle\nu=\frac{1}{2}\ln{\left(1-\frac{2M}{r}\right)},
μ=μ012ln(12Mr),\displaystyle\mu=\mu_{0}-\frac{1}{2}\ln{\left(1-\frac{2M}{r}\right)},
bt=λ0+λ1r,\displaystyle b_{t}=\lambda_{0}+\frac{\lambda_{1}}{r}, (9)

against Eqs. (49) and (50), or equivalently Eq. (6) with the static spherical ansatz, to find that as long as

br2\displaystyle b_{r}^{2} =\displaystyle= e2μ0[1ξ(e2μ01)rr2Mκλ123ξM(r2M)\displaystyle e^{2\mu_{0}}\Bigg{[}\frac{1}{\xi}\frac{\left(e^{2\mu_{0}}-1\right)r}{r-2M}-\frac{\kappa\lambda_{1}^{2}}{3\xi M(r-2M)} (10)
+λ12(2rM)+6λ0λ1Mr+6λ02M2r3M(r2M)2],\displaystyle+\frac{\lambda_{1}^{2}(2r-M)+6\lambda_{0}\lambda_{1}Mr+6\lambda_{0}^{2}M^{2}r}{3M(r-2M)^{2}}\Bigg{]},

Eq. (9) is a solution to the field equations. It is an analytical solution characterized by 4 integral constants, μ0,M,λ0\mu_{0},\,M,\,\lambda_{0} and λ1\lambda_{1}. The 2-parameter solution obtained in Ref. Casana et al. (2018) simply corresponds to λ0=λ1=0\lambda_{0}=\lambda_{1}=0. We also notice that the 3-parameter solution corresponding to μ0=0\mu_{0}=0 has been previously obtained in Ref. Fan (2018).

The following two observations can be made based upon the solution given by Eqs. (9) and (10).

  1. 1.

    The Schwarzschild metric corresponding to μ0=0\mu_{0}=0 is no longer necessarily accompanied by a vanishing bumblebee field. Among the valid choices of λ0\lambda_{0} and λ1\lambda_{1} that give nonnegative br2b_{r}^{2}, an intriguing choice is λ1=2Mλ0\lambda_{1}=-2M\lambda_{0}, which simplifies Eq. (10) to

    br2\displaystyle b_{r}^{2} =\displaystyle= 23(12κξ)λ02Mr2M,\displaystyle\frac{2}{3}\left(1-\frac{2\kappa}{\xi}\right)\frac{\lambda_{0}^{2}M}{r-2M}, (11)

    for μ0=0\mu_{0}=0. As we require br20b_{r}^{2}\geq 0 outside the event horizon, this choice of λ1\lambda_{1} can only be made for ξ2κ\xi\geq 2\kappa. For the special case of ξ=2κ\xi=2\kappa, the Schwarzschild metric is a solution to the field equations together with a nontrivial bumblebee field that has only a temporal component bt12M/rb_{t}\propto 1-2M/r.

  2. 2.

    The Minkowski metric is recovered with μ0=0\mu_{0}=0 and M=0M=0. The limit of Eq. (10) exists if λ1M\lambda_{1}\propto\sqrt{M} while M0M\rightarrow 0. That is to say, the Minkowski metric can be accompanied by a nontrivial bumblebee field that has a radial component

    br2=13(2κξ)Cr,\displaystyle b_{r}^{2}=\frac{1}{3}\left(2-\frac{\kappa}{\xi}\right)\frac{C}{r}, (12)

    where CC being the ratio λ12/M\lambda_{1}^{2}/M has the dimension of length, and might represent the length scale of possible Lorentz violation in an underlying theory.

II.3 General numerical solutions

General solutions to Eqs. (49) and (50) are obtained numerically in two cases enumerated in Appendix A according to the two factors in the rr-component of the vector equation: (i) br=0b_{r}=0; (ii) Rrr=0R_{rr}=0. For the first case, the system of equations reduce to two coupled second-order ODEs, indicating a family of solutions determined by 4 parameters. For the second case, the system of equations reduce to three coupled second-order ODEs, indicating a family of solutions determined by 6 parameters. Both families of solutions turn out to have the asymptotic expansions

ν=n=1n=νnrn,\displaystyle\nu=\sum\limits_{n=1}^{n=\infty}\frac{\nu_{n}}{r^{n}},
μ=n=0n=μnrn,\displaystyle\mu=\sum\limits_{n=0}^{n=\infty}\frac{\mu_{n}}{r^{n}},
bt=n=0n=λnrn.\displaystyle b_{t}=\sum\limits_{n=0}^{n=\infty}\frac{\lambda_{n}}{r^{n}}. (13)

After substituting them into the ODEs to find the recurrence relations for the expansion coefficients νn,μn\nu_{n},\,\mu_{n} and λn\lambda_{n}, these expansions can be utilized to assign the initial values to start the numerical integrations from large enough rr. Note that a constant term in the expansion of ν\nu is unnecessary as it amounts to a change of the time coordinate. Therefore, the number of parameters characterizing the solutions is actually three for the first family (br=0b_{r}=0) and 5 for the second family (Rrr=0R_{rr}=0).

The recurrence relations for the first few coefficients are given in Appendix B. The conclusion is that for the case of br=0b_{r}=0, the solutions must have μ0=0\mu_{0}=0 and are fixed once the three free coefficients μ1,λ0\mu_{1},\,\lambda_{0} and λ1\lambda_{1} are specified, and that for the case of Rrr=0R_{rr}=0, there are 5 free coefficients, μ0,μ1,μ2,λ0\mu_{0},\,\mu_{1},\,\mu_{2},\,\lambda_{0} and λ1\lambda_{1}, describing the solutions. We point out that the analytical solution given by Eqs. (9) and (10) belongs to the second case and is recovered when taking μ2=μ12=M2\mu_{2}=\mu_{1}^{2}=M^{2}.

Before further discussing the numerical solutions, let us relate the parameters μ1\mu_{1} and λ1\lambda_{1} to the mass and the bumblebee charge of the spacetime. With μ0=0\mu_{0}=0, the Arnowitt-Deser-Misner (ADM) mass of the spacetime can be calculated and it is μ1\mu_{1} Arnowitt et al. (2008). However, the definition of the ADM mass does not apply to the solutions with μ00\mu_{0}\neq 0, when the metric is not asymptotically flat. Therefore we adopt the Komar mass as a measure of the spacetime mass when μ00\mu_{0}\neq 0. The Komar mass is defined by the conserved current associated with the time-translation Killing vector Kμ=(1,0,0,0)K^{\mu}=(1,0,0,0). The conserved current is Poisson (2009)

JMμ:=KνRμν=DνDμKν,\displaystyle J_{M}^{\mu}:=K_{\nu}R^{\mu\nu}=D_{\nu}D^{\mu}K^{\nu}, (14)

where the curvature identity [Dμ,Dν]Kα=RλμναKλ[D_{\mu},D_{\nu}]K^{\alpha}=R^{\alpha}_{\phantom{\alpha}\lambda{\mu\nu}}K^{\lambda} and the Killing equation DμKν+DνKμ=0D_{\mu}K_{\nu}+D_{\nu}K_{\mu}=0 have been used. The Komar mass then can be calculated as

MK\displaystyle M_{\rm K} :=\displaystyle:= 14πd3xgJMt\displaystyle-\frac{1}{4\pi}\int d^{3}x\sqrt{-g}\,J_{M}^{t} (15)
=\displaystyle= 14πlimr𝑑θ𝑑ϕgDtKr\displaystyle-\frac{1}{4\pi}\lim\limits_{r\rightarrow\infty}\iint d\theta d\phi\sqrt{-g}\,D^{t}K^{r}
=\displaystyle= eμ0μ1.\displaystyle e^{-\mu_{0}}\mu_{1}.

It is evident that the Komar mass reduces to the ADM mass when μ0=0\mu_{0}=0. The bumblebee charge can be defined in a similar way. Instead of using a Killing vector, the current

JQμ:=ξκbνRμν=Dνbνμ\displaystyle J_{Q}^{\mu}:=\frac{\xi}{\kappa}b_{\nu}R^{\mu\nu}=-D_{\nu}b^{\nu\mu} (16)

is automatically conserved as bμνb_{\mu\nu} is antisymmetric. Therefore the bumblebee charge having the dimension of length can be defined as

Q\displaystyle Q :=\displaystyle:= 14πκ2d3xgJQt\displaystyle-\frac{1}{4\pi}\sqrt{\frac{\kappa}{2}}\int d^{3}x\sqrt{-g}\,J_{Q}^{t} (17)
=\displaystyle= 14πκ2limr𝑑θ𝑑ϕgbtr\displaystyle-\frac{1}{4\pi}\sqrt{\frac{\kappa}{2}}\lim\limits_{r\rightarrow\infty}\iint d\theta d\phi\sqrt{-g}\,b^{tr}
=\displaystyle= κ2eμ0λ1,\displaystyle\sqrt{\frac{\kappa}{2}}\,e^{-\mu_{0}}\lambda_{1},

where the factor κ/2\sqrt{\kappa/2} has been chosen so that the Reissner-Nordström metric in Eq. (54) can be recovered when ξ=0\xi=0.

Back to the numerical results, as we integrate the system of ODEs from a large radius to r=0r=0, there exist three types of solutions. The one that we are interested in has grr=e2μg_{rr}=e^{2\mu} diverging at a certain radius r=rhr=r_{h}. The other two types that we will ignore are solutions of naked singularity and solutions with gtt=e2νg_{tt}=-e^{2\nu} diverging at a finite radius. Concentrating on the interesting solutions that have grrg_{rr} diverging at a finite radius r=rhr=r_{h}, the BH solutions are selected with two extra conditions: (i) vanishing gttg_{tt} at rhr_{h} and (ii) finite curvature scalars at rhr_{h}. It turns out that the finiteness of the curvature scalars at rhr_{h} is trivially satisfied by all the solutions with grrg_{rr} diverging at rhr_{h}. But among these solutions, it turns out that not all of them have gtt=0g_{tt}=0 at rhr_{h}. In fact, for the solutions in the first family (br=0b_{r}=0), most of them have nonvanishing gttg_{tt} at rhr_{h}.

To comprehend what happens near rhr_{h}, we seek expansions of the variables in terms of δ:=rrh\delta:=r-r_{h}. It is after a careful investigation of the numerical solutions that we find the variables taking the following expansions near rhr_{h},

gtt=e2ν=n=0N1nδnδn=0N2nδn,\displaystyle g_{tt}=-e^{2\nu}=-\sum\limits_{n=0}^{\infty}N_{1n}\,\delta^{n}-\sqrt{\delta}\sum\limits_{n=0}^{\infty}N_{2n}\,\delta^{n},
grr=e2μ=1δn=0M1nδn+1δn=0M2nδn,\displaystyle g_{rr}=e^{2\mu}=\frac{1}{\delta}\sum\limits_{n=0}^{\infty}M_{1n}\,\delta^{n}+\frac{1}{\sqrt{\delta}}\sum\limits_{n=0}^{\infty}M_{2n}\,\delta^{n},
bt=n=0L1nδn+δn=0L2nδn,\displaystyle b_{t}=\sum\limits_{n=0}^{\infty}L_{1n}\,\delta^{n}+\sqrt{\delta}\sum\limits_{n=0}^{\infty}L_{2n}\,\delta^{n}, (18)

where {N1n,N2n,M1n,M2n,L1n,L2n}\big{\{}N_{1n},\,N_{2n},\,M_{1n},\,M_{2n},L_{1n},\,L_{2n}\big{\}} is the set of expansion coefficients. The recurrence relations can be obtained by substituting the expansions into the ODEs and setting each order of δ\delta to zero. The recurrence relations for the first few coefficients are given in Appendix C. The most important fact we learn is that the coefficient N10N_{10} is closely related to the coefficients for the terms involving δ\sqrt{\delta} in the following way. If N10=0N_{10}=0, then N2n=M2n=L2n=0N_{2n}=M_{2n}=L_{2n}=0, and vice versa.

The solutions with grrg_{rr} diverging at a finite radius rhr_{h}, namely those can be expanded in the form of Eq. (18), thus can be divided into two categories according to whether N10N_{10} is zero. The BH solutions have N10=0N_{10}=0, while the solutions with N100N_{10}\neq 0, for their unexpected property of bouncing geodesics at rhr_{h}, which will be discussed in Sec. II.4, will be dubbed CHs in the remaining of the paper. Note that both the BH solutions and the CH solutions have finite curvature scalars at rhr_{h} (see Appendix D).

Refer to caption
Figure 1: Contour plots for rhr_{h} on the λ0\lambda_{0}-λ1\lambda_{1} plane for the solutions in the first family (br=0b_{r}=0). The asymptotic parameter μ1\mu_{1} is used as the length unit in the geometrized unit system. The blank region in the first plot consists of uninteresting solutions, namely naked singularities and solutions with singular gttg_{tt}. Notice that when ξ=2κ\xi=2\kappa, metric solutions are identical given λ0\lambda_{0} and λ1\lambda_{1} on a line λ1=2λ0+constant\lambda_{1}=-2\lambda_{0}+{\rm constant}. The line λ1=2λ0\lambda_{1}=-2\lambda_{0} gives the Schwarzschild metric specially.
Refer to caption
Figure 2: Contour plots for rhr_{h} on the λ0\lambda_{0}-λ1\lambda_{1} plane for the solutions in the second family (Rrr=0R_{rr}=0). We have taken μ0=0\mu_{0}=0 and used μ1\mu_{1} as the length unit in the geometrized unit system. The blank regions in the two plots in the first row consist of naked singularities. The red lines are boundaries of the inequality (58), which is a requirement of br20b_{r}^{2}\geq 0 at infinity. Among the 5 plots with the red lines, the region for the inequality to hold contains the origin λ0=λ1=0\lambda_{0}=\lambda_{1}=0 only for the second plot in the first row.

Figures 1 and 2 show how rhr_{h} depends on the asymptotic parameters λ0\lambda_{0} and λ1\lambda_{1}. We take the asymptotic parameter μ1\mu_{1} as the length unit in the numerical calculations. For the first family of solutions (br=0b_{r}=0), a solution is obtained once the values of the coupling constant ξ\xi and the asymptotic parameters λ0\lambda_{0} and λ1\lambda_{1} are specified properly. Sufficient amount of solutions obtained by varying λ0\lambda_{0} and λ1\lambda_{1} then generate the contour plots of rhr_{h} on the λ0\lambda_{0}-λ1\lambda_{1} plane as shown in Fig. 1. For the second family of solutions (Rrr=0R_{rr}=0), with μ1\mu_{1} being the unit, a solution is obtained when the asymptotic parameters μ0,μ2,λ0,λ1\mu_{0},\,\mu_{2},\,\lambda_{0},\,\lambda_{1}, as well as the coupling constant ξ\xi are specified properly. To generate Fig. 2, we have taken μ0=0\mu_{0}=0 and two representative values for μ2\mu_{2}. Solutions with nonzero but small μ0\mu_{0} turn out to produce similar shapes of the contour plots, while solutions with μ0\mu_{0} deviating from zero significantly are uninteresting to us as μ0\mu_{0} has been constrained to be smaller than 101210^{-12} using the advance of perihelia of the Solar-System planets in Ref. Casana et al. (2018).

Figures 3 and 4 show how N10N_{10} depends on λ0\lambda_{0} and λ1\lambda_{1}, using the same solutions producing Figs. 1 and 2. The values of the asymptotic parameters λ0\lambda_{0} and λ1\lambda_{1} for BH solutions and for CH solutions are conveniently revealed in Figs. 3 and 4. For the first family of solutions, the requirement of N10=0N_{10}=0 turns out to force L10L_{10} to vanish, which acts as an extra constraint on the asymptotic parameters λ0\lambda_{0} and λ1\lambda_{1}. Therefore, the BH solutions in the first family only exist along a single line on the λ0\lambda_{0}-λ1\lambda_{1} plane for a given ξ\xi, leaving the majority of the valid λ0\lambda_{0} and λ1\lambda_{1} to generate the CH solutions characterized by N100N_{10}\neq 0. For the second family of solutions, the requirement of N10=0N_{10}=0 forces no extra constraint on λ0\lambda_{0} and λ1\lambda_{1}, so the BH solutions exist on the two-dimensional plane of λ0\lambda_{0} and λ1\lambda_{1}. In fact, as shown in Fig. 4, there is no CH solution for any values of λ0\lambda_{0} and λ1\lambda_{1} as long as one takes μ2μ12\mu_{2}\leq\mu_{1}^{2}.

Refer to caption
Figure 3: Contour plots for N10=gtt(rh)N_{10}=-g_{tt}(r_{h}) on the λ0\lambda_{0}-λ1\lambda_{1} plane for the solutions in the first family (br=0b_{r}=0). The BH solutions, corresponding to N10=0N_{10}=0, are located along the brown lines. The colored contours with N10>0N_{10}>0 are solutions dubbed “CHs” in our work.
Refer to caption
Figure 4: Contour plots for N10=gtt(rh)N_{10}=-g_{tt}(r_{h}) on the λ0\lambda_{0}-λ1\lambda_{1} plane for the solutions in the second family (Rrr=0R_{rr}=0). The BH solutions, corresponding to N10=0N_{10}=0, form the regions colored in brown. We have checked that our numerical integrations have errors less than 5×1055\times 10^{-5}. The contours in other colors represent solutions with N10>0N_{10}>0. These solutions, dubbed “CHs”, do not exist for ξ=2κ\xi=2\kappa or when μ2μ12\mu_{2}\leq\mu_{1}^{2}. The red lines have the same meaning as in Fig. 2.

Comparing the BH solutions in the first family (br=0b_{r}=0) with the charged BH solution in GR, i.e., the Reissner-Nordström metric, is worthwhile. The asymptotic expansions of the metric functions μ\mu and ν\nu given by the recurrence relations in Eq. (52) include the Reissner-Nordström metric as a special case for ξ=0\xi=0, when the bumblebee model recovers the Einstein-Maxwell theory. This suggests that the bumblebee BH solutions in the first family, characterized by the two parameters, M=μ1M=\mu_{1} and Q=κ/2λ1Q=\sqrt{\kappa/2}\,\lambda_{1}, can be viewed as a generalization of the Reissner-Nordström metric. A plot of rhr_{h} versus QQ in unit of μ1\mu_{1} as shown in Fig. 5 demonstrates how the family of solutions change with the coupling constant ξ\xi. We would like to direct the readers’ attention to two special values of ξ\xi, namely ξ=0\xi=0 and ξ=2κ\xi=2\kappa. When ξ<0\xi<0, the maximal charge is less than MM, and when ξ>0\xi>0, the maximal charge is greater than MM. When ξ<2κ\xi<2\kappa, the maximal radius of the event horizon is 2M2M (Schwarzschild metric), and when ξ>2κ\xi>2\kappa, the maximal radius of the event horizon is greater than 2M2M. We notice that when ξ>2κ\xi>2\kappa or ξ<0\xi<0 there exist two black-hole solutions with the same rhr_{h} but different QQ. The two black holes are certainly observationally distinguishable as the difference in QQ makes their metric functions different. In fact, when we calculate the shadow of the black holes in Sec. III.2, we see that they have different shadow radii (see Fig. 10). A perhaps more interesting but puzzling observation is that when ξ=κ\xi=\kappa, rhr_{h} has two values given QQ near its maximum. When this happens, we suspect that the black hole with the smaller rhr_{h} might be unstable. Further study on quasi-normal modes of the black holes found here will shed light on this point.

We end the general discussion on the solutions by tabulating several representative numerical solutions for further reference. Table 1 lists 4 BH solutions and 4 CH solutions in the first family. Table 2 lists 8 BH solutions and 3 CH solutions in the second family. To have a brief picture about how different from the Schwarzschild metric these solutions are, the metric functions, as well as the scalar b2:=gμνbμbνb^{2}:=g^{\mu\nu}b_{\mu}b_{\nu} and the Ricci scalar RR, are plotted in Fig. 6, for several of the representative solutions.

Refer to caption
Figure 5: Event horizon rhr_{h} versus bumblebee charge QQ for the BH solutions in the first family (br=0b_{r}=0). The unit for rhr_{h} and QQ is μ1\mu_{1}, which is the mass of the BH.
Table 1: Representative solutions in the first family (br=0b_{r}=0). The asymptotic parameter μ1\mu_{1} is used as the length unit in the geometrized unit system. In the columns of “Event horizon behavior”, the leading terms of the expansions in Eq. (18) are given.
Solution ξ\xi Asymptotic parameters Event horizon behavior
λ0\lambda_{0} λ1\lambda_{1} rhr_{h} gtt-g_{tt} grrg_{rr} btb_{t}
BH-Ia κ-\kappa 0.13890.1389 0.1974-0.1974 1.7191.719 0.4913δ0.4913\,\delta 3.500δ\frac{3.500}{\delta} 0.1254δ0.1254\,\delta
CH-Ia κ-\kappa 0.20.2 0.25-0.25 1.0791.079 0.046750.05157δ0.04675-0.05157\sqrt{\delta} 0.6751δ0.05777δ\frac{0.6751}{\delta}-\frac{0.05777}{\sqrt{\delta}} 0.07031+0.07584δ-0.07031+0.07584\sqrt{\delta}
BH-Ib κ\kappa 0.25680.2568 0.3637-0.3637 0.61680.6168 0.6754δ0.6754\,\delta 0.8299δ\frac{0.8299}{\delta} 0.2454δ0.2454\,\delta
CH-Ib κ\kappa 0.20.2 0.25-0.25 2.6242.624 0.1839+0.2205δ0.1839+0.2205\sqrt{\delta} 0.4535δ+1.080δ\frac{0.4535}{\delta}+\frac{1.080}{\sqrt{\delta}} 0.09549+0.03323δ0.09549+0.03323\sqrt{\delta}
BH-Ic 2κ2\kappa λ0\lambda_{0} 2λ0-2\lambda_{0} 22 12δ\frac{1}{2}\,\delta 2δ\frac{2}{\delta} λ02δ\frac{\lambda_{0}}{2}\,\delta
CH-Ic 2κ2\kappa 0.20.2 0.25-0.25 3.0733.073 0.1742+0.2380δ0.1742+0.2380\sqrt{\delta} 0.7683δ+1.050δ\frac{0.7683}{\delta}+\frac{1.050}{\sqrt{\delta}} 0.08396+0.04761δ0.08396+0.04761\sqrt{\delta}
BH-Id 3κ3\kappa 0.20550.2055 0.3813-0.3813 2.1682.168 0.4878δ0.4878\,\delta 1.419δ\frac{1.419}{\delta} 0.1112δ0.1112\,\delta
CH-Id 3κ3\kappa 0.20.2 0.25-0.25 2.8802.880 0.1324+0.2289δ0.1324+0.2289\sqrt{\delta} 0.8226δ+1.091δ\frac{0.8226}{\delta}+\frac{1.091}{\sqrt{\delta}} 0.06732+0.05483δ0.06732+0.05483\sqrt{\delta}
Table 2: Representative solutions in the second family (Rrr=0R_{rr}=0). We have taken μ0=0\mu_{0}=0 and set μ1\mu_{1} as the length unit. In the columns of “Event horizon behavior”, the leading terms of the expansions in Eq. (18) are given.
Solution ξ\xi Asymptotic parameters Event horizon behavior
μ2\mu_{2} λ0\lambda_{0} λ1\lambda_{1} rhr_{h} gtt-g_{tt} grrg_{rr} btb_{t}
BH-IIa κ-\kappa 0.90.9 0.20.2 0.25-0.25 1.6801.680 0.1602δ0.1602\,\delta 0.7278δ\frac{0.7278}{\delta} 0.04205+0.07761δ0.04205+0.07761\,\delta
BH-IIb κ-\kappa 1.11.1 0.20.2 0.25-0.25 2.2222.222 0.6537δ0.6537\,\delta 2.699δ\frac{2.699}{\delta} 0.09487+0.03065δ0.09487+0.03065\,\delta
CH-IIa κ-\kappa 1.11.1 0.10.1 0.25-0.25 2.6032.603 0.08048+0.2286δ0.08048+0.2286\sqrt{\delta} 0.5266δ+1.351δ\frac{0.5266}{\delta}+\frac{1.351}{\sqrt{\delta}} 0.001186+0.01286δ0.001186+0.01286\sqrt{\delta}
BH-IIc κ\kappa 0.90.9 0.20.2 0.25-0.25 1.9221.922 0.4691δ0.4691\,\delta 1.841δ\frac{1.841}{\delta} 0.07564+0.05401δ0.07564+0.05401\,\delta
BH-IId κ\kappa 1.11.1 0 0.25-0.25 2.1292.129 0.5445δ0.5445\,\delta 2.279δ\frac{2.279}{\delta} 0.1167+0.04932δ-0.1167+0.04932\,\delta
CH-IIb κ\kappa 1.11.1 0.30.3 0.25-0.25 3.6743.674 0.4059+0.04598δ0.4059+0.04598\sqrt{\delta} 0.003473δ+0.2002δ\frac{0.003473}{\delta}+\frac{0.2002}{\sqrt{\delta}} 0.2223+0.01071δ0.2223+0.01071\sqrt{\delta}
BH-IIe 2κ2\kappa 0.90.9 0.20.2 0.25-0.25 1.9251.925 0.4886δ0.4886\,\delta 1.900δ\frac{1.900}{\delta} 0.08042+0.05239δ0.08042+0.05239\,\delta
BH-IIf 2κ2\kappa 1.11.1 0.20.2 0.25-0.25 2.1092.109 0.5100δ0.5100\,\delta 2.152δ\frac{2.152}{\delta} 0.06976+0.07281δ0.06976+0.07281\,\delta
BH-IIg 3κ3\kappa 0.90.9 0.20.2 0.25-0.25 1.9101.910 0.4865δ0.4865\,\delta 1.883δ\frac{1.883}{\delta} 0.08766+0.04201δ0.08766+0.04201\,\delta
BH-IIh 3κ3\kappa 1.11.1 0.20.2 0.25-0.25 2.1392.139 0.5315δ0.5315\,\delta 2.240δ\frac{2.240}{\delta} 0.06163+0.09186δ0.06163+0.09186\,\delta
CH-IIc 3κ3\kappa 1.11.1 0.30.3 0.5-0.5 2.3922.392 0.03228+0.1915δ0.03228+0.1915\sqrt{\delta} 1.001δ+0.3666δ\frac{1.001}{\delta}+\frac{0.3666}{\sqrt{\delta}} 0.03170+0.08472δ0.03170+0.08472\sqrt{\delta}
Refer to caption
Figure 6: Profiles of several representative solutions. Vertical lines are drawn in the plot of gttg_{tt} to indicate the event horizon of each solution. While the Schwarzschild solution with a zero bumblebee field has vanishing b2:=gμνbμbνb^{2}:=g^{\mu\nu}b_{\mu}b_{\nu} and the Ricci scalar RR, the two scalars are nonzero for general solutions in the bumblebee model as shown in the right panels.
Refer to caption
Figure 7: Geodesics bounce back on the event horizon of a CH; the solution “CH-Ic” in Table 1 is used. The ADM mass of the solution, μ1\mu_{1}, is used as the length unit. The black circles with a radius rh3.073r_{h}\approx 3.073 in both panels indicate the event horizon of the CH. The Schwarzschild radius, rh=2r_{h}=2 is indicated by the solid discs for comparison. In the upper panel, the dotted trajectories are the orbits with the same incident velocities in the Schwarzschild spacetime that fall into the Schwarzschild BH.

II.4 The name “compact hill” and its geodesics

The CH solutions share the same form of the asymptotic expansions with the BH solutions, so they mimic the BH solutions at large radius. But near the event horizon, the nonvanishing gttg_{tt} causes a distinctive feature for them. For CHs, geodesics bounce back on the event horizon, which is the reason for us to name them “compact hill”. The very terminology “event horizon” actually seems improper as this feature of the CH solutions is radically different from the event horizon of BHs. However, in the sense that the null hypersurface r=rhr=r_{h} of the CHs prevents geodesics outside of it from going in while the event horizon of BHs prevents geodesics inside from going out, we still refer r=rhr=r_{h} as the event horizon of the CHs.

Now to show that geodesics bounce back on the event horizon of the CHs, one first finds that under the metric in Eq. (8) the radial components of the four-velocity and the four-acceleration can be written as

(drdλ)2=e2μ(ϵ2e2νl2r2+k),\displaystyle\left(\frac{dr}{d\lambda}\right)^{2}=e^{-2\mu}\left(\epsilon^{2}\,e^{-2\nu}-\frac{l^{2}}{r^{2}}+k\right),
d2rdλ2=e2μ[ϵ2e2ν(μ+ν)+l2r2(μ+1r)kμ],\displaystyle\frac{d^{2}r}{d\lambda^{2}}=e^{-2\mu}\left[-\epsilon^{2}\,e^{-2\nu}(\mu^{\prime}+\nu^{\prime})+\frac{l^{2}}{r^{2}}\left(\mu^{\prime}+\frac{1}{r}\right)-k\mu^{\prime}\right], (19)

where the prime denotes the derivative with respect to rr, and the integral constants ϵ\epsilon and ll are

ϵ\displaystyle\epsilon =\displaystyle= e2νdtdλ,\displaystyle e^{2\nu}\frac{dt}{d\lambda},
l\displaystyle l =\displaystyle= r2dϕdλ,\displaystyle r^{2}\frac{d\phi}{d\lambda}, (20)

while the integral constant kk takes 1, 0-1,\,0, and 11, for timelike, lightlike, and spacelike geodesics as λ\lambda is the suitable affine parameter in each case. With gtt=e2νg_{tt}=-e^{2\nu} being nonzero on the event horizon, we immediately see dr/dλ0dr/d\lambda\rightarrow 0 on the event horizon. Moreover, using the expansions in Eq. (18) we find near the event horizon

(drdλ)2rrhM10(ϵ2N10l2rh2+k),\displaystyle\left(\frac{dr}{d\lambda}\right)^{2}\rightarrow\frac{r-r_{h}}{M_{10}}\left(\frac{\epsilon^{2}}{N_{10}}-\frac{l^{2}}{r_{h}^{2}}+k\right),
d2rdλ212M10(ϵ2N10l2rh2+k).\displaystyle\frac{d^{2}r}{d\lambda^{2}}\rightarrow\frac{1}{2M_{10}}\left(\frac{\epsilon^{2}}{N_{10}}-\frac{l^{2}}{r_{h}^{2}}+k\right). (21)

Because M10M_{10} is positive, the combination in the parentheses must be nonnegative for (dr/dλ)20\left(dr/d\lambda\right)^{2}\geq 0. Therefore near the event horizon the radial acceleration d2r/dλ2d^{2}r/d\lambda^{2} is nonnegative. Together with the fact that the radial velocity dr/dλdr/d\lambda approaches zero on the event horizon, we conclude that geodesics bounce back there.

The gravitational force of a CH exerting on a test mass then can be thought as repulsive near the event horizon while attractive far away. As far as we know there is completely no evidence for this kind of gravitational interaction in experiments and astrophysical observations. Anyway, the effect is academically interesting. The simplest scenario is that a test mass is released at rest and then oscillates between the initial position and the event horizon. More generally, bound orbits exist near the event horizon. The repulsive nature near the event horizon ensures that even the lightlike circular orbit at the radius determined by ν=1/r\nu^{\prime}=1/r is stable. Figure 7 shows example orbits bouncing back on the event horizon of a CH.

Classical trajectories bounce back on the event horizon, suggesting that GWs echo, which is a genuine effect that can be searched for in GW observations Westerweck et al. (2018); Testa and Pani (2018). It turns out that the potentials for waves are finite on the event horizon so waves are not perfectly reflected. What marvelous is that the potentials become complex inside the event horizon, and it is well known in quantum mechanics that complex potentials cause absorption of waves (e.g. see Refs. Avishai and Knoll (1976); Vibók and Balint‐Kurti (1992)). Before we approximately demonstrate the reflection on the event horizon of a CH using a scalar wave, let us admit that not only the potentials for waves but also the Ricci scalar becomes complex inside the event horizon of CHs. The origination of their imaginary parts can be seen from Eq. (18), where δ\sqrt{\delta} becomes imaginary when r<rhr<r_{h}. Though it first appears to us that the CH solutions are unphysical due to the complex Ricci scalar, a second thought about the absorption property related to the imaginary parts of the potentials strikes us the idea that a spacetime region with a complex Ricci scalar might be where information is lost. The idea needs to be further explored, but that is beyond the scope of the present work. In the remaining of this section, we only try to use a brute but manageable approximation to illustrate the reflection of a scalar wave on the event horizon of a CH.

Starting with the Klein-Gordon equation

0=gΦ=1gμ(gμνgνΦ),\displaystyle 0=\Box_{g}\Phi=\frac{1}{\sqrt{-g}}\partial_{\mu}\left(g^{\mu\nu}\sqrt{-g}\,\partial_{\nu}\Phi\right), (22)

and using the spherical metric in Eq. (8) as well as the spherical wave ansatz for the scalar field Φ=eiσtΨ(r)\Phi=e^{i\sigma t}\,\Psi(r), one obtains an equation for Ψ(r)\Psi(r)

Ψ′′+(νμ+2r)Ψ+σ2e2(μν)Ψ=0.\displaystyle\Psi^{\prime\prime}+\left(\nu^{\prime}-\mu^{\prime}+\frac{2}{r}\right)\Psi^{\prime}+\sigma^{2}e^{2(\mu-\nu)}\Psi=0. (23)

The changes of variables

z\displaystyle z =\displaystyle= rΨ,\displaystyle r\Psi,
drdr\displaystyle\frac{dr_{*}}{dr} =\displaystyle= eμν\displaystyle e^{\mu-\nu} (24)

put Eq. (23) into the standard form Berti et al. (2009)

d2zdr2+(σ2Vs)z=0,\displaystyle\frac{d^{2}z}{dr_{*}^{2}}+\left(\sigma^{2}-V_{s}\right)z=0, (25)

with the potential for the scalar wave being

Vs=e2(νμ)νμr.\displaystyle V_{s}=e^{2(\nu-\mu)}\frac{\nu^{\prime}-\mu^{\prime}}{r}. (26)

Fully solving the scattering problem of Eq. (25) is beyond the scope of the present work. We shall focus on the behavior of the wave near the event horizon, where the potential VV and the derivative dr/drdr_{*}/dr have the expansions

Vs=12rhN10M10[1+32(N20N10M20M10)δ+],\displaystyle V_{s}=\frac{1}{2r_{h}}\frac{N_{10}}{M_{10}}\left[1+\frac{3}{2}\left(\frac{N_{20}}{N_{10}}-\frac{M_{20}}{M_{10}}\right)\sqrt{\delta}+...\right],
drdr=M10N101δ[112(N20N10M20M10)δ+].\displaystyle\frac{dr_{*}}{dr}=\sqrt{\frac{M_{10}}{N_{10}}}\frac{1}{\sqrt{\delta}}\left[1-\frac{1}{2}\left(\frac{N_{20}}{N_{10}}-\frac{M_{20}}{M_{10}}\right)\sqrt{\delta}+...\right]. (27)

Besides being complex inside the event horizon for the potential VsV_{s}, Eq. (27) is different from the BH expressions because of two other features. First, for BHs VsV_{s} is zero on the event horizon while Eq. (27) is not. Then, the tortoise radius rr_{*} goes to negative infinity on the event horizon of BHs, but now that dr/drdr_{*}/dr is only divergent as 1/δ1/\sqrt{\delta} for the CHs we find rr_{*} proportional to δ\sqrt{\delta} plus an integral constant.

Taking the distinctive features of Eq. (27) into consideration, we reckon that the step potential

Vstep={Vs1,Re(r)0Vs2+iΓ2,Re(r)<0\displaystyle V_{\rm step}=\begin{cases}V_{s1},&{\rm Re}(r_{*})\geq 0\\ V_{s2}+i\Gamma_{2},&{\rm Re}(r_{*})<0\end{cases} (28)

might be used to approximate the effective potential VsV_{s} near the event horizon. In Eq. (28), the real constant Vs1V_{s1} and the complex constant Vs2+iΓ2V_{s2}+i\Gamma_{2} can be thought as certain average values of VsV_{s} right outside and inside the event horizon. The tortoise radius, which also becomes complex inside the event horizon, has been set to zero at the event horizon. With VsV_{s} brutely approximated by VstepV_{\rm step}, the scattering problem of Eq. (25) is analytically solvable. Assuming no waves coming out of the center at r=0r=0, the plane wave ansatz

z={C+eik1r+Ceik1r,Re(r)0D+eik2r,Re(r)<0\displaystyle z=\begin{cases}C_{+}e^{ik_{1}r_{*}}+C_{-}e^{-ik_{1}r_{*}},&{\rm Re}(r_{*})\geq 0\\ D_{+}e^{ik_{2}r_{*}},&{\rm Re}(r_{*})<0\end{cases} (29)

satisfies the equation when

k1\displaystyle k_{1} =\displaystyle= σ2Vs1,\displaystyle\sqrt{\sigma^{2}-V_{s1}},
k2\displaystyle k_{2} =\displaystyle= σ2Vs2iΓ2.\displaystyle\sqrt{\sigma^{2}-V_{s2}-i\Gamma_{2}}. (30)

The continuity of zz and dz/drdz/dr_{*} at r=0r_{*}=0 fixes the ratio between the amplitudes of the reflected and the incident waves, which is

CC+=k1k2k1+k2.\displaystyle\frac{C_{-}}{C_{+}}=\frac{k_{1}-k_{2}}{k_{1}+k_{2}}. (31)

The reflection rate is then

=|CC+|2=k12+|k2|22k1Re(k2)k12+|k2|2+2k1Re(k2),\displaystyle{\cal{R}}=\bigg{|}\frac{C_{-}}{C_{+}}\bigg{|}^{2}=\frac{k_{1}^{2}+|k_{2}|^{2}-2k_{1}{\rm Re}(k_{2})}{k_{1}^{2}+|k_{2}|^{2}+2k_{1}{\rm Re}(k_{2})}, (32)

where we have assumed the energy of the incident wave σ2\sigma^{2} to be greater than Vs1V_{s1} so that k1k_{1} is real. If we define θ2[π,π)\theta_{2}\in[-\pi,\pi) by

cosθ2=σ2Vs2(σ2Vs2)2+Γ22,\displaystyle\cos\theta_{2}=\frac{\sigma^{2}-V_{s2}}{\sqrt{\left(\sigma^{2}-V_{s2}\right)^{2}+\Gamma_{2}^{2}}},
sinθ2=Γ2(σ2Vs2)2+Γ22,\displaystyle\sin\theta_{2}=\frac{-\Gamma_{2}}{\sqrt{\left(\sigma^{2}-V_{s2}\right)^{2}+\Gamma_{2}^{2}}}, (33)

then k2k_{2} can be explicitly written as

k2=[(σ2Vs2)2+Γ22]14eiθ22.\displaystyle k_{2}=\left[\left(\sigma^{2}-V_{s2}\right)^{2}+\Gamma_{2}^{2}\right]^{\frac{1}{4}}e^{\frac{i\theta_{2}}{2}}. (34)

A brief understanding on how Γ2\Gamma_{2} affects the reflection rate can be achieved by assuming a small Γ2\Gamma_{2} so that Eq. (32), up to the leading order of Γ2\Gamma_{2}, reads

(k1k~2k1+k~2)2+k1(3k~22k12)2k~23(k1+k~2)4Γ22,\displaystyle{\cal{R}}\approx\left(\frac{k_{1}-\tilde{k}_{2}}{k_{1}+\tilde{k}_{2}}\right)^{2}+\frac{k_{1}\left(3\tilde{k}_{2}^{2}-k_{1}^{2}\right)}{2\,\tilde{k}_{2}^{3}(k_{1}+\tilde{k}_{2})^{4}}\Gamma_{2}^{2}, (35)

where k~2=σ2Vs2\tilde{k}_{2}=\sqrt{\sigma^{2}-V_{s2}}. The first term independent of Γ2\Gamma_{2} is the usual reflection rate due to a real step potential. The second term proportional to the square of Γ2\Gamma_{2} represents the contribution from the imaginary part of the potential. The contribution is positive when 3k~22k12>03\tilde{k}_{2}^{2}-k_{1}^{2}>0, namely

σ2>3Vs2Vs12.\displaystyle\sigma^{2}>\frac{3V_{s2}-V_{s1}}{2}. (36)

Because we have required σ2>Vs1\sigma^{2}>V_{s1}, Eq. (36) is always satisfied if Vs1Vs2V_{s1}\geq V_{s2}. If Vs1<Vs2V_{s1}<V_{s2}, then the imaginary potential contributes negatively to the reflection rate when the frequency of the incident wave is lower than the value in Eq. (36).

III Observational tests

The exotic solutions of CHs are not supported by current observations in the strong-field regime such as the GW detections and the imagies of the supermassive black holes. We consider testing the BH solutions in the bumblebee model against observations in this section.

III.1 Classical tests

Classical tests of the weak-field metric in the bumblebee model using observations in the Solar System was first done by Hellings and Nordtvedt (1973) within the PPN framework. It is equivalent to testing the metric asymptotic expansion in Eq. (13) with μ0=0\mu_{0}=0. In Appendix E, we transform the metric expansion in Eq. (13) to the harmonic coordinates so that the PPN parameters β\beta and γ\gamma Poisson and Will (2014) can be directly read off. Existing constraints on the PPN parameters can thus be translated to constrain the free coefficients in the asymptotic expansions in Eq. (13). This is an easy way to obtain constraints on the free parameters in our solutions. But to clearly see how the parameters in our solutions enter the classical effects and thus obtaining tailored constraints for them, we reckon that it is worth leaving the convenient PPN framework but calculating predictions directly using Eq. (13) to compare with observations.

Considering three classical tests of GR, the advance of perihelion, the deflection of light, and the Shapiro time delay, we find that the effect of the advance of perihelion is sensitive to the coefficients μ0,μ1,ν1\mu_{0},\,\mu_{1},\,\nu_{1} and ν2\nu_{2} while the effects of the other two are only sensitive to μ0,μ1\mu_{0},\,\mu_{1} and ν1\nu_{1}.111The other classical test of GR, the gravitational redshift, only tests the equivalence principle Will (2018), and it is left out of the current study because the bumblebee gravity model has already satisfied the Einstein equivalence principle. The calculations are done in the weak-field regime where the ratio of the central mass to the typical length involved is small so that the results are kept to the linear order of this ratio. The results for the deflection of light and the Shapiro time delay are the same as those derived in Ref. Casana et al. (2018), where tight constraints have been set on the corresponding parameter l=e2μ01l=e^{2\mu_{0}}-1. We focus on the new modification introduced by the coefficient ν2\nu_{2} in the effect of the advance of perihelion.

For a bound timelike geodesic with perihelion separation rminr_{\rm min} and aphelion separation rmaxr_{\rm max}, the advance of perihelion per orbit is given by

δϕ=2rminrmax|dϕdr|𝑑r2π.\displaystyle\delta\phi=2\int_{r_{\rm min}}^{r_{\rm max}}\Big{|}\frac{d\phi}{dr}\Big{|}\,dr-2\pi. (37)

Using Eqs. (19) and (20) to calculate dr/dϕdr/d\phi and substituting the expansions for μ\mu and ν\nu in Eq. (13) into it, up to the leading order of μ1/rmin\mu_{1}/r_{\rm min}, the advance of perihelion turns out to be

δϕ2π(eμ01)+πeμ0(μ1ν1+ν2ν1)rmin+rmaxrminrmax.\displaystyle\delta\phi\approx 2\pi\left(e^{\mu_{0}}-1\right)+\pi e^{\mu_{0}}\left(\mu_{1}-\nu_{1}+\frac{\nu_{2}}{\nu_{1}}\right)\frac{r_{\rm min}+r_{\rm max}}{r_{\rm min}r_{\rm max}}. (38)

Because there are two families of solutions with different recurrence relations for the coefficients in the asymptotic expansions, we discuss them separately.

For the family of solutions with br=0b_{r}=0, the first few recurrence relations are given in Eq. (52) with three free parameters, μ1,λ0\mu_{1},\,\lambda_{0} and λ1\lambda_{1}. Using the recurrence relations, Eq. (38) becomes

δϕ\displaystyle\delta\phi \displaystyle\approx π(32ξλ02+2ξλ0λ~1+(ξκ)λ~124ξ(1ξ2κ)λ024)\displaystyle\pi\left(3-\frac{2\xi\lambda_{0}^{2}+2\xi\lambda_{0}\tilde{\lambda}_{1}+(\xi-\kappa)\tilde{\lambda}_{1}^{2}}{4\xi\left(1-\frac{\xi}{2\kappa}\right)\lambda_{0}^{2}-4}\right) (39)
×μ1(rmin+rmax)rminrmax,\displaystyle\times\frac{\mu_{1}\left(r_{\rm min}+r_{\rm max}\right)}{r_{\rm min}r_{\rm max}},

where λ~1=λ1/μ1\tilde{\lambda}_{1}=\lambda_{1}/\mu_{1}. The recurrence relations in Eq. (52) are for both BH solutions and CH solutions. If we restrict our attention to the BH solutions in this family, we should keep in mind that there is only one free parameter among λ0\lambda_{0} and λ1\lambda_{1}. It is clear that Eq. (39) reproduces the well-known result

δϕGR\displaystyle\delta\phi_{\rm GR} \displaystyle\approx 3πμ1(rmin+rmax)rminrmax\displaystyle\frac{3\pi\mu_{1}\left(r_{\rm min}+r_{\rm max}\right)}{r_{\rm min}r_{\rm max}} (40)

for the Schwarzschild metric when λ0=λ1=0\lambda_{0}=\lambda_{1}=0. In addition, the bumblebee model recovers the Einstein-Maxwell theory when ξ=0\xi=0, so Eq. (39) gives the result for the Reissner-Nordström metric when ξ=0\xi=0 with λ1\lambda_{1} set to Q/κ/2Q/\sqrt{\kappa/2}.

For the family of solutions with Rrr=0R_{rr}=0, the first few recurrence relations are given in Eq. (55) with 5 free parameters, μ0,μ1,μ2λ0\mu_{0},\,\mu_{1},\mu_{2}\,\,\lambda_{0} and λ1\lambda_{1}. Using the recurrence relations, Eq. (38) becomes

δϕ2π(eμ01)+eμ0(7+2μ~2)πμ1(rmin+rmax)3rminrmax,\displaystyle\delta\phi\approx 2\pi\left(e^{\mu_{0}}-1\right)+e^{\mu_{0}}\left(7+2\tilde{\mu}_{2}\right)\frac{\pi\mu_{1}\left(r_{\rm min}+r_{\rm max}\right)}{3r_{\rm min}r_{\rm max}}, (41)

where μ~2=μ2/μ12\tilde{\mu}_{2}=\mu_{2}/\mu_{1}^{2}. We see that the bumblebee parameters λ0\lambda_{0} and λ1\lambda_{1} do not affect the advance of perihelion at the leading order of μ1/rmin\mu_{1}/r_{\rm min} in this family. Setting μ~2=1\tilde{\mu}_{2}=1, Eq. (41) reproduces the result in Ref. Casana et al. (2018) which in turn goes back to the Schwarzschild result if taking μ0=0\mu_{0}=0.

Table 3: Orbital parameters of 6 Solar-System planets and the fitting residuals of their perihelion advances.333Orbital parameters are taken from https://nssdc.gsfc.nasa.gov/planetary/factsheet/.
Planet Perihelion (106km10^{6}\,{\rm km}) Aphelion (106km10^{6}\,{\rm km}) Period (days) Perihelion advance residual (mas yr-1)
Mercury 46.00 69.82 87.97 0.020±0.030-0.020\pm 0.030
Venus 107.5 108.9 224.7 0.026±0.0160.026\pm 0.016
Earth 147.1 152.1 365.2 0.0019±0.00190.0019\pm 0.0019
Mars 206.7 249.3 687.0 0.00020±0.00037-0.00020\pm 0.00037
Jupiter 740.6 816.4 4331 0.59±0.280.59\pm 0.28
Saturn 1358 1507 10747 0.0032±0.0047-0.0032\pm 0.0047
Refer to caption
Figure 8: Constraints on the bumblebee parameters λ0\lambda_{0} and λ1\lambda_{1} for the first family of solutions (br=0b_{r}=0) from the fitting residuals of perihelion advances in Table 3. The metric parameter μ1\mu_{1} is used as the unit. In these plots, BHs exist along the black lines passing through the origin in the λ0\lambda_{0}-λ1\lambda_{1} plane (the brown lines in Fig. 3). Note that Jupiter and Saturn give constraints too loose to shown in the plots and they are omitted.
Refer to caption
Figure 9: Constraints on the metric parameters μ0\mu_{0} and μ2\mu_{2} for the second family of solutions (Rrr=0R_{rr}=0) from the fitting residuals of perihelion advances in Table 3. The metric parameter μ1\mu_{1} is used as the unit. Note that Jupiter and Saturn give constraints too loose to shown here and they are omitted.

Pitjeva and Pitjev (2013) fitted the advances of perihelia of 6 planets using standard gravity described by GR. We collect their fitting residuals as well as the orbital parameters in Table 3. We use the larger between a fitting residual and its uncertainty as an upper bound for |δϕδϕGR||\delta\phi-\delta\phi_{\rm GR}|. Then constraints on λ0\lambda_{0} and λ1\lambda_{1} for the first family of solutions can be set using Eq. (39), and constraints on μ0\mu_{0} and μ2\mu_{2} can be set for the second family of solutions using Eq. (41). The results are shown in Figs. 8 and 9. As expected, parameters in both solution families are severely constrained so that the metric outside the Sun is reasonably the Schwarzschild solution to a good approximation. Notice that the constraints on μ0\mu_{0} in Fig. 9 are consistent with the results in Ref. Casana et al. (2018). In the future, a global fit of data directly with expressions given by the bumblebee gravity model is desirable to further eliminate degeneracy among orbital parameters and theory parameters.

Though both pointing to the Schwarzschild solution, we feel that the constraints for the two families of solutions have different interpretations. For the first family of solutions, the constraints are on the bumblebee parameters λ0\lambda_{0} and λ1\lambda_{1}, meaning that the bumblebee charge of the Sun must be very small if not vanishing. Naively, different objects can carry different amount of bumblebee charge, so these constraints are not necessarily valid for other types of astrophysical objects, like BHs or neutron stars. For the second family of solutions, the constraints are directly on the metric parameters μ0\mu_{0} and μ2\mu_{2} regardless of the bumblebee parameters λ0\lambda_{0} and λ1\lambda_{1}. So they are supposed to be general and valid for all types of astrophysical objects.

III.2 BH images

Advances of perihelia of the planets in the Solar System set stringent constraints on the parameters of the static spherical solutions in the bumblebee gravity model. But the results are doubtfully universal for the first family of solutions where the constraints are on the bumblebee parameters λ0\lambda_{0} and λ1\lambda_{1}. While the Sun has tiny or zero bumblebee charge, more compact objects like BHs may still possess a considerable amount of it. So we consider to put bounds on the bumblebee charge carried by the supermassive BHs whose shadow images have been recently taken by the EHT Collaboration Akiyama et al. (2019, 2022a).

The mathematical shape of the shadow of a BH depends on the mass and the spin of it. When extracting from observations, the precise shape is also influenced by the observing resolution as well as the structure of the surrounding matter that emits EM waves. The mass of the BH determines the size of the shadow, while other elements contribute corrections typically of order unity Akiyama et al. (2019). For our purpose of setting preliminary bounds on the bumblebee charge, it is sufficient to consider the simplest model where a circular shadow is created by a static spherical BH. In this case, the radius of the shadow is the impact parameter of a lightlike geodesic infinitely approaching the lightlike circular orbit, i.e., the light ring. In the static spherical spacetime described by Eq. (8), the light ring, if exists, is at a radius rlrr_{\rm lr} given by

ν|r=rlr=1rlr,\displaystyle\nu^{\prime}\big{|}_{r=r_{\rm lr}}=\frac{1}{r_{\rm lr}}, (42)

and the limit of the impact parameter corresponding to the light ring is

σlr=reν|r=rlr.\displaystyle\sigma_{\rm lr}=re^{-\nu}\,\big{|}_{r=r_{\rm lr}}. (43)

Denoting the distance from the observer to the BH as DD, the angular diameter of the shadow is

d=2σlrD=2θgσlrM,\displaystyle d=\frac{2\sigma_{\rm lr}}{D}=2\theta_{g}\frac{\sigma_{\rm lr}}{M}, (44)

where M=μ1M=\mu_{1} is the mass of the BH and θg=M/D\theta_{g}=M/D. Adopting the observational results

d\displaystyle d =\displaystyle= 42±3μas,\displaystyle 42\pm 3\,\mu{\rm as},
θg\displaystyle\theta_{g} =\displaystyle= 3.62±0.60μas\displaystyle 3.62\pm 0.60\,\mu{\rm as} (45)

according to Refs. Gebhardt et al. (2011); Akiyama et al. (2019) for the supermassive BH in M87, we get

σlrM=5.80±1.05.\displaystyle\frac{\sigma_{\rm lr}}{M}=5.80\pm 1.05. (46)

Adopting the observational results444The average of the results from the Very Large Telescope Interferometer (VLTI) and Keck is used as the value for θg\theta_{g}, and their difference is used as the uncertainty.

d\displaystyle d =\displaystyle= 51.8±2.3μas,\displaystyle 51.8\pm 2.3\,\mu{\rm as},
θg\displaystyle\theta_{g} =\displaystyle= 5.02±0.20μas\displaystyle 5.02\pm 0.20\,\mu{\rm as} (47)

according to Refs. Do et al. (2019); Abuter et al. (2022); Akiyama et al. (2022a) for the supermassive BH in the Milky Way, we get

σlrM=5.16±0.31.\displaystyle\frac{\sigma_{\rm lr}}{M}=5.16\pm 0.31. (48)
Refer to caption
Figure 10: Radius of the shadow versus the bumblebee charge for static spherical BHs in the bumblebee gravity model. The two colored bands indicate the observational results in Eqs. (46) and (48), with the orange band for Sgr A and the blue band for M87. The mass of the BH is used as the unit.

In Fig. 10, numerically calculated σlr\sigma_{\rm lr} using Eq. (43) for the solutions in Fig. 5 is plotted with respect to the bumblebee charge. The observational results in Eqs. (46) and (48) are indicated by the shaded bands. We see that for different coupling constant ξ\xi, the bounds on the bumblebee charge can be very different. For example, if we look at the bounds from the result of Sgr A, the bumblebee charge is up to about 0.7 when ξ=κ\xi=-\kappa but to about 3.53.5 when ξ=3κ\xi=3\kappa.

IV Summary

The bumblebee gravity model is an important vector-tensor gravity theory with Lorentz-symmetry violation. Our systematic study of the static spherical vacuum solutions in the bumblebee gravity model reveals two families of solutions that have diverging grrg_{rr} but finite curvature scalars at a finite radius rhr_{h}. The first family, characterized by a bumblebee background field with only the temporal component, consists of solutions given by three free parameters. The second family, characterized by a vanishing rrrr-component of the Ricci tensor, consists of solutions given by 5 free parameters. In each family, BH solutions exist under the requirement that gttg_{tt} vanishes at the event horizon rhr_{h}. The remaining solutions that have nonzero gttg_{tt} at rhr_{h} are dubbed “CHs” for geodesics bounce back at the event horizon (see Sec. II.4).

The fascinating CH solutions provide a gravitational interaction that attracts in the same way as the usual gravity far away from the center but is strongly repulsive near the event horizon. Consequently, unlike BHs, stable bound orbits exist right outside the event horizon of CHs. If waves are considered instead of classical geodesics, one finds that the potentials for waves are finite at the event horizon of CHs so that waves can transit into the event horizon while being partially reflected. Furthermore, because the potentials are complex inside the event horizon of CHs, the waves entering the event horizon will be at least partially absorbed by spacetime itself, causing possible loss of information. We concede that no experiments and observations support any of these features. However, the echoes of waves at the event horizon of CHs and the loss of information associated with spacetime that has complex potentials for waves and complex Ricci scalar are certainly interesting issues worth further study.

Restricting our attention to the BH solutions, we find that in the first family, the number of free parameters reduces from three to two, namely the mass and the charge, while in the second family there are still 5 free parameters. The BH solutions in both families have their merits. In the first family, the solutions recover the Reissner-Nordström metric when ξ=0\xi=0. In this sense, they are the expected BH solutions when the coupling between the Ricci tensor and the vector field is added to the Einstein-Maxwell theory. In the second family, there is an analytical solution given by Eqs. (9) and (10), which deserves a mention because by setting μ0=0\mu_{0}=0 the Schwarzschild metric is recovered together with a nontrivial bumblebee background field. It is amusing to notice the special case of ξ=2κ\xi=2\kappa, for both families give the Schwarzschild metric together with a bumblebee background field that has only a temporal component proportional to 12M/r1-2M/r.

In Sec. III, the BH solutions are tested using Solar-System observations and the images of the supermassive BHs. For the solutions in the first family, constraints on the bumblebee charge carried by the central objects are obtained. The upper bound for the ratio of the bumblebee charge to the mass is at the order of 10310^{-3} for the Sun while it roughly varies from 0.70.7 to 3.53.5 depending on the value of ξ\xi for the supermassive BHs. We need to point out that the bumblebee charge is completely unconstrained when ξ=2κ\xi=2\kappa as the metric is exactly Schwarzschild in this case. For the solutions in the second family, stringent constraints on the metric parameters are obtained, suggesting that if static spherical objects are described by the solutions in this family, the spacetime is most likely to be Schwarzschild but can be accompanied by a nontrivial bumblebee background field.

In conclusion, as a generalization of the Einstein-Maxwell theory and an important example for Lorentz-violating gravitational theories, the bumblebee gravity model in Eq. (2) indeed contains richer solutions. Testing these solutions in experiments and observations is undoubtedly a worthy subject for fundamental physics. On the one hand, the bumblebee background field showing up from nowhere in the vacuum solutions defines special local frames so that local Lorentz symmetry is broken. Therefore testing whether there is any such background field is testing the fundamental symmetry of our Universe. On the other hand, if the bumblebee field is regarded as the EM vector potential, testing the solutions in the bumblebee model will tell us whether the EM field only minimally couples with gravity as in the Einstein-Maxwell theory or the nonminimal coupling between the Ricci tensor and the vector field exists. The tests done in this work have not excluded the possibility of a bumblebee background field though the metric is constrained to be close to the Schwarzschild metric in the spherically symmetric situation. The next step is naturally to put the solutions into test against the GW observations. Future works on extending the spherical BH solutions to rotating BHs, calculating the quasinormal modes of the BHs, as well as solving the two-body problem in the PPN framework, are directions that need to be explored.

Acknowledgements.
We thank Bin Chen and V. Alan Kostelecký for insightful discussions and comments. We are also grateful to the anonymous referee for helpful feedback. This work was supported by the National Natural Science Foundation of China (11991053, 12147120, 11975027, 11721303), the China Postdoctoral Science Foundation (2021TQ0018), the National SKA Program of China (2020SKA0120300), the Max Planck Partner Group Program funded by the Max Planck Society, and the High-Performance Computing Platform of Peking University. R.X. is supported by the Boya Postdoctoral Fellowship at Peking University.

Appendix A Field equations under the static spherical ansatz

Denoting Eμν=Gμνκ(Tb)μνE_{\mu\nu}=G_{\mu\nu}-\kappa\left(T_{b}\right)_{\mu\nu}, with the static spherical ansatz the nonvanishing Einstein field equations are

0=Ett\displaystyle 0=E_{tt} =\displaystyle= e2ν2μ(12rμe2μ)r2ξe2μ[bt′′bt(μ+4ν2r)+bt(2μν4rν2ν′′+ν2)]bt+(κ2ξ)e2μbt 2\displaystyle\frac{e^{2\nu-2\mu}\left(1-2r\mu^{\prime}-e^{2\mu}\right)}{r^{2}}-\xi e^{-2\mu}\left[b_{t}^{\prime\prime}-b_{t}^{\prime}\left(\mu^{\prime}+4\nu^{\prime}-\frac{2}{r}\right)+b_{t}\left(2\mu^{\prime}\nu^{\prime}-\frac{4}{r}\nu^{\prime}-2\nu^{\prime\prime}+\nu^{\prime 2}\right)\right]b_{t}+\left(\frac{\kappa}{2}-\xi\right)e^{-2\mu}b_{t}^{\prime\,2}
+ξe2ν4μ[br′′br(5μ4r)+br(μ′′+3μ26rμ+1r2)]br+ξe2ν4μbr 2,\displaystyle+\xi e^{2\nu-4\mu}\left[b_{r}^{\prime\prime}-b_{r}^{\prime}\left(5\mu^{\prime}-\frac{4}{r}\right)+b_{r}\left(-\mu^{\prime\prime}+3\mu^{\prime 2}-\frac{6}{r}\mu^{\prime}+\frac{1}{r^{2}}\right)\right]b_{r}+\xi e^{2\nu-4\mu}b_{r}^{\prime\,2},
0=Err\displaystyle 0=E_{rr} =\displaystyle= e2μ2rν1r2e2ν(κ2bt 2ξbtbtν+ξbt2ν2)ξe2μ[br(ν+2r)br(ν′′+ν22rν1r2)]br,\displaystyle\frac{e^{2\mu}-2r\nu^{\prime}-1}{r^{2}}-e^{-2\nu}\left(\frac{\kappa}{2}b_{t}^{\prime\,2}-\xi b_{t}b_{t}^{\prime}\nu^{\prime}+\xi b_{t}^{2}\nu^{\prime 2}\right)-\xi e^{-2\mu}\left[b_{r}^{\prime}\left(\nu^{\prime}+\frac{2}{r}\right)-b_{r}\left(\nu^{\prime\prime}+\nu^{\prime 2}-\frac{2}{r}\nu^{\prime}-\frac{1}{r^{2}}\right)\right]b_{r},
0=Eθθ\displaystyle 0=E_{\theta\theta} =\displaystyle= Eϕϕsin2θ=r2e2μ(μν+1rμ1rνν′′ν2)+r2e2(μ+ν)(κ2bt 2ξbtbtν+ξbt2ν2)ξr2e4μbr 2\displaystyle\frac{E_{\phi\phi}}{\sin^{2}\theta}=r^{2}e^{-2\mu}\left(\mu^{\prime}\nu^{\prime}+\frac{1}{r}\mu^{\prime}-\frac{1}{r}\nu^{\prime}-\nu^{\prime\prime}-\nu^{\prime 2}\right)+r^{2}e^{-2(\mu+\nu)}\left(\frac{\kappa}{2}b_{t}^{\prime\,2}-\xi b_{t}b_{t}^{\prime}\nu^{\prime}+\xi b_{t}^{2}\nu^{\prime 2}\right)-\xi r^{2}e^{-4\mu}b_{r}^{\prime\,2}
ξr2e4μ[br′′br(5μ2ν2r)+br(3μ(μν1r)+ν′′μ′′+ν2+νr)]br,\displaystyle\hskip 34.14322pt-\xi r^{2}e^{-4\mu}\left[b_{r}^{\prime\prime}-b_{r}^{\prime}\left(5\mu^{\prime}-2\nu^{\prime}-\frac{2}{r}\right)+b_{r}\left(3\mu^{\prime}\left(\mu^{\prime}-\nu^{\prime}-\frac{1}{r}\right)+\nu^{\prime\prime}-\mu^{\prime\prime}+\nu^{\prime 2}+\frac{\nu^{\prime}}{r}\right)\right]b_{r},
0=Etr\displaystyle 0=E_{tr} =\displaystyle= ξκe2μ(μν+2rμν′′ν2)btbr,\displaystyle\frac{\xi}{\kappa}e^{-2\mu}\left(\mu^{\prime}\nu^{\prime}+\frac{2}{r}\mu^{\prime}-\nu^{\prime\prime}-\nu^{\prime 2}\right)b_{t}b_{r}, (49)

and the vector field equation gives

0\displaystyle 0 =\displaystyle= bt′′bt(μ+ν2r)+ξκbt(μν2rνν′′ν2),\displaystyle b_{t}^{\prime\prime}-b_{t}^{\prime}\left(\mu^{\prime}+\nu^{\prime}-\frac{2}{r}\right)+\frac{\xi}{\kappa}b_{t}\left(\mu^{\prime}\nu^{\prime}-\frac{2}{r}\nu^{\prime}-\nu^{\prime\prime}-\nu^{\prime 2}\right),
0\displaystyle 0 =\displaystyle= ξκ(μν+2rμν′′ν2)br,\displaystyle\frac{\xi}{\kappa}\left(\mu^{\prime}\nu^{\prime}+\frac{2}{r}\mu^{\prime}-\nu^{\prime\prime}-\nu^{\prime 2}\right)b_{r}, (50)

where the prime indicates the derivative with respect to rr. Noticing Rrr=(μν+2μ/rν′′ν2)R_{rr}=\left(\mu^{\prime}\nu^{\prime}+2\mu^{\prime}/r-\nu^{\prime\prime}-\nu^{\prime 2}\right), the second equation in Eq. (50) leads to two possibilities:

  1. 1.

    br=0b_{r}=0, in which case μ\mu and μ\mu^{\prime} can be eliminated from the equations, so we arrive at a system of two second-order ODEs for ν\nu and btb_{t};

  2. 2.

    Rrr=0R_{rr}=0, in which case we find that brb_{r} can be solved in terms of other variables,

    br2\displaystyle b_{r}^{2} =\displaystyle= e2μ2ν3ξr[(2κξ)rν+κ2ξ]bt 2+6ξrνbtbt2ξ[(ξκ1)(rμν+2μ+2ν)+(ξκ+1)rν 2]bt2rμν+2μrν2+ν\displaystyle\frac{e^{2\mu-2\nu}}{3\xi}\frac{r\left[(2\kappa-\xi)r\nu^{\prime}+\kappa-2\xi\right]b_{t}^{\prime\,2}+6\xi r\nu^{\prime}b_{t}b_{t}^{\prime}-2\xi\left[\left(\frac{\xi}{\kappa}-1\right)\left(r\mu^{\prime}\nu^{\prime}+2\mu^{\prime}+2\nu^{\prime}\right)+\left(\frac{\xi}{\kappa}+1\right)r\nu^{\prime\,2}\right]b_{t}^{2}}{r\mu^{\prime}\nu^{\prime}+2\mu^{\prime}-r\nu^{\prime 2}+\nu^{\prime}} (51)
    e2μξrμν+2μrν 2+ν+e2μνrμν+2μrν2+ν,\displaystyle-\frac{e^{2\mu}}{\xi}\frac{r\mu^{\prime}\nu^{\prime}+2\mu^{\prime}-r\nu^{\prime\,2}+\nu^{\prime}+e^{2\mu}\nu^{\prime}}{r\mu^{\prime}\nu^{\prime}+2\mu^{\prime}-r\nu^{\prime 2}+\nu^{\prime}},

    so the system consists of three second-order ODEs for μ,ν\mu,\,\nu and btb_{t}.

Appendix B Asymptotic expansions of the variables

For the case of br=0b_{r}=0, substituting the expansions in Eq. (13) into Eqs. (49) and (50), we find the following recurrence relations

μ0=0,\displaystyle\mu_{0}=0,
ν1=μ1,\displaystyle\nu_{1}=-\mu_{1},
ν2=μ12+2ξμ12λ02+2ξμ1λ0λ1+(ξκ)λ124[ξ(1ξ2κ)λ021],\displaystyle\nu_{2}=-\mu_{1}^{2}+\frac{2\xi\mu_{1}^{2}\lambda_{0}^{2}+2\xi\mu_{1}\lambda_{0}\lambda_{1}+(\xi-\kappa)\lambda_{1}^{2}}{4\left[\xi\left(1-\frac{\xi}{2\kappa}\right)\lambda_{0}^{2}-1\right]},
μ2=μ12+ξ(1ξ2κ)λ02[2ξμ12λ02+2ξμ1λ0λ1+κλ12]4[ξ(1ξ2κ)λ021]\displaystyle\mu_{2}=\mu_{1}^{2}+\frac{\xi(1-\frac{\xi}{2\kappa})\lambda_{0}^{2}\left[2\xi\mu_{1}^{2}\lambda_{0}^{2}+2\xi\mu_{1}\lambda_{0}\lambda_{1}+\kappa\lambda_{1}^{2}\right]}{4\left[\xi(1-\frac{\xi}{2\kappa})\lambda_{0}^{2}-1\right]}
6ξμ12λ02+6ξμ1λ0λ1+(2ξκ)λ124[ξ(1ξ2κ)λ021],\displaystyle\hskip 17.07182pt-\frac{6\xi\mu_{1}^{2}\lambda_{0}^{2}+6\xi\mu_{1}\lambda_{0}\lambda_{1}+(2\xi-\kappa)\lambda_{1}^{2}}{4\left[\xi(1-\frac{\xi}{2\kappa})\lambda_{0}^{2}-1\right]},
λ2=ξλ0[2ξμ12λ02+2ξμ1λ0λ1+(ξκ)λ12]4κ[ξ(1ξ2κ)λ021],\displaystyle\lambda_{2}=\frac{\xi\lambda_{0}\left[2\xi\mu_{1}^{2}\lambda_{0}^{2}+2\xi\mu_{1}\lambda_{0}\lambda_{1}+(\xi-\kappa)\lambda_{1}^{2}\right]}{4\kappa\left[\xi\left(1-\frac{\xi}{2\kappa}\right)\lambda_{0}^{2}-1\right]},
.\displaystyle...\,. (52)

Since the system of equations can be reduced to two second-order ODEs, there should be 4 free coefficients in principle. However, we have dropped the constant term ν0\nu_{0} in the expansion of ν\nu due to the freedom to redefine the time coordinate. There are three independent coefficients unfixed as shown in Eq. (52). They are taken as μ1,λ0\mu_{1},\,\lambda_{0} and λ1\lambda_{1}.

In this family of solutions, the Schwarzschild metric

ν=μ=12ln(12Mr)\displaystyle\nu=-\mu=\frac{1}{2}\ln{\left(1-\frac{2M}{r}\right)} (53)

is recovered by taking μ1=M\mu_{1}=M and λ0=λ1=0\lambda_{0}=\lambda_{1}=0. We also note that the Reissner-Nordström metric

ν=μ=12ln(12Mr+Q2r2)\displaystyle\nu=-\mu=\frac{1}{2}\ln{\left(1-\frac{2M}{r}+\frac{Q^{2}}{r^{2}}\right)} (54)

is recovered by taking μ1=M\mu_{1}=M and λ12=2Q2/κ\lambda_{1}^{2}=2Q^{2}/\kappa specifically for ξ=0\xi=0.

For the case of Rrr=0R_{rr}=0, substituting the expansions in Eq. (13) together with Eq. (51) into Eqs. (49) and (50), we find the following recurrence relations

ν1=μ1,\displaystyle\nu_{1}=-\mu_{1},
ν2=13(μ12+2μ2),\displaystyle\nu_{2}=-\frac{1}{3}\left(\mu_{1}^{2}+2\mu_{2}\right),
λ2=2ξλ03κ(μ12μ2),\displaystyle\lambda_{2}=\frac{2\xi\lambda_{0}}{3\kappa}\left(\mu_{1}^{2}-\mu_{2}\right),
,\displaystyle...\,, (55)

where 5 coefficients, μ0,μ1,μ2,λ0\mu_{0},\,\mu_{1},\,\mu_{2},\,\lambda_{0} and λ1\lambda_{1}, are unfixed. In this case, the system of equations can be reduced to three second-order ODEs, so there are 6 free coefficients in the solutions taking into account the trivial coefficient ν0\nu_{0}. The asymptotic expansion of brb_{r} can also be obtained. It is

br2\displaystyle b_{r}^{2} =\displaystyle= e2μ0(eμ01)ξ+e2μ09ξμ1[2ξκ((κ+4ξ)μ12+(8κ4ξ)μ2)λ02+18ξμ1λ0λ1+3(2ξκ)λ12\displaystyle\frac{e^{2\mu_{0}}\left(e^{\mu_{0}}-1\right)}{\xi}+\frac{e^{2\mu_{0}}}{9\xi\mu_{1}}\Big{[}\frac{2\xi}{\kappa}\left(\left(\kappa+4\xi\right)\mu_{1}^{2}+(8\kappa-4\xi)\mu_{2}\right)\lambda_{0}^{2}+18\xi\mu_{1}\lambda_{0}\lambda_{1}+3(2\xi-\kappa)\lambda_{1}^{2} (56)
+2(15e2μ09)μ1212e2μ0μ2]1r+.\displaystyle+2\left(15e^{2\mu_{0}}-9\right)\mu_{1}^{2}-12e^{2\mu_{0}}\mu_{2}\Big{]}\frac{1}{r}+...\,.

For br20b_{r}^{2}\geq 0 at rr\rightarrow\infty, it implies a constraint

eμ01ξ0.\displaystyle\frac{e^{\mu_{0}}-1}{\xi}\geq 0. (57)

If μ0=0\mu_{0}=0, then there is a constraint

2[(1+4ξκ)μ12+(84ξκ)μ2]λ02+18μ1λ0λ1\displaystyle 2\left[\left(1+\frac{4\xi}{\kappa}\right)\mu_{1}^{2}+\left(8-\frac{4\xi}{\kappa}\right)\mu_{2}\right]\lambda_{0}^{2}+18\mu_{1}\lambda_{0}\lambda_{1}
+3(2κξ)λ12+12ξ(μ12μ2)0.\displaystyle\hskip 28.45274pt+3\left(2-\frac{\kappa}{\xi}\right)\lambda_{1}^{2}+\frac{12}{\xi}\left(\mu_{1}^{2}-\mu_{2}\right)\geq 0. (58)

We point out that this family of solutions includes the analytical solution given by Eqs. (9) and (10). It is recovered by taking μ2=μ12\mu_{2}=\mu_{1}^{2}.

Appendix C Behavior of the variables near the event horizon

For the case of br=0b_{r}=0, substituting the expansions in Eq. (18) into Eqs. (49) and (50), we find the following three equations relating the 7 parameters {rh,N10,N20,M10,M20,L10,L20}\big{\{}r_{h},\,N_{10},\,N_{20},\,M_{10},\,M_{20},\,L_{10},\,L_{20}\big{\}} if N100N_{10}\neq 0,

0\displaystyle 0 =\displaystyle= 16M10N103+ξN202L102rh22ξN10N20L10L20rh2+2κN102L202rh2,\displaystyle-16M_{10}N_{10}^{3}+\xi N_{20}^{2}L_{10}^{2}r_{h}^{2}-2\xi N_{10}N_{20}L_{10}L_{20}r_{h}^{2}+2\kappa N_{10}^{2}L_{20}^{2}r_{h}^{2},
0\displaystyle 0 =\displaystyle= 16κM10N10416κN104rh8ξ(ξ2κ)N103L102rh+3κξN10N202L102rh2+ξ2(ξ2κ)N202L104rh26κξN102N20L10L20rh2\displaystyle 16\kappa M_{10}N_{10}^{4}-16\kappa N_{10}^{4}r_{h}-8\xi(\xi-2\kappa)N_{10}^{3}L_{10}^{2}r_{h}+3\kappa\xi N_{10}N_{20}^{2}L_{10}^{2}r_{h}^{2}+\xi^{2}(\xi-2\kappa)N_{20}^{2}L_{10}^{4}r_{h}^{2}-6\kappa\xi N_{10}^{2}N_{20}L_{10}L_{20}r_{h}^{2}
2ξ2(ξ2κ)N10N20L103L20rh2+2κ(2ξκ)N103L202rh2+2κξ(ξ2κ)N102L102L202rh2,\displaystyle-2\xi^{2}(\xi-2\kappa)N_{10}N_{20}L_{10}^{3}L_{20}r_{h}^{2}+2\kappa(2\xi-\kappa)N_{10}^{3}L_{20}^{2}r_{h}^{2}+2\kappa\xi(\xi-2\kappa)N_{10}^{2}L_{10}^{2}L_{20}^{2}r_{h}^{2},
0\displaystyle 0 =\displaystyle= 64κM10M20N105L10+176κM102N104N20L10+16ξ(ξ2κ)M102N103N20L103128κM102N105L2016κM20N105L10rh\displaystyle 64\kappa M_{10}M_{20}N_{10}^{5}L_{10}+176\kappa M_{10}^{2}N_{10}^{4}N_{20}L_{10}+16\xi(\xi-2\kappa)M_{10}^{2}N_{10}^{3}N_{20}L_{10}^{3}-128\kappa M_{10}^{2}N_{10}^{5}L_{20}-16\kappa M_{20}N_{10}^{5}L_{10}r_{h} (59)
176κM10N104N20L10rh8ξ(ξ2κ)M20N104L103rh56ξ(ξ2κ)M10N103N20L103rh+128κM10N105L20rh\displaystyle-176\kappa M_{10}N_{10}^{4}N_{20}L_{10}r_{h}-8\xi(\xi-2\kappa)M_{20}N_{10}^{4}L_{10}^{3}r_{h}-56\xi(\xi-2\kappa)M_{10}N_{10}^{3}N_{20}L_{10}^{3}r_{h}+128\kappa M_{10}N_{10}^{5}L_{20}r_{h}
+12κξM20N102N202L103rh2+21κξM10N10N203L103rh2+4ξ2(ξ2κ)M20N10N202L105rh2+5ξ2(ξ2κ)M10N203L105rh2\displaystyle+12\kappa\xi M_{20}N_{10}^{2}N_{20}^{2}L_{10}^{3}r_{h}^{2}+21\kappa\xi M_{10}N_{10}N_{20}^{3}L_{10}^{3}r_{h}^{2}+4\xi^{2}(\xi-2\kappa)M_{20}N_{10}N_{20}^{2}L_{10}^{5}r_{h}^{2}+5\xi^{2}(\xi-2\kappa)M_{10}N_{20}^{3}L_{10}^{5}r_{h}^{2}
24κξM20N103N20L102L20rh266κξM10N102N202L102L20rh28ξ2(ξ2κ)M20N102N20L104L20rh2\displaystyle-24\kappa\xi M_{20}N_{10}^{3}N_{20}L_{10}^{2}L_{20}r_{h}^{2}-66\kappa\xi M_{10}N_{10}^{2}N_{20}^{2}L_{10}^{2}L_{20}r_{h}^{2}-8\xi^{2}(\xi-2\kappa)M_{20}N_{10}^{2}N_{20}L_{10}^{4}L_{20}r_{h}^{2}
10ξ2(ξ2κ)M10N10N202L104L20rh2+8κ(2ξκ)M20N104L10L202rh2+2κ(40ξ11κ)M10N103N20L10L202rh2\displaystyle-10\xi^{2}(\xi-2\kappa)M_{10}N_{10}N_{20}^{2}L_{10}^{4}L_{20}r_{h}^{2}+8\kappa(2\xi-\kappa)M_{20}N_{10}^{4}L_{10}L_{20}^{2}r_{h}^{2}+2\kappa(40\xi-11\kappa)M_{10}N_{10}^{3}N_{20}L_{10}L_{20}^{2}r_{h}^{2}
+8κξ(ξ2κ)M20N103L103L202rh2+10κξ(ξ2κ)M10N102N20L103L202rh216κ(2ξκ)M10N104L203rh2.\displaystyle+8\kappa\xi(\xi-2\kappa)M_{20}N_{10}^{3}L_{10}^{3}L_{20}^{2}r_{h}^{2}+10\kappa\xi(\xi-2\kappa)M_{10}N_{10}^{2}N_{20}L_{10}^{3}L_{20}^{2}r_{h}^{2}-16\kappa(2\xi-\kappa)M_{10}N_{10}^{4}L_{20}^{3}r_{h}^{2}.

From Eq. (59), three of the parameters can be solved in terms of the other 4 parameters. Other higher-order coefficients are also fixed by their recurrence relations once the 4 free parameters are specified. If N10=0N_{10}=0, then substituting the expansions in Eq. (18) into Eqs. (49) and (50) simply leads to N2n=M2n=L2n=0N_{2n}=M_{2n}=L_{2n}=0 and L10=0L_{10}=0. The recurrence relations are much simplified with only three free parameters. Taking the three free parameters as N11,L11N_{11},\,L_{11} and rhr_{h}, the recurrence relations for the first few coefficients are

M10\displaystyle M_{10} =\displaystyle= rh+(κ2ξ4)L112rh2N11,\displaystyle r_{h}+\left(\frac{\kappa}{2}-\frac{\xi}{4}\right)\frac{L_{11}^{2}r_{h}^{2}}{N_{11}},
M11\displaystyle M_{11} =\displaystyle= 1(κ24κξ+13ξ2163ξ316κ)L114rh2N112+(3κ2ξ89κξ216+9ξ3323ξ464κ)L116rh3N113,\displaystyle 1-\left(\frac{\kappa^{2}}{4}-\kappa\xi+\frac{13\xi^{2}}{16}-\frac{3\xi^{3}}{16\kappa}\right)\frac{L_{11}^{4}r_{h}^{2}}{N_{11}^{2}}+\left(\frac{3\kappa^{2}\xi}{8}-\frac{9\kappa\xi^{2}}{16}+\frac{9\xi^{3}}{32}-\frac{3\xi^{4}}{64\kappa}\right)\frac{L_{11}^{6}r_{h}^{3}}{N_{11}^{3}},
M12\displaystyle M_{12} =\displaystyle= (κ24κξ+13ξ2163ξ316κ)L114rhN112+(κ3819κ2ξ12+105κξ232173ξ364+191ξ4192κ35ξ5256κ2)L116rh2N113\displaystyle\left(\frac{\kappa^{2}}{4}-\kappa\xi+\frac{13\xi^{2}}{16}-\frac{3\xi^{3}}{16\kappa}\right)\frac{L_{11}^{4}r_{h}}{N_{11}^{2}}+\left(\frac{\kappa^{3}}{8}-\frac{19\kappa^{2}\xi}{12}+\frac{105\kappa\xi^{2}}{32}-\frac{173\xi^{3}}{64}+\frac{191\xi^{4}}{192\kappa}-\frac{35\xi^{5}}{256\kappa^{2}}\right)\frac{L_{11}^{6}r_{h}^{2}}{N_{11}^{3}}
(49κ3ξ96433κ2ξ2192+207κξ364809ξ4384+997ξ51536κ79ξ61024κ2)L118rh3N114\displaystyle-\left(\frac{49\kappa^{3}\xi}{96}-\frac{433\kappa^{2}\xi^{2}}{192}+\frac{207\kappa\xi^{3}}{64}-\frac{809\xi^{4}}{384}+\frac{997\xi^{5}}{1536\kappa}-\frac{79\xi^{6}}{1024\kappa^{2}}\right)\frac{L_{11}^{8}r_{h}^{3}}{N_{11}^{4}}
+(11κ3ξ23255κ2ξ364+55κξ46455ξ5128+55ξ6512κ11ξ71024κ2)L1110rh4N115,\displaystyle+\left(\frac{11\kappa^{3}\xi^{2}}{32}-\frac{55\kappa^{2}\xi^{3}}{64}+\frac{55\kappa\xi^{4}}{64}-\frac{55\xi^{5}}{128}+\frac{55\xi^{6}}{512\kappa}-\frac{11\xi^{7}}{1024\kappa^{2}}\right)\frac{L_{11}^{10}r_{h}^{4}}{N_{11}^{5}},
N12\displaystyle N_{12} =\displaystyle= N11rh+(κ2ξ4)L112+(κξ4ξ24+ξ316κ)L114rhN11,\displaystyle-\frac{N_{11}}{r_{h}}+\left(\frac{\kappa}{2}-\frac{\xi}{4}\right)L_{11}^{2}+\left(\frac{\kappa\xi}{4}-\frac{\xi^{2}}{4}+\frac{\xi^{3}}{16\kappa}\right)\frac{L_{11}^{4}r_{h}}{N_{11}},
L12\displaystyle L_{12} =\displaystyle= L11rh+(ξ4ξ28κ)L113N11+(κξ4ξ24+ξ316κ)L115rhN112.\displaystyle-\frac{L_{11}}{r_{h}}+\left(\frac{\xi}{4}-\frac{\xi^{2}}{8\kappa}\right)\frac{L_{11}^{3}}{N_{11}}+\left(\frac{\kappa\xi}{4}-\frac{\xi^{2}}{4}+\frac{\xi^{3}}{16\kappa}\right)\frac{L_{11}^{5}r_{h}}{N_{11}^{2}}. (60)

Note that for ξ2κ\xi\neq 2\kappa, the Schwarzschild metric is recovered only when L11=0L_{11}=0 and therefore bt=0b_{t}=0. For ξ=2κ\xi=2\kappa, the Schwarzschild metric can exist together with a nonzero btb_{t}.

For the case of Rrr=0R_{rr}=0, using Eq. (51) to eliminate brb_{r} and then substituting the expansions in Eq. (18) into Eqs. (49) and (50), we find the following equation relating the 7 parameters {rh,N10,N20,M10,M20,L10,L20}\big{\{}r_{h},\,N_{10},\,N_{20},\,M_{10},\,M_{20},\,L_{10},\,L_{20}\big{\}} if N100N_{10}\neq 0,

0\displaystyle 0 =\displaystyle= 384κM102N10336κM10M20N102N20rh+12κM102N10N202rh8ξ(2ξκ)M10N202L102rh248κ2M10N102L202rh2\displaystyle 384\kappa M_{10}^{2}N_{10}^{3}-36\kappa M_{10}M_{20}N_{10}^{2}N_{20}r_{h}+12\kappa M_{10}^{2}N_{10}N_{20}^{2}r_{h}-8\xi\left(2\xi-\kappa\right)M_{10}N_{20}^{2}L_{10}^{2}r_{h}^{2}-48\kappa^{2}M_{10}N_{10}^{2}L_{20}^{2}r_{h}^{2} (61)
+16ξ(2ξκ)M10N10N20L10L20rh23κ(ξ2κ)M20N10N20L202rh3+κ(ξ2κ)M10N202L202rh3.\displaystyle+16\xi\left(2\xi-\kappa\right)M_{10}N_{10}N_{20}L_{10}L_{20}r_{h}^{2}-3\kappa\left(\xi-2\kappa\right)M_{20}N_{10}N_{20}L_{20}^{2}r_{h}^{3}+\kappa\left(\xi-2\kappa\right)M_{10}N_{20}^{2}L_{20}^{2}r_{h}^{3}.

From Eq. (61), one of the parameters can be solved in terms of the other 6 parameters. Other higher-order coefficients are then fixed by their recurrence relations once the 6 free parameters are specified. Similar to the first case, if N10=0N_{10}=0 then recurrence relations give N2n=M2n=L2n=0N_{2n}=M_{2n}=L_{2n}=0, but in this case L10L_{10} is no longer fixed to zero so there are still 6 free parameters. Taking the 6 free parameters as N11,M10,M11,L10,L11N_{11},\,M_{10},\,M_{11},\,L_{10},\,L_{11} and rhr_{h}, the recurrence relations for the first few coefficients are

M12\displaystyle M_{12} =\displaystyle= M10M11rhκ2ξ[10(13κ8ξ)27rh210κM10N113ξL102rh2+13(2ξκ)27M11M10rh40(ξ2κ)9L11L10rh10κ(ξ2κ)9ξL112L102],\displaystyle\frac{M_{10}-M_{11}r_{h}}{\kappa-2\xi}\Bigg{[}\frac{10(13\kappa-8\xi)}{27r_{h}^{2}}-\frac{10\kappa M_{10}N_{11}}{3\xi L_{10}^{2}r_{h}^{2}}+\frac{13(2\xi-\kappa)}{27}\frac{M_{11}}{M_{10}r_{h}}-\frac{40(\xi-2\kappa)}{9}\frac{L_{11}}{L_{10}r_{h}}-\frac{10\kappa(\xi-2\kappa)}{9\xi}\frac{L_{11}^{2}}{L_{10}^{2}}\Bigg{]},
N12\displaystyle N_{12} =\displaystyle= M11N113M104N113rh,\displaystyle\frac{M_{11}N_{11}}{3M_{10}}-\frac{4N_{11}}{3r_{h}},
L12\displaystyle L_{12} =\displaystyle= M11L113M104L113rh+2ξM11L103κM10rh2ξL103κrh2.\displaystyle\frac{M_{11}L_{11}}{3M_{10}}-\frac{4L_{11}}{3r_{h}}+\frac{2\xi M_{11}L_{10}}{3\kappa M_{10}r_{h}}-\frac{2\xi L_{10}}{3\kappa r_{h}^{2}}. (62)

Note that the analytical solution in Eqs. (9) and (10) corresponds to taking

rh=2M,\displaystyle r_{h}=2M,
N10=0,\displaystyle N_{10}=0,
N11=12M,\displaystyle N_{11}=\frac{1}{2M},
M10=2Me2μ0,\displaystyle M_{10}=2Me^{2\mu_{0}},
M11=e2μ0,\displaystyle M_{11}=e^{2\mu_{0}},
L10=λ0+λ12M,\displaystyle L_{10}=\lambda_{0}+\frac{\lambda_{1}}{2M},
L11=λ14M2.\displaystyle L_{11}=-\frac{\lambda_{1}}{4M^{2}}. (63)

Appendix D Curvature scalars at the event horizon

If N100N_{10}\neq 0, the metric expansion in Eq. (18) gives finite values for the curvature scalars R,RαβRαβR,\,R_{\alpha\beta}R^{\alpha\beta} and RαβγδRαβγδR_{{\alpha\beta\gamma\delta}}R^{{\alpha\beta\gamma\delta}} when δ=rrh\delta=r-r_{h} approaches zero. They are

R|r=rh=2rh22rhM10+M20N208M102N10+N2028M10N102N112M10N10,\displaystyle R|_{r=r_{h}}=\frac{2}{r_{h}^{2}}-\frac{2}{r_{h}M_{10}}+\frac{M_{20}N_{20}}{8M_{10}^{2}N_{10}}+\frac{N_{20}^{2}}{8M_{10}N_{10}^{2}}-\frac{N_{11}}{2M_{10}N_{10}},
RαβRαβ|r=rh=2rh42rh3M10+32rh2M102M20N208rhM103N10N2028rhM102N102+M202N202128M104N102+M20N20364M103N103+N204128M102N104\displaystyle R_{\alpha\beta}R^{\alpha\beta}|_{r=r_{h}}=\frac{2}{r_{h}^{4}}-\frac{2}{r_{h}^{3}M_{10}}+\frac{3}{2r_{h}^{2}M_{10}^{2}}-\frac{M_{20}N_{20}}{8r_{h}M_{10}^{3}N_{10}}-\frac{N_{20}^{2}}{8r_{h}M_{10}^{2}N_{10}^{2}}+\frac{M_{20}^{2}N_{20}^{2}}{128M_{10}^{4}N_{10}^{2}}+\frac{M_{20}N_{20}^{3}}{64M_{10}^{3}N_{10}^{3}}+\frac{N_{20}^{4}}{128M_{10}^{2}N_{10}^{4}}
+N112rhM102N10M20N11N2016M103N102N202N1116M102N103+N1128M102N102,\displaystyle\hskip 51.21504pt+\frac{N_{11}}{2r_{h}M_{10}^{2}N_{10}}-\frac{M_{20}N_{11}N_{20}}{16M_{10}^{3}N_{10}^{2}}-\frac{N_{20}^{2}N_{11}}{16M_{10}^{2}N_{10}^{3}}+\frac{N_{11}^{2}}{8M_{10}^{2}N_{10}^{2}},
RαβγδRαβγδ|r=rh=4rh4+2rh2M102+M20N20332M103N103+M202N20264M104N102+N20464M102N104M20N20N118M103N102N202N118M102N103+N1124M102N102.\displaystyle R_{{\alpha\beta\gamma\delta}}R^{{\alpha\beta\gamma\delta}}|_{r=r_{h}}=\frac{4}{r_{h}^{4}}+\frac{2}{r_{h}^{2}M_{10}^{2}}+\frac{M_{20}N_{20}^{3}}{32M_{10}^{3}N_{10}^{3}}+\frac{M_{20}^{2}N_{20}^{2}}{64M_{10}^{4}N_{10}^{2}}+\frac{N_{20}^{4}}{64M_{10}^{2}N_{10}^{4}}-\frac{M_{20}N_{20}N_{11}}{8M_{10}^{3}N_{10}^{2}}-\frac{N_{20}^{2}N_{11}}{8M_{10}^{2}N_{10}^{3}}+\frac{N_{11}^{2}}{4M_{10}^{2}N_{10}^{2}}. (64)

Note that M10M_{10} is always nonzero in our numerical solutions. The relations in Eqs. (59) and (61) as well as the corresponding recurrence relations for N11N_{11} might be used to rewrite the results in Eq. (64) for each family of solutions. While the rewriting does not bring any simplification for the first family of solutions (br=0b_{r}=0), for the second family of solutions (Rrr=0R_{rr}=0), the results in Eq. (64) remarkably simplify to

R|r=rh=2rh2,\displaystyle R|_{r=r_{h}}=\frac{2}{r_{h}^{2}},
RαβRαβ|r=rh=2rh42rh3M10+32rh2M102,\displaystyle R_{\alpha\beta}R^{\alpha\beta}|_{r=r_{h}}=\frac{2}{r_{h}^{4}}-\frac{2}{r_{h}^{3}M_{10}}+\frac{3}{2r_{h}^{2}M_{10}^{2}},
RαβγδRαβγδ|r=rh=4rh4+6rh2M102.\displaystyle R_{{\alpha\beta\gamma\delta}}R^{{\alpha\beta\gamma\delta}}|_{r=r_{h}}=\frac{4}{r_{h}^{4}}+\frac{6}{r_{h}^{2}M_{10}^{2}}. (65)

The coordinate scalar b2:=gμνbμbνb^{2}:=g_{\mu\nu}b^{\mu}b^{\nu} is also finite at rhr_{h}. It is

b2|r=rh=L102N10\displaystyle b^{2}|_{r=r_{h}}=-\frac{L_{10}^{2}}{N_{10}} (66)

for the first family of solutions, and

b2|r=rh=2M10ξrh1ξ+(κ2ξ)L1023κN10+(ξ2κ)rhL2026ξN10\displaystyle b^{2}|_{r=r_{h}}=\frac{2M_{10}}{\xi r_{h}}-\frac{1}{\xi}+\frac{(\kappa-2\xi)L_{10}^{2}}{3\kappa N_{10}}+\frac{(\xi-2\kappa)r_{h}L_{20}^{2}}{6\xi N_{10}} (67)

for the second family of solutions.

If N10=0N_{10}=0 and thus N2n=M2n=L2n=0N_{2n}=M_{2n}=L_{2n}=0 in Eq. (18), the finite values for the curvature scalars R,RαβRαβR,\,R_{\alpha\beta}R^{\alpha\beta} and RαβγδRαβγδR_{{\alpha\beta\gamma\delta}}R^{{\alpha\beta\gamma\delta}} are then

R|r=rh=2rh24rhM10+M112M1023N122M10N11,\displaystyle R|_{r=r_{h}}=\frac{2}{r_{h}^{2}}-\frac{4}{r_{h}M_{10}}+\frac{M_{11}}{2M_{10}^{2}}-\frac{3N_{12}}{2M_{10}N_{11}},
RαβRαβ|r=rh=2rh44rh3M10+4rh2M102M11rhM103+3N12rhM102N11+M1128M1043M11N124M103N11+9N1228M102N112,\displaystyle R_{\alpha\beta}R^{\alpha\beta}|_{r=r_{h}}=\frac{2}{r_{h}^{4}}-\frac{4}{r_{h}^{3}M_{10}}+\frac{4}{r_{h}^{2}M_{10}^{2}}-\frac{M_{11}}{r_{h}M_{10}^{3}}+\frac{3N_{12}}{r_{h}M_{10}^{2}N_{11}}+\frac{M_{11}^{2}}{8M_{10}^{4}}-\frac{3M_{11}N_{12}}{4M_{10}^{3}N_{11}}+\frac{9N_{12}^{2}}{8M_{10}^{2}N_{11}^{2}},
RαβγδRαβγδ|r=rh=4rh4+4rh2M102+M1124M1043M11N122M103N11+9N1224M102N112,\displaystyle R_{{\alpha\beta\gamma\delta}}R^{{\alpha\beta\gamma\delta}}|_{r=r_{h}}=\frac{4}{r_{h}^{4}}+\frac{4}{r_{h}^{2}M_{10}^{2}}+\frac{M_{11}^{2}}{4M_{10}^{4}}-\frac{3M_{11}N_{12}}{2M_{10}^{3}N_{11}}+\frac{9N_{12}^{2}}{4M_{10}^{2}N_{11}^{2}}, (68)

when δ\delta approaches zero. The recurrence relations for the first family of solutions do not bring any simplification for the results in Eq. (68). The recurrence relations for the second family of solutions simplify the results to

R|r=rh=2rh22rhM10,\displaystyle R|_{r=r_{h}}=\frac{2}{r_{h}^{2}}-\frac{2}{r_{h}M_{10}},
RαβRαβ|r=rh=2rh44rh3M10+2rh2M102,\displaystyle R_{\alpha\beta}R^{\alpha\beta}|_{r=r_{h}}=\frac{2}{r_{h}^{4}}-\frac{4}{r_{h}^{3}M_{10}}+\frac{2}{r_{h}^{2}M_{10}^{2}},
RαβγδRαβγδ|r=rh=4rh4+8rh2M102,\displaystyle R_{{\alpha\beta\gamma\delta}}R^{{\alpha\beta\gamma\delta}}|_{r=r_{h}}=\frac{4}{r_{h}^{4}}+\frac{8}{r_{h}^{2}M_{10}^{2}}, (69)

which recovers the results for the Schwarzschild metric by taking M10=rhM_{10}=r_{h}. With N10=0N_{10}=0, we have b2|r=rh=0b^{2}|_{r=r_{h}}=0 for the first family of solutions, and

b2|r=rh\displaystyle b^{2}|_{r=r_{h}} =\displaystyle= M10ξrh(4ξ+7κ)L1029κrhN11+2(2ξκ)M11L1029κM10N11\displaystyle\frac{M_{10}}{\xi r_{h}}-\frac{(4\xi+7\kappa)L_{10}^{2}}{9\kappa r_{h}N_{11}}+\frac{2(2\xi-\kappa)M_{11}L_{10}^{2}}{9\kappa M_{10}N_{11}} (70)
+(ξ2κ)rhL1123ξN112L10L11N111ξ\displaystyle+\frac{(\xi-2\kappa)r_{h}L_{11}^{2}}{3\xi N_{11}}-\frac{2L_{10}L_{11}}{N_{11}}-\frac{1}{\xi}

for the second family of solutions.

Appendix E Asymptotic expansion of the metric in the harmonic coordinates

The PPN metric expansion is constructed in the harmonic coordinates. For static spherical spacetime, it is Poisson and Will (2014)

g¯00=1+2Mr¯2β(Mr¯)2+,\displaystyle\bar{g}_{00}=-1+\frac{2M}{\bar{r}}-2\beta\left(\frac{M}{\bar{r}}\right)^{2}+...,
g¯ij=(1+2γMr¯)δij+,\displaystyle\bar{g}_{ij}=\left(1+2\gamma\frac{M}{\bar{r}}\right)\delta_{ij}+..., (71)

where r¯=x¯2+y¯2+z¯2\bar{r}=\sqrt{\bar{x}^{2}+\bar{y}^{2}+\bar{z}^{2}} is the Euclidean length of the position vector in the harmonic coordinates, and β\beta and γ\gamma are the PPN parameters that turn out to be 2+ν2/μ122+\nu_{2}/\mu_{1}^{2} and 11 to match the expansion in Eq. (13) when μ0=0\mu_{0}=0. In the following derivation, we show the match by transforming the metric expansion in Eq. (13) to the harmonic coordinates.

The transformation includes two steps: (i) from the coordinates (t,r,θ,ϕ)(t,\,r,\,\theta,\,\phi) to (t,r¯,θ,ϕ)(t,\,\bar{r},\,\theta,\,\phi), and (ii) then to the harmonic coordinates (t,x¯,y¯,z¯)(t,\,\bar{x},\,\bar{y},\,\bar{z}). Denoting

gr¯r¯=e2μ(drdr¯)2,\displaystyle g_{\bar{r}\bar{r}}=e^{2\mu}\left(\frac{dr}{d\bar{r}}\right)^{2}, (72)

the metric components in the harmonic coordinates are

g¯00=e2ν,\displaystyle\bar{g}_{00}=-e^{2\nu},
g¯ij=r¯x¯ir¯x¯jgr¯r¯+θx¯iθx¯jr2+ϕx¯iϕx¯jr2sin2θ,\displaystyle\bar{g}_{ij}=\frac{\partial\bar{r}}{\partial\bar{x}^{i}}\frac{\partial\bar{r}}{\partial\bar{x}^{j}}g_{\bar{r}\bar{r}}+\frac{\partial\theta}{\partial\bar{x}^{i}}\frac{\partial\theta}{\partial\bar{x}^{j}}r^{2}+\frac{\partial\phi}{\partial\bar{x}^{i}}\frac{\partial\phi}{\partial\bar{x}^{j}}r^{2}\sin^{2}\theta, (73)

where the Jacobian matrix for the transformation between (r¯,θ,ϕ)(\bar{r},\,\theta,\,\phi) and (x¯,y¯,z¯)(\bar{x},\,\bar{y},\,\bar{z}) has the usual elements

(r¯x¯r¯y¯r¯z¯θx¯θy¯θz¯ϕx¯ϕy¯ϕz¯)=(sinθcosϕsinθsinϕcosθ1r¯cosθcosϕ1r¯cosθsinϕ1r¯sinθ1r¯sinϕsinθ1r¯cosϕsinθ0).\displaystyle\begin{pmatrix}\frac{\partial\bar{r}}{\partial\bar{x}}&\frac{\partial\bar{r}}{\partial\bar{y}}&\frac{\partial\bar{r}}{\partial\bar{z}}\\ \frac{\partial\theta}{\partial\bar{x}}&\frac{\partial\theta}{\partial\bar{y}}&\frac{\partial\theta}{\partial\bar{z}}\\ \frac{\partial\phi}{\partial\bar{x}}&\frac{\partial\phi}{\partial\bar{y}}&\frac{\partial\phi}{\partial\bar{z}}\end{pmatrix}=\begin{pmatrix}\sin\theta\cos\phi&\sin\theta\sin\phi&\cos\theta\\ \frac{1}{\bar{r}}\cos\theta\cos\phi&\frac{1}{\bar{r}}\cos\theta\sin\phi&-\frac{1}{\bar{r}}\sin\theta\\ -\frac{1}{\bar{r}}\frac{\sin\phi}{\sin\theta}&\frac{1}{\bar{r}}\frac{\cos\phi}{\sin\theta}&0\end{pmatrix}. (74)

Calculating the inverse metric g¯αβ\bar{g}^{\alpha\beta} and the Christoffel symbols Γ¯αβλ\bar{\Gamma}^{\lambda}_{\phantom{\mu}\alpha\beta} in the harmonic coordinates, we find that the harmonic condition

g¯αβΓ¯αβλ=0\displaystyle\bar{g}^{\alpha\beta}\,\bar{\Gamma}^{\lambda}_{\phantom{\lambda}\alpha\beta}=0 (75)

simplifies to an ODE for r¯\bar{r} as a function of rr,

r¯′′+(νμ+2r)r¯2e2μr2r¯=0,\displaystyle\bar{r}^{\prime\prime}+\left(\nu^{\prime}-\mu^{\prime}+\frac{2}{r}\right)\bar{r}^{\prime}-\frac{2e^{2\mu}}{r^{2}}\bar{r}=0, (76)

where the prime denotes the derivative with respect to rr. With μ\mu and ν\nu expressed as the expansions in Eq. (13), Eq. (76) has the following asymptotic solutions

r¯=rs(1+c1r+c2r2+),\displaystyle\bar{r}=r^{s}\left(1+\frac{c_{1}}{r}+\frac{c_{2}}{r^{2}}+...\right), (77)

where the constant ss and the coefficients c1,c2,c_{1},\,c_{2},\,..., can be determined by substituting the solution back into Eq. (76). We find the indicial equation for ss to be

s2+s2e2μ0=0,\displaystyle s^{2}+s-2e^{2\mu_{0}}=0, (78)

and the coefficients to be

c1=s(μ1ν1)4e2μ0μ12s,\displaystyle c_{1}=\frac{s(\mu_{1}-\nu_{1})-4e^{2\mu_{0}}\mu_{1}}{2s},
c2=e2μ0[4e2μ0μ12+(14s)μ12+(2s1)μ1ν12sμ2]s(2s1)\displaystyle c_{2}=\frac{e^{2\mu_{0}}\left[4e^{2\mu_{0}}\mu_{1}^{2}+(1-4s)\mu_{1}^{2}+(2s-1)\mu_{1}\nu_{1}-2s\mu_{2}\right]}{s(2s-1)}
+(s1)(μ1ν1)2+4s(μ2ν2)4(2s1),\displaystyle\hskip 19.91684pt+\frac{(s-1)(\mu_{1}-\nu_{1})^{2}+4s(\mu_{2}-\nu_{2})}{4(2s-1)},
.\displaystyle...\,. (79)

Note that the two roots of Eq. (78) give two independent solutions whose linear combinations produce general solutions to Eq. (76). But for our purpose, it is sensible to have s=1s=1 when μ0=0\mu_{0}=0, suggesting to take the special solution corresponding to the root

s=1+1+8e2μ02.\displaystyle s=\frac{-1+\sqrt{1+8e^{2\mu_{0}}}}{2}. (80)

To eliminate rr in Eq. (73) so that it can be compared with Eq. (71), the inverse of Eq. (77) is required. We find

r=u(1c1s1uc2+1s2sc12s1u2+),\displaystyle r=u\left(1-\frac{c_{1}}{s}\frac{1}{u}-\frac{c_{2}+\frac{1-s}{2s}c_{1}^{2}}{s}\frac{1}{u^{2}}+...\right), (81)

where u=r¯1/su=\bar{r}^{1/s}. Focusing on the transformation when μ0=0\mu_{0}=0, we have s=1s=1 and Eq. (81) is simplified to

r=r¯(1+μ1r¯+μ2+ν2r¯2+),\displaystyle r=\bar{r}\left(1+\frac{\mu_{1}}{\bar{r}}+\frac{\mu_{2}+\nu_{2}}{{\bar{r}}^{2}}+...\right), (82)

where the relation ν1=μ1\nu_{1}=-\mu_{1}, which is valid for both families of solutions, has also been used. Using Eq. (82) to eliminate rr in Eq. (73) then leads to

g¯00=1+2μ1r¯4μ12+2ν2r¯2+,\displaystyle\bar{g}_{00}=-1+\frac{2\mu_{1}}{\bar{r}}-\frac{4\mu_{1}^{2}+2\nu_{2}}{\bar{r}^{2}}+...,
g¯ij=(1+2μ1r¯)δij+.\displaystyle\bar{g}_{ij}=\left(1+\frac{2\mu_{1}}{\bar{r}}\right)\delta_{ij}+...\,. (83)

With μ1=M\mu_{1}=M, we see

β=2+ν2μ12,γ=1.\displaystyle\beta=2+\frac{\nu_{2}}{\mu_{1}^{2}},\quad\gamma=1. (84)

Using the recurrence relations for ν2\nu_{2} in Eqs. (52) and (55), β\beta can be related to the free coefficients in each family of solutions.

References