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

Geodesic Motion of Test Particles around the Scalar Hairy Black Holes with Asymmetric Vacua

Hongyu Chen hongyuchen0306@qq.com School of Science, Jiangsu University of Science and Technology, 212100, Zhenjiang, Jiangsu Province, China    Wei Fan fanwei@just.edu.cn School of Science, Jiangsu University of Science and Technology, 212100, Zhenjiang, Jiangsu Province, China    Xiao Yan Chew xiao.yan.chew@just.edu.cn School of Science, Jiangsu University of Science and Technology, 212100, Zhenjiang, Jiangsu Province, China
Abstract

An asymptotically flat hairy black hole (HBH) can exhibit distinct characteristics when compared to the Schwarzschild black hole, due to the evasion of no-hair theorem by minimally coupling the Einstein gravity with a scalar potential which possesses asymmetric vacua, i.e, a false vacuum (ϕ=0)(\phi=0) and a true vacuum (ϕ=ϕ1)(\phi=\phi_{1}). In this paper, we investigate the geodesic motion of both massive test particles and photons in the vicinity of HBH with ϕ1=0.5\phi_{1}=0.5 and ϕ1=1.0\phi_{1}=1.0 by analyzing their effective potentials derived from the geodesic equation. By fixing ϕ1\phi_{1}, the effective potential of a massive test particle increases monotonically when its angular momentum LL is very small. When LL increases to a critical value, the effective potential possesses an inflection point which is known as the innermost stable of circular orbit (ISCO), where the test particle can still remain stable in a circular orbit with a minimal radius without being absorbed by the HBH or fleeing to infinity. Beyond the critical value of LL, the effective potential possesses a local minimum and a local maximum, indicating the existence of unstable and stable circular orbits, respectively. Moreover, the HBH possesses an unstable photon sphere but its location slightly deviates from the Schwarzschild black hole. The trajectories of null geodesics in the vicinity of HBH can also be classified into three types, which are the direct, lensing and photon sphere, based on the deflection angle of light, but the values of impact parameters can vary significantly than the Schwarzschild black hole.

I Introduction

In General Relativity (GR), the no-hair theorem describes that the characteristics of a black hole can be completely determined by only three global charges: mass, electric charge, and angular momentum [1, 2, 3], thus only electrovacuum black holes such as the Schwarzschild and Reissner-Nordstrom black holes can satisfy this theorem. However, there exists a class of black holes known as hairy black holes (HBHs), which manage to circumvent the no-hair theorem. The event horizon of these black holes are supported by nontrivial matter fields, allowing them to possess additional parameters referred to as ”hair”, which are associated to the corresponding theory. As a result, the solutions of HBHs can be bifurcated from the electrovacuum black holes, demonstrating unique features in the strong gravity regime, while becoming indistinguishable from them in the weak gravity regime.

In particular, various types of asymptotically flat HBHs have been constructed numerically in the Einstein-Klein-Gordon (EKG) theory [4, 5, 7, 6, 8, 9], where the Einstein gravity is minimally coupled to scalar potentials V(ϕ)V(\phi) without the inclusion of other matter fields. Hence, the existence of HBHs is fully determined by the characteristics of V(ϕ)V(\phi) to bypass the no-hair theorem [10]. In general, V(ϕ)V(\phi) cannot be strictly positive. It should become negative in some regions in order to violate the weak energy condition - one of the basic criteria of the no-hair theorem. For instance, a HBH is constructed from one V(ϕ)V(\phi) of asymmetric vacua [4, 5], which has the false and true vacuums to describe the first-order phase transition of our universe [11] and whose negative region is bounded by two roots of V(ϕ)V(\phi). Besides, that corresponding HBH has been generalized to become a charged HBH which can be bifurcated from the Reissner-Nordstrom black hole [6]. Furthermore, a HBH is constructed by V(ϕ)=λϕ4+μϕ2V(\phi)=-\lambda\phi^{4}+\mu\phi^{2} with λ\lambda and μ\mu are the positive constants, where the negative regions of V(ϕ)V(\phi) lies in the region |ϕ|>μ/λ|\phi|>\sqrt{\mu/\lambda} [7, 8]. The HBH can be connected smoothly with its counterpart gravitating scalaron in the small horizon limit [12]. Recently, a HBH can be constructed by V(ϕ)=α2ϕ6V(\phi)=-\alpha^{2}\phi^{6} where the scalar field is massless, thus it serves as a first step to explore the possibilities of existence of HBH with generalized V(ϕ)=α2ϕnV(\phi)=-\alpha^{2}\phi^{n} with nn is an integer [9]. On the other hand, other constructions of HBHs in the EKG theory can be found in Refs. [13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26]. Moreover, other classes of scalar HBHs can be constructed by minimally coupling a real scalar field with a U(1)-gauged complex scalar field [27, 28, 29] and Proca field associated with some potentials [30].

So far GR has successfully passed numerous observational tests with remarkable precision since the groundbreaking proposal by A. Einstein at 1915, ranging from the perihelion shift of mercury at 1920, the discovery of the binary pulsars by R. Hulse and J. Taylor [31], and then the recent detection of gravitational waves from the merger of binary compact objects by LIGO-VIRGO-KAGRA (LVK) collaboration and imaging of shadow for supermassive black holes in the galaxies M87 and Sgr A by Event Horizon Telescope (EHT). Hence, these observational tests are actually the basic manifestation of GR about the notion of how the spacetime around a compact object can be bent by its own gravity and how does the corresponding spacetime affect the motion of a test particle, as summarized by the famous physicist J. Wheeler in his quotation “Matter tells space how to curve. Space tells matter how to move”. Therefore, we study the influence of nontrivial scalar field on the geodesic motion of test particles in the vicinity of HBH with asymmetric vacua [4, 5] and report the results in this paper. Besides, the geodesic motion of test particles in the vicinity of other types of black holes have been extensively studied, for instance, Schwarzschild black hole [32, 33], Schwarzschild-(A)dS black hole [34], swirling universe [35, 36, 37, 38], Myers-Perry black hole [39], regular black hole by Ayón-Beato and Gracía [40], black hole in braneworld [41], (2+1)-rotating black holes [42, 43], SU(2)-(A)dS black hole in conformal gravity [44], supersymmetric AdS5\hbox{AdS}_{5} black hole [45], Kerr black hole [46, 47, 48, 49], Kerr-Newman black hole [50, 51, 52, 53, 54], U(1)2U(1)^{2} dyonic rotating black holes [55], Kerr-Newman-Taub-NUT black hole [56], black hole in the Einstein-Maxwell-dilaton-axion theory [57], polymer black hole in the loop quantum gravity [58], black hole in the Einsteinian cubic gravity [59], Carrollian Reissner-Nordstrom black hole [60], Euclidean Schwarzschild geometry [61]. The trajectories of test particles in these black holes can be described analytically by the elliptic functions, since the closed form of these black holes are known. However, we can only numerically obtain the trajectories of test particles in this paper since we couldn’t obtain the closed form of the HBH, which can only be constructed numerically by solving the Einstein equation.

This paper is organized as follows. In Sec. II, we briefly introduce some basic theoretical setups to construct the HBH and properties of the HBH. In Sec III, we introduce the Lagrangian, derive the effective potential for the test particles around the HBH and then introduce our approach to calculate the trajectories of particles. In Sec IV, we present and discuss our numerical findings. Finally, in Sec. V, we summarize our work and present an outlook.

II Scalar Hairy Black Holes with Asymmetric Vacua

In the EKG theory, we consider a scalar potential V(ϕ)V(\phi) with asymmetric vacua [4, 5] minimally coupled with the Einstein gravity,

S=d4xg[R16πG12μϕμϕV(ϕ)],S=\int d^{4}x\sqrt{-g}\left[\frac{R}{16\pi G}-\frac{1}{2}\nabla_{\mu}\phi\nabla^{\mu}\phi-V(\phi)\right]\,, (1)

where the explicit form of V(ϕ)V(\phi) is given by

V(ϕ)=V012(ϕa)2[3(ϕa)24(ϕa)(ϕ0+ϕ1)+6ϕ0ϕ1],V(\phi)=\frac{V_{0}}{12}\left(\phi-a\right)^{2}\left[3\left(\phi-a\right)^{2}-4(\phi-a)(\phi_{0}+\phi_{1})+6\phi_{0}\phi_{1}\right]\,, (2)

with aa, V0V_{0}, ϕ0\phi_{0}, and ϕ1\phi_{1} being constants. The asymptotic value of ϕ\phi at the spatial infinity is fixed by ϕ=a\phi=a, and we choose a=0a=0 such that ϕ=0\phi=0 at the spatial infinity in this paper. Hence, as shown in Fig. 1, V(ϕ)V(\phi) possesses a local minimum with V(0)=0V(0)=0 when ϕ=0\phi=0. V(ϕ)V(\phi) also possesses a local maximum at ϕ=ϕ0\phi=\phi_{0} and a global minimum at ϕ=ϕ1\phi=\phi_{1}. In the past decades V(ϕ)V(\phi) has been employed to study the first-order phase transition of our universe from the false vacuum (ϕ=0)(\phi=0) to the true vacuum (ϕ=ϕ1)(\phi=\phi_{1}).

Refer to caption

Figure 1: The authors have considered the scalar potential V(ϕ)V(\phi) with a false vacuum at ϕ=0\phi=0, a barrier at ϕ=ϕ0\phi=\phi_{0} and a true vacuum at ϕ=ϕ1\phi=\phi_{1} to construct the hairy black holes [5].

The negative region of V(ϕ)V(\phi) shown in Fig. 1 can violate the weak energy condition, which is one of the basic assumptions in the no-hair theorem. Hence, a few decades ago a class of HBHs can be constructed by Eq. (2) [4] to evade the no-hair theorem, and recently its properties such as the Hawking temperature, Ricci scalar, Kretschmann scalar and so on just have been studied systematically [5]. Moreover, it can be bifurcated from the Reissner-Nordstrom black hole to possess an electric charge by minimally coupling the Maxwell field to the EKG theory [6]. Besides, Eq. (2) has been employed to construct a fermionic star and its tidal deformability just has been investigated recently [62, 63].

Note that the asymmetrical profile of V(ϕ)V(\phi) is caused by the appearance of cubic term ϕ3\phi^{3}, thus V(ϕ)V(\phi) can become symmetric if the cubic term disappears. For instance, the authors recently have employed a symmetric profile of V(ϕ)=λϕ4+μϕ2V(\phi)=-\lambda\phi^{4}+\mu\phi^{2} (λ\lambda and μ\mu are the positive constants), which possesses two degenerate global maxima and a local minimum at ϕ=0\phi=0 to construct the HBH [8] and gravitating scalaron [12].

We then obtain the Einstein equation and Klein-Gordon (KG) equation from the variation of Eq. (1) with respect to the metric and scalar field, respectively

Rμν12gμνR=8πG(12gμναϕαϕgμνV+μϕνϕ),μμϕ=dVdϕ.R_{\mu\nu}-\frac{1}{2}g_{\mu\nu}R=8\pi G\left(-\frac{1}{2}g_{\mu\nu}\nabla_{\alpha}\phi\nabla^{\alpha}\phi-g_{\mu\nu}V+\nabla_{\mu}\phi\nabla_{\nu}\phi\right)\,,\quad\nabla_{\mu}\nabla^{\mu}\phi=\frac{dV}{d\phi}\,. (3)

The following spherically symmetric metric is employed as an Ansatz to construct the numerical solutions of HBH,

ds2=N(r)e2σ(r)dt2+dr2N(r)+r2(dθ2+sin2θdφ2),ds^{2}=-N(r)e^{-2\sigma(r)}dt^{2}+\frac{dr^{2}}{N(r)}+r^{2}\left(d\theta^{2}+\sin^{2}\theta d\varphi^{2}\right)\,, (4)

where N(r)=12m(r)/rN(r)=1-2m(r)/r with m(r)m(r) is the Misner-Sharp mass function [64]. The ADM mass of a HBH can be read off simply by the condition m()=Mm(\infty)=M. The direct substitution of Eq. (4) into the Eq. (3) yields a set of nonlinear ordinary differential equations (ODEs) for the functions,

m=2πG(Nϕ2+2V),σ=4πGrϕ2,(eσr2Nϕ)=eσr2dVdϕ,m^{\prime}=2\pi G\left(N\phi^{\prime 2}+2V\right)\,,\quad\sigma^{\prime}=-4\pi Gr\phi^{\prime 2}\,,\quad\left(e^{-\sigma}r^{2}N\phi^{\prime}\right)^{\prime}=e^{-\sigma}r^{2}\frac{dV}{d\phi}\,, (5)

where the prime denotes the derivative of the functions with respect to the radial coordinate rr. Here we only consider the solutions of HBH for the range of rr from the horizon radius rHr_{H} to infinity. Although the form of ODEs looks very simple, it is almost impossible to obtain the analytical solutions for the HBH, hence we integrate the ODEs numerically with the appropriate boundary conditions by the professional ODE solver package Colsys [65], which adopts the Newton-Raphson method to solve the boundary value problem for a set of nonlinear ODEs by utilizing the adaptive mesh refinement to greatly enhance the accuracy of solutions with more than 1000 points, and estimating the errors of solutions for the users’ reference.

Here we impose the boundary conditions for the ODEs at the horizon and infinity. All functions are required to be finite at the horizon, which can be expressed in term of the power series expansion with few leading terms,

m(r)\displaystyle m(r) =rH2+m1(rrH)+O((rrH)2),\displaystyle=\frac{r_{H}}{2}+m_{1}(r-r_{H})+O\left((r-r_{H})^{2}\right)\,, (6)
σ(r)\displaystyle\sigma(r) =σH+σ1(rrH)+O((rrH)2),\displaystyle=\sigma_{H}+\sigma_{1}(r-r_{H})+O\left((r-r_{H})^{2}\right)\,, (7)
ϕ(r)\displaystyle\phi(r) =ϕH+ϕH,1(rrH)+O((rrH)2),\displaystyle=\phi_{H}+\phi_{H,1}(r-r_{H})+O\left((r-r_{H})^{2}\right)\,, (8)

where

m1=4πGrH2V(ϕH),σ1=4πGrHϕH,12,ϕH,1=rHdV(ϕH)dϕ18πGrH2V(ϕH),m_{1}=4\pi Gr^{2}_{H}V(\phi_{H})\,,\quad\sigma_{1}=-4\pi Gr_{H}\phi^{2}_{H,1}\,,\quad\phi_{H,1}=\frac{r_{H}\frac{dV(\phi_{H})}{d\phi}}{1-8\pi Gr_{H}^{2}V(\phi_{H})}\,, (9)

Here σH\sigma_{H} and ϕH\phi_{H} are the values of σ(r),ϕ(r)\sigma(r),\phi(r) at the horizon respectively. The denominator of ϕH,1\phi_{H,1} has to fulfill the condition 18πGrH2V(ϕH)01-8\pi Gr_{H}^{2}V(\phi_{H})\neq 0 such that σ(r)\sigma(r) and ϕ(r)\phi(r) are finite at the horizon. At infinity all functions have to satisfy the asymptotically flat condition, which are given by m(r)=Mm(r\rightarrow\infty)=M with MM is the ADM mass of the HBH, σ(r)=ϕ(r)=0\sigma(r\rightarrow\infty)=\phi(r\rightarrow\infty)=0. Furthermore, we perform the numerical calculation in the compactified coordinate x[0,1]x\in[0,1] via the transformation x=1rH/rx=1-r_{H}/r, which maps the horizon and infinity to the numerical value 0 and 11 respectively. Besides, we introduce the dimensionless parameters: rr/βr\rightarrow r/\sqrt{\beta}, mm/βm\rightarrow m/\sqrt{\beta}, ϕβϕ\phi\rightarrow\sqrt{\beta}\phi, ϕ1βϕ1\phi_{1}\rightarrow\sqrt{\beta}\phi_{1}, ϕ0βϕ0\phi_{0}\rightarrow\sqrt{\beta}\phi_{0} and VβVV\rightarrow\sqrt{\beta}V. Therefore, the remaining free parameters in the calculation are ϕ0\phi_{0}, ϕ1\phi_{1}, rHr_{H}, σH\sigma_{H}, ϕH\phi_{H}, and MM. The parameters σH\sigma_{H} and MM can be determined exactly when the solutions of HBH satisfy the boundary conditions, so the input parameters for the calculation are given by ϕ0\phi_{0}, ϕ1\phi_{1}, rHr_{H}, and ϕH\phi_{H}.

(a) Refer to caption (b) Refer to caption (c) Refer to caption (d) Refer to caption

Figure 2: Two basic properties of HBH with ϕ1=0.5\phi_{1}=0.5 and ϕ1=1.0\phi_{1}=1.0 when the scalar field ϕH\phi_{H} is nontrivial at the horizon rHr_{H} : (a) The reduced area of horizon aHa_{H} vs ϕH\phi_{H} (b) The reduced Hawking temperature vs ϕH\phi_{H}; (c) The parameter ϕ0\phi_{0} as the function of ϕH\phi_{H} for ϕ1=0.5\phi_{1}=0.5 and ϕ1=1.0\phi_{1}=1.0; (d) The solutions m(x)m(x), σ(x)\sigma(x) and ϕ(x)\phi(x) of the HBH with ϕH=0.7,ϕH=1.0,rH=1.0\phi_{H}=0.7,\phi_{H}=1.0,r_{H}=1.0 in the compactified coordinate xx.

Since we have 4 free parameters to describe our HBHs, it would be convenient when we study the geodesics motion of test particles around the HBHs with rH=1r_{H}=1 by fixing ϕ1=0.5,1.0\phi_{1}=0.5,1.0 for simplicity. Hence, the basic properties of HBH can be briefly summarized in Fig. 2, where we have defined the reduced area of horizon aH=AH/(16πM2)a_{H}=A_{H}/(16\pi M^{2}) and reduced Hawking temperature tH=8πTHMt_{H}=8\pi T_{H}M with the area of horizon AH=4πrH2A_{H}=4\pi r^{2}_{H} and Hawking temperature TH=14πN(rH)eσHT_{H}=\frac{1}{4\pi}N^{\prime}(r_{H})e^{-\sigma_{H}}. The purpose to introduce such quantities is to easily see the connection between our HBH model with the Schwarzschild black hole, since the trivial solution to the Eq. (3) is the Schwarzschild black hole when the scalar field at the horizon ϕH\phi_{H} doesn’t exist. Recall that the values of aHa_{H} and tHt_{H} are unity for Schwarzchild black hole when ϕH=0\phi_{H}=0. However, when ϕH\phi_{H} increases, the HBH bifurcates from the Schwarzschild black hole where aHa_{H} decreases from unity but tHt_{H} increases from unity. In the limit ϕH=ϕ1\phi_{H}=\phi_{1}, aHa_{H} reaches to zero but tHt_{H} increases very sharply, therefore the HBH could disappear in that limit.

Figs. 2(c) demonstrates that ϕ0\phi_{0} increases monotonically as ϕH\phi_{H} increases from zero for ϕ1=0.5,1.0\phi_{1}=0.5,1.0, and is almost indistinguishable for small ϕH\phi_{H}. We also find the values of ϕ0<<ϕ1\phi_{0}<<\phi_{1} in the limit ϕH=ϕ1\phi_{H}=\phi_{1}. Fig. 2(d) shows the typical profiles of the solutions of HBH with ϕH=0.7\phi_{H}=0.7, ϕ1=1.0\phi_{1}=1.0 and rH=1r_{H}=1 in the compactified coordinate xx. The functions behave almost constant inside the bulk, corresponding to the global minimum of V(ϕ)V(\phi) which is the true vacuum ϕ1\phi_{1}. However, a sharp boundary developed at some intermediate region of the spacetime, when the solutions move away from the horizon, to drastically change to a new set of almost constant functions. This region corresponds to the false vacuum (ϕ=0)(\phi=0) at infinity, where ϕ\phi sits at the local minimum of V(ϕ)V(\phi).

III Geodesic Motion of Test Particles Around the HBH

The geodesic motion of test particles in the vicinity of HBH can be studied in the Lagrangian formalism, where the Lagrangian is given by

=12x˙μx˙μ=ϵ,\mathscr{L}=\frac{1}{2}\dot{x}^{\mu}\dot{x}_{\mu}=\epsilon\,, (10)

where ϵ=0\epsilon=0 refers to massless particle and ϵ=1\epsilon=-1 refers to massive particle. The dot denotes the derivative of a function with respect to the affine parameter λ\lambda. Since the spacetime of HBH is static and stationary, it possesses two conserved quantities which are the energy EE and angular momentum LL,

E=t˙=e2σNt˙,L=φ˙=r2φ˙.E=-\frac{\partial\mathscr{L}}{\partial\dot{t}}=e^{-2\sigma}N\dot{t}\,,\quad L=\frac{\partial\mathscr{L}}{\partial\dot{\varphi}}=r^{2}\dot{\varphi}\,. (11)

Since our HBH is spherically symmetric, then we just have to concentrate on the motion of particles on the equatorial plane (θ=π/2)(\theta=\pi/2). Therefore, we can obtain the radial equation r˙\dot{r} as shown in the below,

e2σr˙2=E2Veff(r),e^{-2\sigma}\dot{r}^{2}=E^{2}-V_{\text{eff}}(r)\,, (12)

where the effective potential VeffV_{\text{eff}} of a test particle is given by

Veff(r)=e2σN(L2r2ϵ).V_{\text{eff}}(r)=e^{-2\sigma}N\left(\frac{L^{2}}{r^{2}}-\epsilon\right)\,. (13)

There are several types of orbits can exist in the spacetime of HBH by analyzing Veff(r)V_{\text{eff}}(r), such as the circular orbit, bound orbit, escape orbit and etc. In general, a test particle is in a circular orbit when it is moving in a circular motion with a fixed radius from the HBH. It can be determined from the condition Veff(r)=0V^{\prime}_{\text{eff}}(r)=0. If Veff(r)V_{\text{eff}}(r) possesses a local minimum, then the circular orbit is stable but the circular orbit is unstable if Veff(r)V_{\text{eff}}(r) possesses a local maximum. A test particle is in a bound orbit where it can travel between the minimal radius and maximal radius from the HBH with E<1E<1. However, a test particle can flee to the infinity from a point in HBH when E>1E>1. If Veff(r)V_{\text{eff}}(r) contains an inflection point which connects the curves with concave downward and concave upward, this implies the test particle is located in the innermost stable circular orbit (ISCO), which can be determined by the condition Veff(r)=Veff′′(r)=0V^{\prime}_{\text{eff}}(r)=V^{\prime\prime}_{\text{eff}}(r)=0, and the explicit form is given by

Veff(r)\displaystyle V^{\prime}_{\text{eff}}(r) =(NN2σ)Veff2L2e2σNr3,\displaystyle=\left(\frac{N^{\prime}}{N}-2\sigma^{\prime}\right)V_{\text{eff}}-\frac{2L^{2}e^{-2\sigma}N}{r^{3}}\,, (14)
Veff′′(r)\displaystyle V^{\prime\prime}_{\text{eff}}(r) =(N′′N4σNN+4Nσ22σ′′)Veff+2L2e2σ(3Nr4+4Nσr32Nr3).\displaystyle=\left(\frac{N^{\prime\prime}}{N}-\frac{4\sigma^{\prime}N^{\prime}}{N}+4N\sigma^{\prime 2}-2\sigma^{\prime\prime}\right)V_{\text{eff}}+2L^{2}e^{-2\sigma}\left(\frac{3N}{r^{4}}+\frac{4N\sigma^{\prime}}{r^{3}}-\frac{2N^{\prime}}{r^{3}}\right)\,. (15)

The location of ISCO, rISCOr_{\text{ISCO}} can be calculated numerically by solving the above equations.

Moreover, we derive the following expression,

dφdr=φ˙r˙=Lr2eσE2Veff(r),\frac{d\varphi}{dr}=\frac{\dot{\varphi}}{\dot{r}}=\frac{L}{r^{2}e^{\sigma}\sqrt{E^{2}-V_{\text{eff}}(r)}}\,, (16)

which can allow us to obtain the trajectories of test particles by numerically integrating the above expression, as shown in the below,

φ(r)=rminrmaxLr2eσE2Veff(r)𝑑r.\varphi(r)=\int_{r_{\text{min}}}^{r_{\text{max}}}\frac{L}{r^{2}e^{\sigma}\sqrt{E^{2}-V_{\text{eff}}(r)}}dr\,. (17)

The above integration range [rmin,rmax][r_{\text{min}},r_{\text{max}}] is only valid when E2Veff(r)0E^{2}-V_{\text{eff}}(r)\geq 0. Then we can present their trajectories in the two-dimensional Cartesian coordinate using the following relation,

x=rcos(φ(r)),y=rsin(φ(r)).x=r\cos(\varphi(r))\,,\quad y=r\sin(\varphi(r))\,. (18)

For photon (ϵ=0)(\epsilon=0), we can introduce the parameter bL/Eb\equiv L/E which is known as the impact parameter in the radial equation r˙\dot{r},

e2σr˙2=E2(1b2veff(r)),e^{-2\sigma}\dot{r}^{2}=E^{2}\left(1-b^{2}v_{\text{eff}}(r)\right)\,, (19)

where

veff(r)=e2σNr2.v_{\text{eff}}(r)=\frac{e^{-2\sigma}N}{r^{2}}\,. (20)

The extremum of veff(r)v_{\text{eff}}(r) indicates the presence of a photon sphere, where the null geodesics can theoretically orbit arbitrary amount of time. Since the imaging of shadow for the supermassive black holes in the center of some galaxies can also be described by the framework of backward ray-tracing, therefore it might be convenient to show the trajectories of null geodesics around the HBH according to the total number of orbits of null geodesics nn, which can be defined as [66]

n=φ2π.n=\frac{\varphi}{2\pi}\,. (21)

This allows us to classify the trajectories of null geodesics into three types: direct emission with 0<n<3/40<n<3/4 for b(0,b2)(b2+,)b\in\left(0,b_{2}^{-}\right)\cup\left(b_{2}^{+},\infty\right) where the null geodesics only intersect with the accretion disk once (m=1)(m=1); lensing with 3/4<n<5/43/4<n<5/4 for b(b2,b3)(b3+,b2+)b\in\left(b_{2}^{-},b_{3}^{-}\right)\cup\left(b_{3}^{+},b_{2}^{+}\right) where the null geodesics intersect with the accretion disk twice (m=2)(m=2); and photon sphere with n>5/4n>5/4 for b(b3,b3+)b\in\left(b_{3}^{-},b_{3}^{+}\right) where the null geodesics intersect with the accretion disk three times (m=3)(m=3). Note that bm<bcb_{m}^{-}<b_{c} and bm+>bcb_{m}^{+}>b_{c}, where mm represents the number of intersections of null geodesics with the accretion disk and bcb_{c} is the critical value of bb for the photon sphere.

IV Results and Discussions

(a) Refer to caption (b) Refer to caption (c) Refer to caption (d) Refer to caption (e)    Refer to caption  (f)    Refer to caption

Figure 3: The effective potential Veff(x)V_{\text{eff}}(x) in the compactified coordinate x=1rH/rx=1-r_{H}/r for a massive particle around the HBH with rH=1r_{H}=1 for a) ϕ1=0.5\phi_{1}=0.5 and ϕH=0.3\phi_{H}=0.3. Note that in the figure, “Orbit 1” and “Orbit 2” refer to the motion of particle with E2=0.95E^{2}=0.95 in (e) and E2=1.2E^{2}=1.2 in (f), respectively; b) ϕ1=0.5\phi_{1}=0.5 and ϕH=0.4999\phi_{H}=0.4999; c) ϕ1=1.0\phi_{1}=1.0 and ϕH=0.3\phi_{H}=0.3; d) ϕ1=1.0\phi_{1}=1.0 and ϕH=0.995\phi_{H}=0.995; e) The bound orbit of a massive particle with E2=0.95E^{2}=0.95, labeled by “Orbit 1” in (a) around the HBH with ϕ1=0.5\phi_{1}=0.5, ϕH=0.3\phi_{H}=0.3, rH=1r_{H}=1; f) A massive particle with E2=1.2E^{2}=1.2, labeled by “Orbit 2” in (a) fleeing away from the HBH with ϕ1=0.5\phi_{1}=0.5, ϕH=0.3\phi_{H}=0.3, rH=1r_{H}=1. Note that the black disc represents the HBH.

We present our research findings for the investigation of geodesic motion of test particles around the HBH with rH=1r_{H}=1 by fixing ϕ1=0.5\phi_{1}=0.5 and ϕ1=1.0\phi_{1}=1.0. Thus, we begin with the massive test particle by analyzing its effective potentials Veff(x)V_{\text{eff}}(x) with several values of the angular momentum squared L2L^{2} in the compactified coordinate x=1rH/rx=1-r_{H}/r as shown in Figs. 3(a) for ϕH=0.3\phi_{H}=0.3 and ϕ1=0.5\phi_{1}=0.5. When we increase L2=0L^{2}=0 until LISCO2=3.137L^{2}_{\text{ISCO}}=3.137 (red dashed curve), Veff(x)V_{\text{eff}}(x) changes its monotonically increasing behaviour until possessing an inflection point, which is known as the location of ISCO, denoted as xISCOx_{\text{ISCO}} when LISCO2=3.137L^{2}_{\text{ISCO}}=3.137, causing Veff(xISCO)=Veff′′(xISCO)=0V^{\prime}_{\text{eff}}(x_{\text{ISCO}})=V^{\prime\prime}_{\text{eff}}(x_{\text{ISCO}})=0. This implies that the massive test particle can stay in the ISCO without being absorbed by HBH or escaped to infinity. When L2>3.137L^{2}>3.137, Veff(x)V_{\text{eff}}(x) starts to develop a local maximum and a local minimum which indicate the massive test particle possesses unstable and stable circular orbits, respectively. Note that the location of unstable circular orbit is always less than the location of stable circular orbit, possibly due to the strong gravitational pull by the HBH causing the instability of the circular orbit in the bulk region.

Fig. 3 (b) shows that Veff(x)V_{\text{eff}}(x) of a massive test particle around the HBH with ϕH=0.4999\phi_{H}=0.4999 (in the limit ϕH=ϕ1\phi_{H}=\phi_{1}), which also behaves qualitatively similar to Fig. 3(a). It possesses the ISCO when we increase L2L^{2} from zero until LISCO2=3.221L^{2}_{\text{ISCO}}=3.221 and two circular orbits which are stable and unstable when L2>3.221L^{2}>3.221. Recall that the values of xISCO=2/30.6667x_{\text{ISCO}}=2/3\approx 0.6667 [67] and LISCO2=3rH2L^{2}_{\text{ISCO}}=3r^{2}_{H} for the Schwarzschild black hole, hence we find that xISCOx_{\text{ISCO}} and LISCO2L^{2}_{\text{ISCO}} of the HBH with ϕ1=0.5\phi_{1}=0.5 don’t deviate too much from the Schwarzschild black hole. Other values of xISCOx_{\text{ISCO}} and LISCO2L^{2}_{\text{ISCO}} correspond to different ϕH\phi_{H} can be found in Table. 1.

ϕH\phi_{H} 0 (Schwarzschild) 0.2 0.4 0.45 0.4999
xISCOx_{\text{ISCO}} 0.6667 0.6607 0.6428 0.6398 0.6509
LISCO2L^{2}_{\text{ISCO}} 3.0000 3.0545 3.2200 3.2340 3.2110
Table 1: Several values of xISCOx_{\text{ISCO}} and LISCO2L^{2}_{\text{ISCO}} correspond to a massive test particle around the HBH with rH=1r_{H}=1 and ϕ1=0.5\phi_{1}=0.5.

Figs. 3 (c) and (d) exhibit Veff(x)V_{\text{eff}}(x) of a massive test particle around the HBH with ϕ1=1.0\phi_{1}=1.0 for ϕH=0.3\phi_{H}=0.3 and ϕH=0.995\phi_{H}=0.995 (in the limit ϕH=ϕ1\phi_{H}=\phi_{1}), respectively. We observe that the profiles of Veff(x)V_{\text{eff}}(x) in these two figures also behave qualitatively similar to Figs. 3(a) and (b), where they possess the ISCO when L2=LISCO2L^{2}=L^{2}_{\text{ISCO}} and two circular orbits which are stable and unstable when L2>LISCO2L^{2}>L^{2}_{\text{ISCO}}. However, we find that the deviation of LISCO2L^{2}_{\text{ISCO}} and xISCOx_{\text{ISCO}} of the HBH from the Schwarzschild black hole can become larger when ϕH\phi_{H} increases, as shown in Table. 2. Besides, Figs. 3 (a)-(d) show that VeffV_{\text{eff}}(x) approaches to unity when xx approaches to infinity, since the HBH is asymptotically flat.

ϕH\phi_{H} 0 (Schwarzschild) 0.2 0.4 0.6 0.9 0.95 0.995
xISCOx_{\text{ISCO}} 0.6667 0.6624 0.6680 0.6828 0.5731 0.5644 0.5671
LISCO2L^{2}_{\text{ISCO}} 3.0000 3.0988 3.5276 4.5204 5.7238 5.6721 5.1242
Table 2: Several values of xISCOx_{\text{ISCO}} and LISCO2L^{2}_{\text{ISCO}} correspond to a massive test particle around the HBH with rH=1r_{H}=1 and ϕ1=1.0\phi_{1}=1.0.

Moreover, Fig. 3 (e) demonstrates that a massive test particle with energy E2=0.95E^{2}=0.95, labeled by “Orbit 1” in Fig. 3 (a) can move in a bound orbit around the HBH with ϕ1=0.5\phi_{1}=0.5, ϕH=0.3\phi_{H}=0.3 and rH=1.0r_{H}=1.0. Nevertheless, it can flee to infinity with E2=1.2>1E^{2}=1.2>1, labeled by “Orbit 2” in Fig. 3 (a) from the same HBH, as shown in Fig. 3 (f). These two figures can be obtained by numerically integrating Eq. (17).

Fig. 4 (a) and (b) shows the effective potential veff(x)v_{\text{eff}}(x) of the null geodesics in the compactified coordinate xx in the HBH with ϕ1=0.5\phi_{1}=0.5 and ϕ1=1.0\phi_{1}=1.0, respectively. Both figures depict that veff(x)v_{\text{eff}}(x) possesses a local maximum, which indicates that the photon sphere is unstable. Recall that the location of photon sphere for the Schwarzchild black hole in the compactified coordinate xx is given by xph=1/3x_{\text{ph}}=1/3 [68], thus Fig. 4(c) shows the location of photon sphere for HBH with ϕ1=0.5\phi_{1}=0.5 and ϕ1=1.0\phi_{1}=1.0 don’t deviate too much from the Schwarzschild black hole. Fig. 4 (d) compares the total number of orbit nn for the Schwarzschild black hole and the HBH with ϕ1=0.5\phi_{1}=0.5. Although the trajectories of null geodesics around the HBH still can be divided into three types: direct, lensing and photon sphere, we observe that when ϕH\phi_{H} increases, the values of impact parameters b2b_{2}^{-}, b2+b_{2}^{+}, b3b_{3}^{-} and b3+b_{3}^{+} associated to different types of trajectories as shown in Table 3 for the HBH can increase significantly from the Schwarzcshild black hole, this can impact the brightness and size of the optical appearance (shadow) of the HBH. Besides, the three types of trajectories of null geodesics around the HBH with ϕ1=0.5\phi_{1}=0.5, ϕH=0.3\phi_{H}=0.3, rH=1.0r_{H}=1.0 can be visualized in Fig. 4 (e).

ϕH\phi_{H} bcb_{c} b2b^{-}_{2} b2+b^{+}_{2} b3b^{-}_{3} b3+b^{+}_{3}
0 (Schwarzschild) 2.5981 2.5005 3.0330 2.5860 2.6050
0.2 2.6391 2.5490 3.0996 2.6350 2.6546
0.4 2.8100 2.7128 3.3219 2.8056 2.8269
0.45 2.9223 2.8205 3.4652 2.9177 2.9402
0.4999 3.9452 3.7951 4.7619 3.9383 3.9716
Table 3: The values of impact parameters associated to different types of the null geodesics around the HBH with ϕ1=0.5\phi_{1}=0.5, rH=1.0r_{H}=1.0.

(a) Refer to caption (b) Refer to caption (c) Refer to caption (d) Refer to caption (e) Refer to caption

Figure 4: The effective potential veff(x)v_{\text{eff}}(x) in the compactified coordinate x=1rH/rx=1-r_{H}/r for null geodesics around the HBH with rH=1r_{H}=1 for a) ϕ1=0.5\phi_{1}=0.5; b) ϕ1=1.0\phi_{1}=1.0; c) The location of photon sphere in the compactified coordinate xx around the HBH with rH=1r_{H}=1, ϕ1=0.5\phi_{1}=0.5 and ϕ1=1.0\phi_{1}=1.0 versus ϕH\phi_{H}; d) Based on n=φ/(2π)n=\varphi/(2\pi), the null geodesics around the Schwarzschild black hole with rH=1r_{H}=1 and HBH with ϕ1=0.5\phi_{1}=0.5, rH=1r_{H}=1 can be categorized into three emissions as a function of impact parameter bb: direct (black), lensing (yellow) and photon ring (red). e) The trajectories of null geodesics consist of direct (black curves), lensing (yellow curves) and photon sphere (blue curve) around the HBH with ϕ1=0.5\phi_{1}=0.5, ϕH=0.3\phi_{H}=0.3, rH=1r_{H}=1.

V Conclusion and Outlook

In conclusion, we have studied the geodesic motion of test particles around a spherically symmetric and asymptotically flat hairy black hole (HBH) in the Einstein-Klein-Gordon (EKG) theory. The scalar potential V(ϕ)V(\phi) contains asymmetric vacua with the true vacuum being ϕ1\phi_{1} and the false vacuum being ϕ=0\phi=0. When the scalar field ϕH\phi_{H} is nontrivial at the horizon for 0ϕH<ϕ10\leq\phi_{H}<\phi_{1}, the solutions of HBH can exhibit some new properties different from the Schwarzschild black hole, as a consequence from the evasion of the no-hair theorem. In our investigation, we fix ϕ1=0.5,1.0\phi_{1}=0.5,1.0. Since the spacetime of HBH is spherically symmetric, we can obtain the effective potential Veff(r)V_{\text{eff}}(r) of test particles as a function of radial coordinate rr, and this allows us to focus on the motion of test particles on the equatorial plane. The analysis on the profiles of Veff(r)V_{\text{eff}}(r) could allow us to know the possible orbits of test particles and the integration of the geodesic equation allows us to obtain their trajectories.

For the case of a massive test particle around the HBHs with rH=1r_{H}=1 and ϕ1=0.5,1.0\phi_{1}=0.5,1.0, initially Veff(r)V_{\text{eff}}(r) is a strictly increasing function when the angular momentum squared L2L^{2} is very small. However, the increase of L2L^{2} can change this monotonically increasing behaviour, until it possesses an inflection point when L2L^{2} reaches to a critical value, LISCO2L^{2}_{\text{ISCO}} which is known as the innermost stable circular orbit (ISCO) where the particle can stay in a circular orbit with a minimal radius rISCOr_{\text{ISCO}} without being absorbed by the HBH and escaping to infinity. The deviation for the values of LISCO2L^{2}_{\text{ISCO}} and rISCOr_{\text{ISCO}} of the HBHs from the Schwarzschild black hole can be increased as ϕH\phi_{H} increases for large ϕ1\phi_{1}. When L2L^{2} exceeds LISCO2L^{2}_{\text{ISCO}}, Veff(r)V_{\text{eff}}(r) develops a local maximum and local minimum, indicating the presence of unstable and stable circular orbits, respectively. Besides, the particle can travel in a bound orbit and an escape orbit.

For the photon around the HBHs with rH=1r_{H}=1 and ϕ1=0.5,1.0\phi_{1}=0.5,1.0, its Veff(r)V_{\text{eff}}(r) possesses a local maximum, which indicates the presence of unstable photon sphere. The location of photon sphere around the HBHs doesn’t deviate too much from the Schwarzschild black hole, perhaps the increment of ϕ1\phi_{1} could exhibit large deviation. According to the total number of orbits, the null geodesics can be divided into three types which are the direct, lensing and photon sphere with their associated impact parameters bb. When the strength of scalar field at the horizon increases, the values of bb correspond to these three types of null geodesics can vary significantly from the Schwarzschild black hole.

Finally, several new research directions can be carried out from this work for the future research. First, we can continue to study the imaging of shadow with different emission profiles for this HBH, and compare them with the Schwarzschild black hole, since we have already calculated the null geodesics in this paper. Recently, the shadow of HBH with inverted Higgs-like scalar potential [8] has been studied [69]. Then, we can also study the imaging of shadow and then provide some constraints for some parameters for different HBHs [6, 9] with different scalar potentials in the EKG theory.

Acknowledgement

XYC is supported by the starting grant of Jiangsu University of Science and Technology (JUST). WF is supported in part by the National Natural Science Foundation of China under Grant No. 12105121. We acknowledge to have useful discussions with Carlos Benavides-Gallego, Jose Luis Blázquez-Salcedo, Jutta Kunz, Eduard Alexis Larranaga.

References

  • [1] W. Israel, Phys. Rev. 164, 1776-1779 (1967)
  • [2] B. Carter, Phys. Rev. Lett. 26, 331-333 (1971)
  • [3] R. Ruffini and J. A. Wheeler, Phys. Today 24, no.1, 30 (1971)
  • [4] A. Corichi, U. Nucamendi and M. Salgado, Phys. Rev. D 73, 084002 (2006)
  • [5] X. Y. Chew, D. h. Yeom and J. L. Blázquez-Salcedo, Phys. Rev. D 108, no.4, 044020 (2023)
  • [6] X. Y. Chew and D. h. Yeom, Phys. Rev. D 110 (2024) no.4, 044036
  • [7] S. S. Gubser, Class. Quant. Grav. 22 (2005), 5121-5144
  • [8] X. Y. Chew and K. G. Lim, Phys. Rev. D 109 (2024) no.6, 064039
  • [9] X. Y. Chew and Y. S. Myung, Phys. Rev. D 110 (2024) no.4, 044011
  • [10] C. A. R. Herdeiro and E. Radu, Int. J. Mod. Phys. D 24 (2015) no.09, 1542014
  • [11] S. R. Coleman and F. De Luccia, Phys. Rev. D 21, 3305 (1980).
  • [12] X. Y. Chew and K. G. Lim, Universe 10 (2024), 212
  • [13] O. Bechmann and O. Lechtenfeld, Class. Quant. Grav. 12 (1995), 1473-1482
  • [14] H. Dennhardt and O. Lechtenfeld, Int. J. Mod. Phys. A 13 (1998), 741-764
  • [15] K. A. Bronnikov and G. N. Shikin, Grav. Cosmol. 8 (2002), 107-116
  • [16] C. Martinez, R. Troncoso and J. Zanelli, Phys. Rev. D 70 (2004), 084035
  • [17] V. V. Nikonov, J. V. Tchemarina and A. N. Tsirulev, Class. Quant. Grav. 25 (2008), 138001
  • [18] A. Anabalon and J. Oliva, Phys. Rev. D 86 (2012), 107501
  • [19] O. S. Stashko and V. I. Zhdanov, Gen. Rel. Grav. 50 (2018), 105
  • [20] C. Gao and J. Qiu, Gen. Rel. Grav. 54 (2022) no.12, 158
  • [21] T. Karakasis, G. Koutsoumbas and E. Papantonopoulos, Phys. Rev. D 107 (2023) no.12, 124047
  • [22] A. N. Atmaja, Eur. Phys. J. C 84 (2024) no.5, 456
  • [23] X. Li and J. Ren, Phys. Rev. D 109 (2024) no.10, 104061
  • [24] X. P. Rao, H. Huang and J. Yang, [arXiv:2403.11770 [gr-qc]].
  • [25] M. P. Wijayanto, E. S. Fadhilla, F. T. Akbar and B. E. Gunara, Gen. Rel. Grav. 55 (2023) no.1, 19
  • [26] M. P. Wijayanto, F. T. Akbar and B. E. Gunara, Gen. Rel. Grav. 56 (2024) no.2, 22
  • [27] J. Kunz and Y. Shnir, Phys. Rev. D 107 (2023) no.10, 104062
  • [28] J. Kunz, V. Loiko and Y. Shnir, Phys. Rev. D 110 (2024) no.12, 125020
  • [29] Y. Brihaye, F. Buisseret, B. Hartmann and O. Layfield, [arXiv:2410.22231 [gr-qc]].
  • [30] C. Herdeiro, E. Radu and E. dos Santos Costa Filho, JCAP 07 (2024), 081
  • [31] R. A. Hulse and J. H. Taylor, Astrophys. J. Lett. 191 (1974), L59
  • [32] Y. Hagihara, Japan. J. Astron. Geophys. 8 (1931) 67.
  • [33] V. Witzany and G. A. Piovano, Phys. Rev. Lett. 132 (2024) no.17, 171401
  • [34] E. Hackmann and C. Lammerzahl, Phys. Rev. Lett. 100 (2008), 171101
  • [35] Y. K. Lim, Phys. Rev. D 91 (2015) no.2, 024048
  • [36] Y. K. Lim, Phys. Rev. D 101 (2020) no.10, 104031
  • [37] R. Capobianco, B. Hartmann and J. Kunz, Phys. Rev. D 109 (2024) no.6, 064042
  • [38] R. Capobianco, B. Hartmann and J. Kunz, Phys. Rev. D 110 (2024) no.8, 084078
  • [39] V. Diemer, J. Kunz, C. Lämmerzahl and S. Reimers, Phys. Rev. D 89 (2014) no.12, 124026
  • [40] A. García, E. Hackmann, J. Kunz, C. Lämmerzahl and A. Macías, J. Math. Phys. 56 (2015), 032501
  • [41] Y. Liu and X. Zhang, Phys. Rev. D 111 (2025) no.4, 044056
  • [42] S. Soroushfar, R. Saffari and A. Jafari, Phys. Rev. D 93 (2016) no.10, 104037
  • [43] S. Kazempour and S. Soroushfar, Chin. J. Phys. 65 (2020), 579-592
  • [44] B. Hoseini, R. Saffari and S. Soroushfar, Eur. Phys. J. Plus 136 (2021) no.5, 489
  • [45] J. C. Drawer and S. Grunau, Eur. Phys. J. C 80 (2020) no.6, 536
  • [46] X. Yang and J. Wang, Astrophys. J. Suppl. 207 (2013), 6
  • [47] S. W. Wei and Y. X. Liu, Phys. Rev. D 98 (2018) no.2, 024042
  • [48] Y. Liu and B. Sun, Chin. Phys. C 48 (2024) no.4, 045107
  • [49] A. Cieślik, E. Hackmann and P. Mach, Phys. Rev. D 108 (2023) no.2, 024056
  • [50] X. L. Yang and J. C. Wang, Astron. Astrophys. 561 (2014), A127
  • [51] Y. P. Zhang, S. W. Wei, W. D. Guo, T. T. Sui and Y. X. Liu, Phys. Rev. D 97 (2018) no.8, 084056
  • [52] C. Y. Wang, D. S. Lee and C. Y. Lin, Phys. Rev. D 106 (2022) no.8, 084048
  • [53] Y. T. Li, C. Y. Wang, D. S. Lee and C. Y. Lin, Phys. Rev. D 108 (2023) no.4, 044010
  • [54] Y. C. Ko, D. S. Lee and C. Y. Lin, Phys. Rev. D 109 (2024) no.8, 084043
  • [55] K. Flathmann and S. Grunau, Phys. Rev. D 94 (2016) no.12, 124013
  • [56] H. Cebeci, N. Özdemir and S. Şentorun, Phys. Rev. D 93 (2016) no.10, 104031
  • [57] K. Flathmann and S. Grunau, Phys. Rev. D 92 (2015) no.10, 104027
  • [58] K. Chen and S. W. Wei, Phys. Rev. D 110 (2024) no.2, 024041
  • [59] Y. Z. Li, X. M. Kuang and Y. Sang, Eur. Phys. J. C 84 (2024) no.5, 529
  • [60] B. Chen, H. Sun and J. Xu, Phys. Rev. D 110 (2024) no.12, 125016
  • [61] E. Battista and G. Esposito, Eur. Phys. J. C 82 (2022) no.12, 1088
  • [62] L. Del Grosso and P. Pani, Phys. Rev. D 108 (2023) no.6, 064042
  • [63] E. Berti, V. De Luca, L. Del Grosso and P. Pani, Phys. Rev. D 109 (2024) no.12, 124008
  • [64] C. W. Misner and D. H. Sharp, Phys. Rev. 136 (1964), B571-B576
  • [65] U. Ascher, J. Christiansen and R. D. Russell, Math. Comput. 33, no.146, 659-679 (1979)
  • [66] S. E. Gralla, D. E. Holz and R. M. Wald, Phys. Rev. D 100 (2019) no.2, 024018
  • [67] The value for the location of ISCO of Schwarzcshild black hole in the radial coordinate rr is given by rISCO=3rHr_{\text{ISCO}}=3r_{H}.
  • [68] The value for the location of photon sphere of Schwarzcshild black hole in the radial coordinate rr is given by rph=3rH/2r_{\text{ph}}=3r_{H}/2.
  • [69] K. G. Lim and X. Y. Chew, [arXiv:2501.07029 [gr-qc]].