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

Dynamics of null particles and shadow for general rotating black hole

Kun Meng
School of Physics and Electronic Information, Weifang University, Weifang 261061, China
Xi-Long Fan111xilong.fan@whu.edu.cn
School of Physics and Technology, Wuhan University, Wuhan, Hubei 430072, China
Song Li
Shanghai Astronomical Observatory, Shanghai, 200030, China;
School of Astronomy and Space Science, University of Chinese Academy of Sciences,
Beijing, 100049, China
Wen-Biao Han
School of Fundamental Physics and Mathematical Sciences,
Hangzhou Institute for Advanced Study, UCAS, Hangzhou 310024, China;
Shanghai Astronomical Observatory, Shanghai, 200030, China;
School of Astronomy and Space Science, University of Chinese Academy of Sciences,
Beijing, 100049, China;
International Centre for Theoretical Physics Asia-Pacific, Beijing/Hangzhou, China
Hongsheng Zhang222sps_zhanghs@ujn.edu.cn
School of Physics and Technology, University of Jinan, Jinan, China
Abstract

The Johannsen black hole (BH) is a generic rotating BH admitting three constants of motions ( energy, angular momentum, and Carter constant) and is characterized by four deviation parameters besides mass and spin, which could be a model-independent probe of the no-hair theorem. We systematically study the dynamics of null particles around Johannsen BH, revealing the effects of the deviation parameters on the BH shadow as well as the effects of spin. By using the shadow boundaries of M87* and SgrA*, for the first time, the deviation parameters of those BHs are constrained. The detailed results depend on the spin aa and inclination angle θ0\theta_{0}. Assuming a=0.2a=0.2 and θ0=15\theta_{0}=15^{\circ}, the deviation parameter α13\alpha_{13} are constained within \sim [-3.5, 6] for M87* observation and [-3, 0.5] for SgrA* observation. We also show the images of a Johannsen BH surrounded by a Page-Thorne thin accretion disk observed by a remote observer with a ray-tracing method and discuss the effects of the deviation parameters on deforming the accretion disk image, which could be tested by observations with higher sensitivities in the future.

1 Introduction

Black hole may be the most thoroughly-studied object before its discovery in the history of science. Recently, the Event Horizon Telescope (EHT) collaboration announced the images of the supermassive BHs M87* and SgrA*[1, 2], these discoveries together with the detections of gravitational waves (GWs)[3, 4] confirm the existence of black holes, and open a new era of testing gravity in strong-field regime and far beyond the scale of solar system.

Shadow is the image formed by photons which are sent off by the central BH and reach the observer finally. All the unstable orbits of photons form light rings around BH, the projection of the light rings on the observing screen is what called shadow. The unstable orbits are the critical ones, on which small perturbations would cause the photons to be either captured by the central compact object or sent off to infinity. The light rays forming shadow pass very close to the event horizon and are bent a lot due to the strong gravitational lensing effect, therefore observations of shadow make it possible to test the strong-field properties of gravity.

As we know, according to no-hair theorem[5, 6, 7], Kerr metric is the unique neutral, stationary, and asymptotically flat BH solution of vacuum Einstein equations in four dimensions. While, it is possible BHs are not isolated and are surrounded and influenced by other fields, as well as, the underlying gravity theory is possible to be some modified gravity theory, which means Kerr hypothesis may be be deviated. Recently, non-Kerr BHs have been studied extensively by considering coupling matter fields to gravity[8, 9, 10, 11], or modifying theories of gravity [12, 13, 14, 15, 16, 17], or immersing the BHs in astrophysical environment [18, 19, 20, 21, 22, 23]. Testing no-hair theorem, and thus testing general relativity, is one of the most important topics in black hole physics and astrophysics.

Observationally, it is quite sensible to treat the deviations of Kerr BH in a model-independent way, so that the effects of the deviations can be studies universally. The model-independent strong-field tests require a modified spacetime which deviates from the Kerr metric in a parametrized form. In Ref.[24], Johannsen constructed a general asymptotically flat rotating BH that depends nonlinearly on the deviation functions, and the spacetime symmetries are held meanwhile. The Johannsen BH admits three constants of motion, i.e., the energy, angular momentum and Carter constant, which is regular outside the event horizon and free of closed timelike curves. The BH possesses correct Newtonian limit in the non-relativistic regime, and is consistent with all current weak-field tests. Up to leading truncation order, the BH possesses four deviation parameters that measure potential deviations from the Kerr metric in the strong-field regime. Apparently, the BH is not solution of any particular theories of gravity, but can be embedded into several known modified theories of gravity invoked in astrophysics and cosmology by suitably choosing the deviation parameters. Significantly Johannsen BH does not suffer from pathologies in the exterior domain, it hence fairly serves as a phenomenological framework for strong-field tests of the no-hair theorem in general classes of gravity theories.

The shadow of Schwarzschild BH was first studied by Synge[25], and Bardeen first studied the shadow of Kerr[26], the result showed that, compared to the spherical BH, the shadow of Kerr is no longer circular, the distortion of shadow is understood as the effect of frame dragging. Aside from the distortion effects of rotating parameter, studies show that the deviation parameters lead to deformation of the shadows significantly as well, thus shadow can be used to extract information of BHs so as to detect the deviations from Kerr and constrain the BH parameters in the coming observations[27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55]. The announcement of BH shadow inspire many attempts to test no-hair theorem, alternative theories of gravity, the candidates of dark matter, and quantum effects of gravity utilizing the data of shadow [56, 57, 58, 59, 60, 61, 62, 63, 64, 65, 66, 67, 68, 69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99, 100]. With the improvements of detection accuracy, BH shadow is expected to serve as independent measurement or complementary to GW detections to provide a powerful way of constraining the BH parameters and distinguishing different BHs.

Aside from the deviation effects, the luminous source is another factor that affect the shadow. We know BHs accrete matters and form accretion disk surrounding them, the matters in accretion disk would emit photons and serve as luminous source of the shadow, which is expected to be affected significantly by accretion disk. Compared to the ideal treatment of BH shadow, taking into account accretion disk is more realistic in astrophysics.

In this paper, we study the shadow of Johannsen BH, our results show that the shadow are affected significantly by rotating and deviation parameters, including the shape of shadow boundaries, and the luminosity, the central dark region and the region with high luminosity of the accretion disk images. With the observational results of M87* and SgrA* we constrain the deviation parameters further based on the theoretical constraints. Due to the universal features of Johannsen BH, this study is helpful to test no-hair theorem and modified gravities in a model-independent way.

The paper is organised as, in section 2 we review the general rotating BH proposed by Johannsen. In section 3, we plot the shadow boundaries on celestial plane, calculate the shadow radius, and constrain the deviation parameters with observational data. In section 4, we give the accretion disk images of Johannsen BHs by ray-tracing method. We summarize our results in the last section 5. In this work, the geometrized units are used with G=c=1G=c=1.

2 General rotating BH

In this section we review Johannsen BH proposed in Ref.[24]. Johannsen BH is a general neutral, stationary and asymptoticall flat BH, which is a generalization of Kerr. Let’s start with Kerr, in Boyer-Lindquist coordinates Kerr BH is given by

ds2\displaystyle ds^{2} =(12MrΣ)dt22Marsin2θΣdtdϕ+ΣΔdr2\displaystyle=-\left(1-\frac{2Mr}{\Sigma}\right)dt^{2}-\frac{2Mar\sin^{2}\theta}{\Sigma}dtd\phi+\frac{\Sigma}{\Delta}dr^{2}
+Σdθ2+(r2+a2+2Ma2rsin2θΣ)sin2θdϕ2,\displaystyle+\Sigma d\theta^{2}+\left(r^{2}+a^{2}+\frac{2Ma^{2}r\sin^{2}\theta}{\Sigma}\right)\sin^{2}\theta d\phi^{2}, (1)

with

Δr22Mr+a2,Σr2+a2cos2θ.\displaystyle\Delta\equiv r^{2}-2Mr+a^{2},\quad\quad\quad\Sigma\equiv r^{2}+a^{2}\cos^{2}\theta. (2)

We know general stationary and axisymmetric BHs are of Petrov type I which admit two constants of motion, the energy EE and angular momentum LzL_{z}. While Kerr BH is of Petrov type D, since Carter found there exist a third constant of motion, the Carter constant QQ[101], by separating variables of the Hamilton-Jacobi equations of test particles

Sτ=12gαβSxαSxβ.\displaystyle-\frac{\partial S}{\partial\tau}=\frac{1}{2}g^{\alpha\beta}\frac{\partial S}{\partial x^{\alpha}}\frac{\partial S}{\partial x^{\beta}}. (3)

The successful separation of variables enable the geodesic motions of test particles in Kerr spacetime to be integrable. If the Kerr metric is required to be modified in the manner by holding the Carter symmetry, i.e., the motions of test particle in this novel spacetime is expected to be integrable as well, a metric constructed by Johannsen satisfies this requirement, the metric can be expressed in contravariant form as

gαβxαxβ\displaystyle g^{\alpha\beta}\frac{\partial}{\partial x^{\alpha}}\frac{\partial}{\partial x^{\beta}} (4)
=\displaystyle= 1ΔΣ~[(r2+a2)A1(r)t+aA2(r)ϕ]2\displaystyle-\frac{1}{\Delta\tilde{\Sigma}}\left[(r^{2}+a^{2})A_{1}(r)\frac{\partial}{\partial t}+aA_{2}(r)\frac{\partial}{\partial\phi}\right]^{2}
+ΔΣ~A5(r)(r)2+1Σ~A6(θ)(θ)2\displaystyle+\frac{\Delta}{\tilde{\Sigma}}A_{5}(r)\left(\frac{\partial}{\partial r}\right)^{2}+\frac{1}{\tilde{\Sigma}}A_{6}(\theta)\left(\frac{\partial}{\partial\theta}\right)^{2}
+1Σ~sin2θ[A3(θ)ϕ+asin2θA4(θ)t]2,\displaystyle+\frac{1}{\tilde{\Sigma}\sin^{2}\theta}\left[A_{3}(\theta)\frac{\partial}{\partial\phi}+a\sin^{2}\theta A_{4}(\theta)\frac{\partial}{\partial t}\right]^{2},

where

Σ~Σ+f(r)+g(θ).\displaystyle\tilde{\Sigma}\equiv\Sigma+f(r)+g(\theta). (5)

The explicit form of the deviation functions Ai(r)A_{i}(r) and Aj(θ)A_{j}(\theta) will be discussed below.

Refer to caption
Figure 1: Shadow boundaries of Johannsen BH with different values of α13\alpha_{13}, aa and θ0\theta_{0}. All plots in the same row correspond to the same inclination angle. All plots in the same column correspond to the same rotating parameter. The red circle on every plot corresponds to α13=2\alpha_{13}=-2, the yellow circle corresponds to α13=0\alpha_{13}=0, the green circle corresponds to α13=10\alpha_{13}=10, the blue circle corresponds to α13=20\alpha_{13}=20, and the purple circle corresponds to α13=30\alpha_{13}=30. Masses of the black holes are set to be unity M=1M=1.

.

To see the motions of test particle in the spacetime described by the metric (4) are integrable, we suppose the Hamilton-Jacobi function of a particle with mass m0m_{0} to be of the form

S=12m02τEt+Lzϕ+Sr(r)+Sθ(θ),\displaystyle S=\frac{1}{2}m_{0}^{2}\tau-Et+L_{z}\phi+S_{r}(r)+S_{\theta}(\theta), (6)

which indicates pr=dSr/drp_{r}=\textmd{d}S_{r}/\textmd{d}r and pθ=dSθ/dθp_{\theta}=\textmd{d}S_{\theta}/\textmd{d}\theta. Inserting the metric (4) into the Hamilton-Jacobi equation (3), we have

m02=\displaystyle-m_{0}^{2}= 1ΔΣ~[(r2+a2)A1(r)E+aA2(r)Lz]2\displaystyle-\frac{1}{\Delta\tilde{\Sigma}}\left[-(r^{2}+a^{2})A_{1}(r)E+aA_{2}(r)L_{z}\right]^{2} (7)
+1Σ~sin2θ[LzaEsin2θ]2\displaystyle+\frac{1}{\tilde{\Sigma}\sin^{2}\theta}\left[L_{z}-aE\sin^{2}\theta\right]^{2}
+ΔΣ~A5(r)(Srr)2+1Σ~(Sθθ)2.\displaystyle+\frac{\Delta}{\tilde{\Sigma}}A_{5}(r)\left(\frac{\partial S_{r}}{\partial r}\right)^{2}+\frac{1}{\tilde{\Sigma}}\left(\frac{\partial S_{\theta}}{\partial\theta}\right)^{2}.

After separating variables of Eq. (7) we have

C=\displaystyle C= 1Δ[(r2+a2)A1(r)E+aA2(r)Lz]2\displaystyle\frac{1}{\Delta}\left[-(r^{2}+a^{2})A_{1}(r)E+aA_{2}(r)L_{z}\right]^{2}
m02[r2+f(r)]ΔA5(r)(Srr)2,\displaystyle-m_{0}^{2}\left[r^{2}+f(r)\right]-\Delta A_{5}(r)\left(\frac{\partial S_{r}}{\partial r}\right)^{2}, (8)
C=\displaystyle C= 1sin2θ[LzaEsin2θ]2+m02(a2cos2θ+g(θ))+(Sθθ)2,\displaystyle\frac{1}{\sin^{2}\theta}\left[L_{z}-aE\sin^{2}\theta\right]^{2}+m_{0}^{2}(a^{2}\cos^{2}\theta+g(\theta))+\left(\frac{\partial S_{\theta}}{\partial\theta}\right)^{2}, (9)

where CC is the separation constant, through which we could define the Carter constant

Q=C(LzaE)2.\displaystyle Q=C-(L_{z}-aE)^{2}. (10)

Since Carter symmetry is a critical property of Johannsen metric, it is quite sensible to discuss the underlying physics of Carter symmetry/Carter constant. Carter constant reveals a deep and complicated symmetry in spacetime, it corresponds to a Killing tensor. For Johannsen BH, we find the Killing tensor is given by

Kμν=a2cos2θgμν+1sin2θδϕμδϕν+a2sin2θδtμδtν2aδtμδϕν+δθμδθν.K^{\mu\nu}=a^{2}\cos^{2}\theta g^{\mu\nu}+\frac{1}{\sin^{2}\theta}\delta^{\mu}_{\phi}\delta^{\nu}_{\phi}+a^{2}\sin^{2}\theta\delta^{\mu}_{t}\delta^{\nu}_{t}-2a\delta^{\mu}_{t}\delta^{\nu}_{\phi}+\delta^{\mu}_{\theta}\delta^{\nu}_{\theta}. (11)

One checks that the Killing equation K(μν;λ)=0K_{(\mu\nu;\lambda)}=0 is satisfied, where the bracket denotes symmetrization of the indices. As Killing vectors t\partial_{t} and ϕ\partial_{\phi} correspond to energy and angular momentum of the spacetime, Killing tensor describes a nontrivial symmetry of the spacetime and makes the general geodesic motion of probe particle to be integrable, for more motivations and applications of Carter constant see Cater’s original paper and related works [102, 103, 104, 105, 106].

Generally it relates to energy and angular momentum at zz-direction. Here we concentrate on the physical implications of Cater constant in this article. The separation constant CC and Carter constant QQ in our paper are related through Eq.(10). We take new separation constant following [105]

Λ=C+2aELzm02a2,\Lambda=C+2aEL_{z}-m_{0}^{2}a^{2}, (12)

then the separation equation (9) can be rewritten as

pθ2+pϕ2sin2θ+m02a2[(Em0)21]sin2θ=Λ,p_{\theta}^{2}+\frac{p_{\phi}^{2}}{\sin^{2}\theta}+m_{0}^{2}a^{2}\left[\left(\frac{E}{m_{0}}\right)^{2}-1\right]\sin^{2}\theta=\Lambda, (13)

where pϕ=Lzp_{\phi}=L_{z}. Since we will see later g(θ)g(\theta) is set to be zero by observation requirement, here we ignore this term to clarify the physical meaning of Carter constant. Note that eq.(13) does not depend on the mass MM of gravitational source. Now we try to explain the terms on the left side of above equation. In the a0a\rightarrow 0 limit, i.e., the spherical symmetry case, the physical meaning of the first two terms are clear. They are the square of angular momentum L2=pθ2+pϕ2sin2θL^{2}=p_{\theta}^{2}+\frac{p_{\phi}^{2}}{\sin^{2}\theta}, in which the separation constant Λ\Lambda equals to L2L^{2}. When a0a\neq 0, we consider the weak-field approximation, which implies that the observer locates at far field region, the spatial metric can be rewritten as

dσ2=Σr2+a2dr2+Σdθ2+(r2+a2)sin2θdϕ2,d\sigma^{2}=\frac{\Sigma}{r^{2}+a^{2}}dr^{2}+\Sigma d\theta^{2}+(r^{2}+a^{2})\sin^{2}\theta d\phi^{2}, (14)

with oblate spheroidal coordinates

x\displaystyle x =\displaystyle= (r2+a2)1/2sinθcosϕ,\displaystyle(r^{2}+a^{2})^{1/2}\sin\theta\cos\phi, (15)
y\displaystyle y =\displaystyle= (r2+a2)1/2sinθsinϕ,\displaystyle(r^{2}+a^{2})^{1/2}\sin\theta\sin\phi, (16)
z\displaystyle z =\displaystyle= rcosθ.\displaystyle r\cos\theta. (17)

With the asymptotic metric (14) the square modulus of particle’s linear momentum observed at infinity is given by

p2=grrpr2+gθθpθ2+gϕϕpϕ2.p^{2}=g^{rr}_{\infty}p_{r}^{2}+g^{\theta\theta}_{\infty}p_{\theta}^{2}+g^{\phi\phi}_{\infty}p_{\phi}^{2}. (18)

In the limit rr\rightarrow\infty, we have

gθθpθ20,gϕϕpϕ20.g^{\theta\theta}_{\infty}p_{\theta}^{2}\longrightarrow 0,\quad\quad\quad\quad g^{\phi\phi}_{\infty}p_{\phi}^{2}\longrightarrow 0. (19)

Thus

grrpr2p2=m02[(Em0)21].g^{rr}_{\infty}p_{r}^{2}\longrightarrow p^{2}=m_{0}^{2}\left[\left(\frac{E}{m_{0}}\right)^{2}-1\right]. (20)

Now the separation equation (13) can be rewritten as

pθ2+pϕ2sin2θ+|pr|2a2sin2θ=Λ.p_{\theta}^{2}+\frac{p_{\phi}^{2}}{\sin^{2}\theta}+|p_{r}|^{2}a^{2}\sin^{2}\theta=\Lambda. (21)

This equation implies that, if a particle at infinity possess nonvanishing momentum prp_{r} it also possesses nonvanishing angular momentum with respect to the center OO of r=0r=0 disk. In the Boyer-Lindquist coordinates, the θ=constant\theta=constant surfaces are not cones with vertices at the origin, but are hyperboloids of rotation, which cross the r=0r=0 disk into a circle with radius asinθa\sin\theta. At infinity the vector pr\vec{p}_{r} coincide with the asymptotes of the hyperboloid, which is to say pr\vec{p}_{r} is tangent to θ=constant\theta=constant surfaces, thus parallel transport along θ=constant\theta=constant surface maintains pr\vec{p}_{r} to be tangent to the hyperboloid. When crossing the r=0r=0 disk, the point of maximum approach to the center OO is reached, now pr\vec{p}_{r} is orthogonal to the r=0r=0 disk. So its momentum with respect to the center OO has a square modulus given by |pr|2a2sin2θ|p_{r}|^{2}a^{2}\sin^{2}\theta, which is just the term appearing in eq.(21). It should be mentioned, for bounded orbits of particles the sign before the third term on the left side of eq.(21) should be changed, for more details please refer to Ref.[105].

From eq.(21) we see that, the separation constant has the meaning of ‘extended’ angular momentum which come not only from the angular motions but also from the radial motion. The radial angular momentum is not obvious, since for the spacetime of black hole with spherical symmetry, angular momentum comes only from angular motions. While, for the spacetime of rotating black hole, the radial angular momentum appears. It arises from the particular choice of coordinates which in turn ensures the separability of the Hamilton-Jacobi equation.

Solving the variables-separated Eqs. (8) and (9) we have

Sr(r)=±𝑑r1ΔR(r)A5(r),Sθ(θ)=±𝑑θΘ(θ),\displaystyle S_{r}(r)=\pm\int dr\frac{1}{\Delta}\sqrt{\frac{R(r)}{A_{5}(r)}},\;S_{\theta}(\theta)=\pm\int d\theta\sqrt{\Theta(\theta)}, (22)

with

R(r)\displaystyle R(r) P2Δ{m02[r2+f(r)]+(LzaE)2+Q},\displaystyle\equiv P^{2}-\Delta\left\{m_{0}^{2}\left[r^{2}+f(r)\right]+(L_{z}-aE)^{2}+Q\right\}, (23)
Θ(θ)\displaystyle\Theta(\theta) Q+(LzaE)2m02(a2cos2θ+g(θ))1sin2θ[LzaEsin2θ]2,\displaystyle\equiv Q+(L_{z}-aE)^{2}-m_{0}^{2}(a^{2}\cos^{2}\theta+g(\theta))-\frac{1}{\sin^{2}\theta}\left[L_{z}-aE\sin^{2}\theta\right]^{2}, (24)
P\displaystyle P (r2+a2)A1(r)EaA2(r)Lz.\displaystyle\equiv(r^{2}+a^{2})A_{1}(r)E-aA_{2}(r)L_{z}. (25)

The explicit integration expression of coordinates and proper time can be obtained by setting the partial derivative of Hamilton-Jacobi function with respect to the constants of motion to be zero, that’s to say the motions of particles in spacetime (4) are integrable.

Refer to caption
Figure 2: Shadow boundaries of Johannsen BH with different values of α22\alpha_{22}, aa and θ0\theta_{0}. All plots in the same row correspond to the same inclination angle. All plots in the same column correspond to the same rotating parameter. The red circle on every plot corresponds to α22=2\alpha_{22}=-2, the green circle corresponds to α22=0\alpha_{22}=0, the yellow circle corresponds to α22=10\alpha_{22}=10, the blue circle corresponds to α22=20\alpha_{22}=20, the purple circle corresponds to α22=30\alpha_{22}=30. Masses of the black holes are set to be unity M=1M=1.

In order to write the metric in explicit form, the derivation functions Ai(r)A_{i}(r), i=1,2,5i=1,2,5, can be expanded as a power series of M/rM/r[24]:

Ai(r)n=0αin(Mr)n,i=1,2,5,\displaystyle A_{i}(r)\equiv\sum_{n=0}^{\infty}\alpha_{in}\left(\frac{M}{r}\right)^{n},~{}~{}~{}~{}i=1,2,5, (26)

as well as

f(r)n=0ϵnMnrn2,g(θ)M2k,l=0γklsinkθcoslθ,\displaystyle f(r)\equiv\sum_{n=0}\epsilon_{n}\frac{M^{n}}{r^{n-2}},\;\;g(\theta)\equiv M^{2}\sum_{k,l=0}^{\infty}\gamma_{kl}\sin^{k}\theta\cos^{l}\theta, (27)

Expanding the metric in power of 1/r1/r and requiring the metric to be asymptotically flat, the undetermined parameters and functions are fixed to be α10=α20=α50=1\alpha_{10}=\alpha_{20}=\alpha_{50}=1, ϵ0=ϵ1=0\epsilon_{0}=\epsilon_{1}=0, and A3(θ)=A4(θ)=A6(θ)=1A_{3}(\theta)=A_{4}(\theta)=A_{6}(\theta)=1. If the parameter MM is required to be the mass of the central object, the parameters can be further constrained to be α11=α21=α51=0\alpha_{11}=\alpha_{21}=\alpha_{51}=0. The metric should be consistent with the the weak-field tests, in the parameterized post-Newtonian (PPN) formulism, the general metric can be written in the form [107, 108]

ds2=[12Mr+2(βPPNγPPN)M2r2]dt2+(1+2γPPNMr)dr2+r2(dθ2+sin2θdϕ2).ds^{2}=-\left[1-\frac{2M}{r}+2(\beta_{PPN}-\gamma_{PPN})\frac{M^{2}}{r^{2}}\right]dt^{2}+(1+2\gamma_{PPN}\frac{M}{r})dr^{2}+r^{2}(d\theta^{2}+\sin^{2}\theta d\phi^{2}). (28)

For general relativity, βPPN=γPPN=1\beta_{PPN}=\gamma_{PPN}=1. Values of βPPN,γPPN\beta_{PPN},\gamma_{PPN} corresponding to some other metric theories can be found in Chapter 5 of Ref.[108]. We perform the large rr expansion of Johannsen metric and find

ds2=[12MrM2(2α12ϵ2)g(θ)r2]dt2+(1+2Mr)dr2+r2(dθ2+sin2θdϕ2).ds^{2}=-\left[1-\frac{2M}{r}-\frac{M^{2}(2\alpha_{12}-\epsilon_{2})-g(\theta)}{r^{2}}\right]dt^{2}+(1+\frac{2M}{r})dr^{2}+r^{2}(d\theta^{2}+\sin^{2}\theta d\phi^{2}). (29)

Note that the spin parameter aa appear in the higher order expansion of 1r\frac{1}{r}. Comparing metric (29) with metric (28) we have

2(βPPNγPPN)\displaystyle 2(\beta_{PPN}-\gamma_{PPN}) =\displaystyle= ϵ22α12+g(θ)M2,\displaystyle\epsilon_{2}-2\alpha_{12}+\frac{g(\theta)}{M^{2}}, (30)
γPPN\displaystyle\gamma_{PPN} =\displaystyle= 1,\displaystyle 1, (31)

which implies

βPPN1=12[ϵ22α12+g(θ)M2]\beta_{PPN}-1=\frac{1}{2}\left[\epsilon_{2}-2\alpha_{12}+\frac{g(\theta)}{M^{2}}\right] (32)

This quantity is constrained by observations to be [108]

βPPN1=(0.2±2.5)×105.\beta_{PPN}-1=(0.2\pm 2.5)\times 10^{-5}. (33)

In order to avoid any fine-tuning between the parameters ϵ2\epsilon_{2}, α12\alpha_{12} and the function g(θ)g(\theta), one can set ϵ2=α12=g(θ)=0\epsilon_{2}=\alpha_{12}=g(\theta)=0 for simplicity.

Now the metric can be expressed in covariant form as

gtt\displaystyle g_{tt} =Σ~[Δa2A2(r)2sin2θ][(r2+a2)A1(r)a2A2(r)sin2θ]2,\displaystyle=-\frac{\tilde{\Sigma}[\Delta-a^{2}A_{2}(r)^{2}\sin^{2}\theta]}{[(r^{2}+a^{2})A_{1}(r)-a^{2}A_{2}(r)\sin^{2}\theta]^{2}},
gtϕ\displaystyle g_{t\phi} =a[(r2+a2)A1(r)A2(r)Δ]Σ~sin2θ[(r2+a2)A1(r)a2A2(r)sin2θ]2,\displaystyle=-\frac{a[(r^{2}+a^{2})A_{1}(r)A_{2}(r)-\Delta]\tilde{\Sigma}\sin^{2}\theta}{[(r^{2}+a^{2})A_{1}(r)-a^{2}A_{2}(r)\sin^{2}\theta]^{2}},
grr\displaystyle g_{rr} =Σ~ΔA5(r),\displaystyle=\frac{\tilde{\Sigma}}{\Delta A_{5}(r)},
gθθ\displaystyle g_{\theta\theta} =Σ~,\displaystyle=\tilde{\Sigma},
gϕϕ\displaystyle g_{\phi\phi} =Σ~sin2θ[(r2+a2)2A1(r)2a2Δsin2θ][(r2+a2)A1(r)a2A2(r)sin2θ]2.\displaystyle=\frac{\tilde{\Sigma}\sin^{2}\theta\left[(r^{2}+a^{2})^{2}A_{1}(r)^{2}-a^{2}\Delta\sin^{2}\theta\right]}{[(r^{2}+a^{2})A_{1}(r)-a^{2}A_{2}(r)\sin^{2}\theta]^{2}}.

with

A1(r)\displaystyle A_{1}(r) =1+n=3α1n(Mr)n,\displaystyle=1+\sum_{n=3}^{\infty}\alpha_{1n}\left(\frac{M}{r}\right)^{n}, (34)
A2(r)\displaystyle A_{2}(r) =1+n=2α2n(Mr)n,\displaystyle=1+\sum_{n=2}^{\infty}\alpha_{2n}\left(\frac{M}{r}\right)^{n},
A5(r)\displaystyle A_{5}(r) =1+n=2α5n(Mr)n,\displaystyle=1+\sum_{n=2}^{\infty}\alpha_{5n}\left(\frac{M}{r}\right)^{n},
Σ~\displaystyle\tilde{\Sigma} =r2+a2cos2θ+f(r),\displaystyle=r^{2}+a^{2}\cos^{2}\theta+f(r),
f(r)\displaystyle f(r) =n=3ϵnMnrn2.\displaystyle=\sum_{n=3}^{\infty}\epsilon_{n}\frac{M^{n}}{r^{n-2}}.

Thus to leading order, the metric is characterized by mass MM, spin aa and four deviation parameters α13\alpha_{13}, α22\alpha_{22}, α52\alpha_{52}, and ϵ3\epsilon_{3}. By requiring the metric to be regular outside the event horizon, the determinant of the metric to be non-negative, and the absence of closed timelike curves, the four deviation parameters can be further constrained to be

α52\displaystyle\alpha_{52} >(M+M2a2)2M2,\displaystyle>-\frac{\left(M+\sqrt{M^{2}-a^{2}}\right)^{2}}{M^{2}}, (35)
ϵ3\displaystyle\epsilon_{3} >(M+M2a2)3M3,\displaystyle>-\frac{\left(M+\sqrt{M^{2}-a^{2}}\right)^{3}}{M^{3}},
α13\displaystyle\alpha_{13} >(M+M2a2)3M3,\displaystyle>-\frac{\left(M+\sqrt{M^{2}-a^{2}}\right)^{3}}{M^{3}},
α22\displaystyle\alpha_{22} >(M+M2a2)2M2.\displaystyle>-\frac{\left(M+\sqrt{M^{2}-a^{2}}\right)^{2}}{M^{2}}.

For more details on constraining the deviation parameters from theoretical aspect please refer to Ref.[24]. It should be mentioned that, generally the Johannsen metric (2) is not a solution of any particular gravity theory, but can be mapped to known four-dimensional BH solution of theories of modified gravity by suitably choosing the deviation parameters, e.g., the metric can be mapped to modified gravity bumpy Kerr metric, rotating braneworld BHs, slowly rotating BHs in dynamical Chern-Simons gravity, and static BHs in Einstein-Dilaton-Gauss-Bonnet gravity[106, 110, 111, 112].

Refer to caption
Refer to caption
Figure 3: Shadow radius as function of α13\alpha_{13} (left plot) or α22\alpha_{22} (right plot). The dashed and solid lines correspond to θ0=15\theta_{0}=15^{\circ} and 9090^{\circ} respectively. The red, green and blue lines correspond to a=0.2,0.5a=0.2,0.5 and 0.9 respectively. The blue region on each plot represents the observation bound of M87*. The orange vertical lines represent the case a=0.2a=0.2 and θ0=15\theta_{0}=15^{\circ}. Masses of the black holes are set to be unity M=1M=1.

3 Shadow boundary

In this section, we study the effects of the BI parameters on the shadow boundary. Shadow boundary is expected to be formed by all the unstable orbits of photons known as light rings. The unstable orbits are the critical ones which separate the photons to escape to infinity or to be captured by the central BH.

3.1 Shadow boundary on celestial plane

With the results given in the last section, one obtains the geodesic equations of photon in the spacetime of Johannsen BH

Σ~dtdτ\displaystyle\tilde{\Sigma}\frac{dt}{d\tau} =a[aEsin2θLz]+(r2+a2)A1(r)ΔP,\displaystyle=-a\left[aE\sin^{2}\theta-L_{z}\right]+\frac{(r^{2}+a^{2})A_{1}(r)}{\Delta}P, (36)
Σ~drdτ\displaystyle\tilde{\Sigma}\frac{dr}{d\tau} =±A5(r)R(r),\displaystyle=\pm\sqrt{A_{5}(r)R(r)},
Σ~dθdτ\displaystyle\tilde{\Sigma}\frac{d\theta}{d\tau} =±Θ(θ),\displaystyle=\pm\sqrt{\Theta(\theta)},
Σ~dϕdτ\displaystyle\tilde{\Sigma}\frac{d\phi}{d\tau} =[aELzsin2θ]+aA2(r)ΔP.\displaystyle=-\left[aE-\frac{L_{z}}{\sin^{2}\theta}\right]+\frac{aA_{2}(r)}{\Delta}P.

The unstable circular orbit is determined by the equations

R=0,dRdr=0.\displaystyle R=0,\quad\quad\quad\quad\quad\quad\frac{\textmd{d}R}{\textmd{d}r}=0. (37)

Solving the above equation we have

Lz\displaystyle L_{z} =\displaystyle= E[3a4M3α13a2r(Mr3+r4+7M4α135M3rα13)r3(3Mr3+r4+3M4α13\displaystyle E\left[3a^{4}M^{3}\alpha_{13}-a^{2}r(Mr^{3}+r^{4}+7M^{4}\alpha_{13}-5M^{3}r\alpha_{13})-r^{3}(-3Mr^{3}+r^{4}+3M^{4}\alpha_{13}\right. (38)
2M3rα13)]/[ar(Mr3+r45M3rα22+M2(2a2+3r2)α22)],\displaystyle\quad\left.-2M^{3}r\alpha_{13})\right]/\left[ar(-Mr^{3}+r^{4}-5M^{3}r\alpha_{22}+M^{2}(2a^{2}+3r^{2})\alpha_{22})\right],
Q\displaystyle Q =\displaystyle= (E2(r10(3Mr3+r4+3M4α132M3rα13)2+a8M7α13α22(6Mr2α13\displaystyle(E^{2}(-r^{10}(-3Mr^{3}+r^{4}+3M^{4}\alpha_{13}-2M^{3}r\alpha_{13})^{2}+a^{8}M^{7}\alpha_{13}\alpha_{22}(6Mr^{2}\alpha_{13} (39)
4r3α22+M3α13α22)a6M4r(12M5r2α132α22+6Mr6(α132α22)α22\displaystyle-4r^{3}\alpha_{22}+M^{3}\alpha_{13}\alpha_{22})-a^{6}M^{4}r(12M^{5}r^{2}\alpha_{13}^{2}\alpha_{22}+6Mr^{6}(\alpha_{13}-2\alpha_{22})\alpha_{22}
8r7α222+2M7α132α222+M6rα132α2222M4r3α13α22(α13+4α22)+M2r5α13\displaystyle-8r^{7}\alpha_{22}^{2}+2M^{7}\alpha_{13}^{2}\alpha_{22}^{2}+M^{6}r\alpha_{13}^{2}\alpha_{22}^{2}-2M^{4}r^{3}\alpha_{13}\alpha_{22}(\alpha_{13}+4\alpha_{22})+M^{2}r^{5}\alpha_{13}
(15α13+34α22)+8M3r4α13(3α13α222))+a2Mr5(4r102M2r8(7α132α22)\displaystyle(15\alpha_{13}+34\alpha_{22})+8M^{3}r^{4}\alpha_{13}(-3\alpha_{13}-\alpha_{22}^{2}))+a^{2}Mr^{5}(4r^{10}-2M^{2}r^{8}(7\alpha_{13}-2\alpha_{22})
+10Mr9α222M10α132α222+M9rα132α222M5r5α13(19α13+30α22)2M3r7\displaystyle+10Mr^{9}\alpha_{22}-2M^{10}\alpha_{13}^{2}\alpha_{22}^{2}+M^{9}r\alpha_{13}^{2}\alpha_{22}^{2}-M^{5}r^{5}\alpha_{13}(19\alpha_{13}+30\alpha_{22})-2M^{3}r^{7}
(24α13+(158α22)α22)2M7r3α13(8α222+α13(21+α22))+2M6r4α13(28α13\displaystyle(24\alpha_{13}+(15-8\alpha_{22})\alpha_{22})-2M^{7}r^{3}\alpha_{13}(8\alpha_{22}^{2}+\alpha_{13}(21+\alpha_{22}))+2M^{6}r^{4}\alpha_{13}(28\alpha_{13}
+α22(15+4α22))+2M4r6(16α222+α13(21+4α22)))+a4M2r3(8r9α22\displaystyle+\alpha_{22}(15+4\alpha_{22}))+2M^{4}r^{6}(-16\alpha_{22}^{2}+\alpha_{13}(21+4\alpha_{22})))+a^{4}M^{2}r^{3}(8r^{9}\alpha_{22}
+8M7r2α132α22+2M3r6(α13α22)α22+4M9α132α222M8rα132α222+2Mr8\displaystyle+8M^{7}r^{2}\alpha_{13}^{2}\alpha_{22}+2M^{3}r^{6}(\alpha_{13}-\alpha_{22})\alpha_{22}+4M^{9}\alpha_{13}^{2}\alpha_{22}^{2}-M^{8}r\alpha_{13}^{2}\alpha_{22}^{2}+2Mr^{8}
(5α13+8α22)5M4r5(6α132+8α13α22+5α222)+M2r7(22α13+23α222)\displaystyle(5\alpha_{13}+8\alpha_{22})-5M^{4}r^{5}(6\alpha_{13}^{2}+8\alpha_{13}\alpha_{22}+5\alpha_{22}^{2})+M^{2}r^{7}(-22\alpha_{13}+23\alpha_{22}^{2})
+2M5r4α13(38α13+α22(35+2α22))+M6r3α13(8α222α13(49+6α22)))))/\displaystyle+2M^{5}r^{4}\alpha_{13}(38\alpha_{13}+\alpha_{22}(35+2\alpha_{22}))+M^{6}r^{3}\alpha_{13}(8\alpha_{22}^{2}-\alpha_{13}(49+6\alpha_{22})))))/
(a2r6(Mr3+r45M3rα22+M2(2a2+3r2)α22)2).\displaystyle(a^{2}r^{6}(-Mr^{3}+r^{4}-5M^{3}r\alpha_{22}+M^{2}(2a^{2}+3r^{2})\alpha_{22})^{2}).

We introduce celestial coordinates to visualize the shadow of BH[56, 113]

α\displaystyle\alpha =limr0(r02sinθ0dϕdr)=ξcscθ0,\displaystyle=\lim_{r_{0}\rightarrow\infty}\left(-r_{0}^{2}\sin\theta_{0}\frac{d\phi}{dr}\right)=-\xi\csc\theta_{0}, (40)
β\displaystyle\beta =limr0(r02dθdr)=±(η+a2cos2θ0ξ2cot2θ0)1/2,\displaystyle=\lim_{r_{0}\rightarrow\infty}\left(r_{0}^{2}\frac{d\theta}{dr}\right)=\pm\left(\eta+a^{2}\cos^{2}\theta_{0}-\xi^{2}\cot^{2}\theta_{0}\right)^{1/2},

where r0r_{0} is the distance between the BH and the observer, θ0\theta_{0} is the inclination angle between the rotating axis of the BH and the observer’s line of sight. ξ\xi and η\eta are two new parameters which are introduced as ξ=Lz/E\xi=L_{z}/E and η=Q/E2\eta=Q/E^{2}. For the θ=π/2\theta=\pi/2 plane, the celestial coordinates take the simple form: α=ξ,β=±η\alpha=-\xi,\beta=\pm\sqrt{\eta}.

From Eqs.(38) and (39) we see that the shadow boundaries are not affected by the deviation parameters ϵ3\epsilon_{3} and α52\alpha_{52}, but are only affected by α13\alpha_{13} and α22\alpha_{22}. In Fig.1 and Fig.2, we plot the shadow boundaries of Johannsen BH with different values of α13\alpha_{13} and α22\alpha_{22}. It’s interesting to see the effects of BH parameters by comparing Fig.1 and Fig.2. From the two figures we see that, the shadow size grow with both α13\alpha_{13} and α22\alpha_{22}. Fig.2 exhibits the shadow size grow explicitly with aa and θ0\theta_{0} as well, while the effects of shadow size growing with aa and θ0\theta_{0} are not so evident seen from Fig.1. This is in agreement with the direct calculations of shadow radii as shown in Fig.3 and Fig.4, in Fig.3 the curves are closer together while in Fig.4 the curves are more widely separated. Actually, as we will discuss below, there exist intersection points for the curves in Fig.3 and Fig.4, on the two sides of intersection points the behaviors of shadow size are distinct. If the shadow size grow with aa and θ0\theta_{0} on one side of the intersection point, it shrink with aa and θ0\theta_{0} on the other side of the intersection point, and vice versa. Fig.1 and Fig.2 show that shadows are deformed more greatly by larger aa and θ0\theta_{0}, and α13\alpha_{13}, α22\alpha_{22} contribute oppositely on deforming the shadows, the shadows are deformed more greatly by smaller α13\alpha_{13} but are more deformed by larger α22\alpha_{22}.

3.2 Constrain the BH parameters with observations

Aside from the theoretical constraints on BH parameters given by Eq.(35), the two relevant parameters that affect shadow boundaries, α13\alpha_{13} and α22\alpha_{22}, can be further constrained by observation data. For instance, the shadow radius observed by EHT collaboration can be used to constrain the BH parameters. For rotating BHs, the shadows of which are often not round, thus the definition of shadow radius for static BHs is not applicable here, we adopt the definition of shadow radius given in Refs.[113, 62]

Rs=(αtαr)2+βt22(αrαt).\displaystyle R_{s}=\frac{\left(\alpha_{t}-\alpha_{r}\right)^{2}+\beta_{t}^{2}}{2(\alpha_{r}-\alpha_{t})}. (41)

Here (αt,βt)(\alpha_{t},\beta_{t}), (αr,0)(\alpha_{r},0) are coordinates of the shadow vertices at top and right edges, which depend on the BH parameters through Eqs.(37) and (40), therefore the shadow radii would be affected by the BH parameters.

In this paper, we adopt the two observation results on shadow size of M87* and SgrA* released by EHT collaboration. For M87*, the observed angular gravitational radius reported by EHT collaboration is consistent within the range of 17%17\% with the one that was estimated through the analysis of stellar dynamics[1], which means that the shadow radius is bounded as [58]

4.31MRs6.08M.\displaystyle 4.31M\leq R_{s}\leq 6.08M. (42)

The estimates on mass-to-distance ratio of SgrA* were predominantly produced by two teams of EHT collaboration, Keck Observatory and Very Large Telescope Interferometer (VLTI)[2]. The shadow size bound given by Keck is

4.5MRs5.5M.\displaystyle 4.5M\leq R_{s}\leq 5.5M. (43)

and the one given by VLTI is

4.3MRs5.3M.\displaystyle 4.3M\leq R_{s}\leq 5.3M. (44)
Refer to caption
Refer to caption
Figure 4: Shadow radius as function of α13\alpha_{13} (left plot) or α22\alpha_{22} (right plot). The dashed and solid lines correspond to θ0=15\theta_{0}=15^{\circ} and 9090^{\circ} respectively. The red, green and blue lines correspond to a=0.2,0.5a=0.2,0.5 and 0.9 respectively. The green region represents the observation bound of SgrA* given by VLTI, the purple region represents the observation bound of SgrA* given by Keck, the blue region is the intersection sector of the green and purple regions. The orange vertical lines represent the case a=0.2a=0.2 and θ0=15\theta_{0}=15^{\circ}. Masses of the black holes are set to be unity M=1M=1.

In Fig.3 we plot shadow radii with different values of aa, θ0\theta_{0} and deviation parameters. For the left plot we fix α22=0\alpha_{22}=0 with α13\alpha_{13} nonvanishing, and for the right plot we fix α13=0\alpha_{13}=0 with α22\alpha_{22} nonvanishing. The blue regions describe the observation bound of M87*. Fig.3 exibits explicitly that, the shadow radii grow with the deviation parameters, within the theoretical allowed ranges of BH parameters, the theoretical results may exceed the observation bound. On the left plot, one sees the theoretical results of shadow radius may exceed both the upper and the lower observation bounds, i.e., too large and too small values of α13\alpha_{13} are all disfavored. On the right plot, one sees that the theoretical shadow radius corresponding to the lower theoretical bound of α22\alpha_{22} always lie in the range of observation bound, as α22\alpha_{22} increases the upper observation bound may be exceeded, which implies too large values of α22\alpha_{22} are disfavored. To see the observation constraints on the deviation parameters more explicitely, we plot two orange vertical lines for the case a=0.2a=0.2 and θ0=15\theta_{0}=15^{\circ} which is described by red dashed line. The values of α13\alpha_{13} between the two orange vertical lines are preferred by observation, while the values of α13\alpha_{13} beyond this range are disfavored by observation.

From Fig.3 one also sees that, for larger rotating parameter the shadow radii grow faster with the deviation parameters, this is in agreement with what one sees in Figs.1 and 2. Since the growth rates of the curves are different for different aa, there exist intersection points between the curves as shown in Fig.3. On the left side of the intersection points, the shadow radii are larger for smaller rotating parameters with fixed deviation parameter, while on the right side of the intersection points, the shadow radii are smaller for smaller rotating parameters. The effects of the rotating parameter are less significant for larger inclination angle, e.g., when θ0=90\theta_{0}=90^{\circ} the three curves with different aa almost coincide as shown on the left plot in Fig.3.

In Fig.4, we use the data of SgrA* shadow to constrain the BH parameters. The green region corresponds to the observation bound given by Keck, the purple region corresponds to the observation bound given by VLTI, and the blue region is the overlapping region of the green and purple regions. From the figure it’s easy to see, too large and too small values of α13\alpha_{13} are disfavored likewise, so as to the theoretical shadow radius does not exceed the observation bounds. Too large values of α22\alpha_{22} are disfavored since the observation upper bounds may be exceeded. While for the theoretical allowed minimum of α22\alpha_{22}, the corresponding shadow radii always lie in the range of the observation bounds. Likewise, we also plot two orange vertical lines to indicate the observation preferred range of the deviation parameter for the case a=0.2a=0.2 and θ0=15\theta_{0}=15^{\circ} corresponding to the red dashed line. Now, the values of α13\alpha_{13} between the two orange vertical lines are preferred by both Keck and VLTI bounds.

4 The image of an accretion disk of Johannsen BH

BHs accrete surrounding matter to form the accretion disk, which is expected to be an important luminous source of shadow, thus the information of accretion disk can be extracted through observing photons on the image plane of observers. We simulate the photons emitted from accretion disk by tracing backward along the trajectory of photon from the image plane to the accretion disk plane. The radiation flux from accretion disk corresponds to the luminosity on the image plane. Since accretion disks are quite complex physical systems which are difficult to simulate, for the purpose of this work, we use the simple Page-Thorne thin disk model to describe the radiation flux of accretion disk[114].

The Lagrangian of a free particle is given by

=12gαβq˙αq˙β,{\cal L}=\frac{1}{2}g_{\alpha\beta}\dot{q}^{\alpha}\dot{q}^{\beta}, (45)

where q˙αdqαdτ\dot{q}^{\alpha}\equiv\frac{dq^{\alpha}}{d\tau}. The Hamiltonian could be defined in the standard way

H=pσq˙σ=12gαβpαpβ,H=p_{\sigma}\dot{q}^{\sigma}-{\cal L}=\frac{1}{2}g^{\alpha\beta}p_{\alpha}p_{\beta}, (46)

where pαq˙α=gαβq˙βp_{\alpha}\equiv\frac{\partial{\cal L}}{\partial\dot{q}^{\alpha}}=g_{\alpha\beta}\dot{q}^{\beta} is the momentum conjugate to qαq^{\alpha}. For the Johannsen metric we obtain

H=ΔA5(r)2Σ~pr2+pθ22Σ~R(r)2ΔΣ~Θ2Σ~,\displaystyle H=\frac{\Delta A_{5}(r)}{2\tilde{\Sigma}}p_{r}^{2}+\frac{p_{\theta}^{2}}{2\tilde{\Sigma}}-\frac{R(r)}{2\Delta\tilde{\Sigma}}-\frac{\Theta}{2\tilde{\Sigma}}, (47)

from which we obtain the equations of motion of photon

r˙\displaystyle\dot{r} =ΔA5(r)Σ~pr,\displaystyle=\frac{\Delta A_{5}(r)}{\tilde{\Sigma}}p_{r}, (48)
p˙r\displaystyle\dot{p}_{r} =(ΔA5(r)2Σ~)pr2(12Σ~)pθ2+(R+ΔΘ2ΔΣ~),\displaystyle=-\left(\frac{\Delta A_{5}(r)}{2\tilde{\Sigma}}\right)^{\prime}p_{r}^{2}-\left(\frac{1}{2\tilde{\Sigma}}\right)^{\prime}p_{\theta}^{2}+\left(\frac{R+\Delta\Theta}{2\Delta\tilde{\Sigma}}\right)^{\prime},
θ˙\displaystyle\dot{\theta} =pθΣ~,\displaystyle=\frac{p_{\theta}}{\tilde{\Sigma}},
p˙θ\displaystyle\dot{p}_{\theta} =(ΔA5(r)2Σ~),θpr2(12Σ~),θpθ2+(R+ΔΘ2ΔΣ~),θ,\displaystyle=-\left(\frac{\Delta A_{5}(r)}{2\tilde{\Sigma}}\right)_{,\theta}p_{r}^{2}-\left(\frac{1}{2\tilde{\Sigma}}\right)_{,\theta}p_{\theta}^{2}+\left(\frac{R+\Delta\Theta}{2\Delta\tilde{\Sigma}}\right)_{,\theta},
t˙\displaystyle\dot{t} =12ΔΣ~E(R+ΔΘ),\displaystyle=\frac{1}{2\Delta\tilde{\Sigma}}\frac{\partial}{\partial E}\left(R+\Delta\Theta\right),
pt˙\displaystyle\dot{p_{t}} =0,\displaystyle=0,
ϕ˙\displaystyle\dot{\phi} =12ΔΣ~L(R+ΔΘ),\displaystyle=-\frac{1}{2\Delta\widetilde{\Sigma}}\frac{\partial}{\partial L}\left(R+\Delta\Theta\right),
pϕ˙\displaystyle\dot{p_{\phi}} =0,\displaystyle=0,

where denotes derivation with respect to rr, and the subscript ,θ,\theta denotes derivation with respect to θ\theta. We consider an observer viewing the central BH from a large distance dd (see Fig.5). According to the equations of motion (48), the photons with initial positions and momenta on the observer’s image plane can be traced backwards to the accretion disk so as to extract information of the accretion disk. For convenience, we convert the coordinates of the photon on the image plane at (α0,β0)(\alpha_{0},\beta_{0}) to Boyer-Lindquist coordinates[115, 116]

ri\displaystyle r_{i} =α02+β02+d2,\displaystyle=\sqrt{\alpha_{0}^{2}+\beta_{0}^{2}+d^{2}}, (49)
θi\displaystyle\theta_{i} =arccosβ0sinθ0+dcosθ0ri,\displaystyle=\arccos{\frac{\beta_{0}\sin\theta_{0}+d\cos\theta_{0}}{r_{i}}},
ϕi\displaystyle\phi_{i} =arctanα0dsinθ0β0cosθ0,\displaystyle=\arctan{\frac{\alpha_{0}}{d\sin\theta_{0}-\beta_{0}\cos\theta_{0}}},

where θ0\theta_{0} is the inclination angle of the photon as shown in Fig.5. The photons that can be observed on the image plane are those with momentum perpendicular to the image plane and against the line of sight, the components of the photon momentum k0\vec{k}_{0} are given by

kr\displaystyle k_{r} =drik0,\displaystyle=\frac{d}{r_{i}}k_{0},
kθ\displaystyle k_{\rm\theta} =cosθ0+(β0sinθ0+dcosθ0)dri2α02+(dsinθ0ycosθ0)2k0,\displaystyle=\frac{-\cos\theta_{0}+(\beta_{0}\sin\theta_{0}+d\cos\theta_{0})\frac{d}{r_{i}^{2}}}{\sqrt{\alpha_{0}^{2}+\left(d\sin\theta_{0}-y^{\prime}\cos\theta_{0}\right)^{2}}}k_{0},
kϕ\displaystyle k_{\rm\phi} =α0sinθ0α02+(dsinθ0β0cosθ0)2k0.\displaystyle=\frac{-\alpha_{0}\sin\theta_{0}}{\alpha_{0}^{2}+(d\sin\theta_{0}-\beta_{0}\cos\theta_{0})^{2}}k_{0}.
Refer to caption
Figure 5: The geometry used for tracing trajectories of photons.

.

The radiation flux of accretion disk is closely related to circular geodesic motion of baryons on the equatorial plane. Obviously there are two conserved quantities for the baryons on circular orbits, energy and angular momentum. With the external spacetime metric, energy and angular momentum can be expressed as pt=E=gttt˙+gtϕϕ˙p_{t}=-E=g_{tt}\dot{t}+g_{t\phi}\dot{\phi}, pϕ=L=gtϕt˙+gϕϕϕ˙p_{\phi}=L=g_{t\phi}\dot{t}+g_{\phi\phi}\dot{\phi}, with which the geodesic motion on the equatorial plane can be expressed as

dtdτ\displaystyle\frac{dt}{d\tau} =Egϕϕ+Lgtϕgtϕ2gttgϕϕ,\displaystyle=\frac{Eg_{\phi\phi}+Lg_{t\phi}}{g_{t\phi}^{2}-g_{tt}g_{\phi\phi}}\,, (50)
dϕdτ\displaystyle\frac{d\phi}{d\tau} =Egtϕ+Lgttgtϕ2gttgϕϕ,\displaystyle=-\frac{Eg_{t\phi}+Lg_{tt}}{g_{t\phi}^{2}-g_{tt}g_{\phi\phi}}\,, (51)
grr(drdτ)2\displaystyle g_{rr}\left(\frac{dr}{d\tau}\right)^{2} =1+E2gϕϕ+2ELgtϕ+L2gttgtϕ2gttgϕϕ.\displaystyle=-1+\frac{E^{2}g_{\phi\phi}+2ELg_{t\phi}+L^{2}g_{tt}}{g_{t\phi}^{2}-g_{tt}g_{\phi\phi}}\,. (52)

In the above derivation we use the normalization condition pαpα=m02p_{\alpha}p^{\alpha}=m_{0}^{2} and set the mass of the massive particle m0m_{0} to be 1. Note that θ˙=0\dot{\theta}=0, there is no particle motion in the direction of θ\theta. Eqs.(50) and (51) allow us to re-express energy and angular momentum as

E\displaystyle E =\displaystyle= gttgtϕΩgtt2gtϕΩgϕϕΩ2,\displaystyle\frac{g_{tt}-g_{t\phi}\Omega}{\sqrt{g_{tt}-2g_{t\phi}\Omega-g_{\phi\phi}\Omega^{2}}}\,, (53)
L\displaystyle L =\displaystyle= gtϕ+gϕϕΩgtt2gtϕΩgϕϕΩ2,\displaystyle\frac{g_{t\phi}+g_{\phi\phi}\Omega}{\sqrt{g_{tt}-2g_{t\phi}\Omega-g_{\phi\phi}\Omega^{2}}}\,, (54)

where Ω=dϕ/dt\Omega=d\phi/dt is the angular velocity of the massive particle on the equatorial plane. The circular equatorial orbits require r˙=r¨=0\dot{r}=\ddot{r}=0, this condition along with Eq.(52) give rise to the circular orbits condition gtt,r+2gtϕ,rΩ+gϕϕ,rΩ2=0g_{tt,r}+2g_{t\phi,r}\Omega+g_{\phi\phi,r}\Omega^{2}=0, from which one obtains the angular velocity

Ω=dϕdt=gtϕ,r±(gtϕ,r)2gtt,rgϕϕ,rgϕϕ,r.\displaystyle\Omega=\frac{d\phi}{dt}=\frac{-g_{t\phi,r}\pm\sqrt{(g_{t\phi,r})^{2}-g_{tt,r}g_{\phi\phi,r}}}{g_{\phi\phi,r}}\,. (55)

From Eq.(52) one may introduce an effective potential

Veff(r)=1+E2gϕϕ+2ELgtϕ+L2gttgtϕ2gttgϕϕ.V_{\textmd{\scriptsize{eff}}}(r)=-1+\frac{E^{2}g_{\phi\phi}+2ELg_{t\phi}+L^{2}g_{tt}}{g_{t\phi}^{2}-g_{tt}g_{\phi\phi}}\,. (56)

The position of the innermost stable circular orbit (ISCO) is determined by the effective potential through the conditions

Veff(r)=Veff(r)=Veff′′(r)=0.\displaystyle V_{\textmd{\scriptsize{eff}}}(r)=V_{\textmd{\scriptsize{eff}}}^{\prime}(r)=V_{\textmd{\scriptsize{eff}}}^{\prime\prime}(r)=0\,. (57)

In Ref.[114], Page and Thorne considered the geometrically thin accretion disk, where the matter content is distributed on the equatorial plane. Under some reasonable assumptions and by application of the conservation laws of energy, angular momentum and rest mass, the radiation flux can be expressed as

F(r)=M˙4πgΩ,r(EΩL)2rmsr(EΩL)L,r𝑑r,\displaystyle F(r)=-\frac{\dot{M}}{4\pi\sqrt{-g}}\frac{\Omega_{,r}}{\left(E-\Omega L\right)^{2}}\int_{r_{ms}}^{r}\left(E-\Omega L\right)L_{,r}dr\,, (58)

which is the time-averaged flux of radiant energy flowing out of the face of accretion disk. Here M˙\dot{M} is the accretion rate, rmsr_{ms} is the radius of the ISCO or marginally stable circular geodesic orbit. We know the effect of redshift should be taken into account for the observed radiation intensity, the redshift factor is defined as g=pμuoμ/pνueνg=p_{\mu}u_{o}^{\mu}/p_{\nu}u_{e}^{\nu}, where pμp_{\mu} is the conjugate momentum of photon, uoμu_{o}^{\mu} is the four velocity of observer, and ueμu_{e}^{\mu} is the four velocity of emitting source.

To obtain shadow of Johannsen BH we need to trace the trajectories of photons that reach the observer. Since most of the photons would not reach the observer, direct method of tracing photons from the light source would waste lots of computation resources. Thus we use the open-source code Gyoto which traces the light rays backwards from the image plane of observer to the plane of accretion disk[117]. The trajectory integrations are performed according to Eqs.(48), which are solved numerically by the fourth order Runge-Kutta algorithm with an adaptive step. The initial momentum of the photon on the image plane is given by Eq.(4). We modify the code to be applicable to Johannsen BH. Since Johannsen BH possesses 4 deviation parameters, we would like to study the effects of the parameters on the accretion disk image one by one, like what we done in studying shadow boundary when examining one parameter we set other three parameters to be 0. We assume the image plane to be at large distance from the source BH, in the code we set the distance d=100Md=100M. We take two inclination angles θ0=15\theta_{0}=15^{\circ} and θ0=60\theta_{0}=60^{\circ} to show the accretion disk images on the image plane.

Refer to caption
Refer to caption
Figure 6: Accretion disk images of Johannsen BH with different values of α13\alpha_{13}, aa and θ0\theta_{0}. The left panel correspond to θ0=15\theta_{0}=15^{\circ}, the right one correspond to θ0=60\theta_{0}=60^{\circ}. For the row a=0a=0, from left to right α13=7,0,50,100\alpha_{13}=-7,0,50,100 and 200200 respectively; for the row a=0.2a=0.2, α13=7.5,0,10,50\alpha_{13}=-7.5,0,10,50, 100100; for the row a=0.5a=0.5, α13=6,0,10,30\alpha_{13}=-6,0,10,30, 5050; and for the row a=0.7a=0.7, α13=7.5,0,5,10\alpha_{13}=-7.5,0,5,10, 2020. As comparision, the second columns on both panels correspond to the images of Kerr BH with α13=0\alpha_{13}=0. Masses of the black holes have been set to be M=1M=1.
Refer to caption
Refer to caption
Figure 7: Accretion disk images of Johannsen BH with different values of α22\alpha_{22}, aa and θ0\theta_{0}. The left panel correspond to θ0=15\theta_{0}=15^{\circ}, the right one correspond to θ0=60\theta_{0}=60^{\circ}. For the row a=0a=0, α22=3.9,0,100,200\alpha_{22}=-3.9,0,100,200, 300300; for the row a=0.2a=0.2, α22=3.8,0,50,100\alpha_{22}=-3.8,0,50,100, 200200; for the row a=0.5a=0.5, α22=3,0,20,30\alpha_{22}=-3,0,20,30, 4040; and for the row a=0.7a=0.7, α22=2.9,0,5,10\alpha_{22}=-2.9,0,5,10, 1515. The second column on each panel corresponds to Kerr BH with α22=0\alpha_{22}=0. We have set M=1M=1.

.

Refer to caption
Refer to caption
Figure 8: Accretion disk images of Johannsen BH with different values of α52\alpha_{52}, aa and θ0\theta_{0}. The left panel correspond to θ0=15\theta_{0}=15^{\circ}, the right one correspond to θ0=60\theta_{0}=60^{\circ}. For the row a=0a=0, α52=3.5,0,100,200\alpha_{52}=-3.5,0,100,200, 300300; for the row a=0.5a=0.5, α52=3,0,100,200,300\alpha_{52}=-3,0,100,200,300; for the row a=0.9a=0.9, α52=2,0,100,200,300\alpha_{52}=-2,0,100,200,300; and for a=0.99a=0.99, α52=1,0,100,200,300\alpha_{52}=-1,0,100,200,300. The second column on each panel corresponds to Kerr BH with α52=0\alpha_{52}=0. We have set M=1M=1.

.

Refer to caption
Refer to caption
Figure 9: Accretion disk images of Johannsen BH with different values of ϵ3\epsilon_{3}, aa and θ0\theta_{0}. The left panel correspond to θ0=15\theta_{0}=15^{\circ}, the right one correspond to θ0=60\theta_{0}=60^{\circ}. For the row a=0a=0, ϵ3=7,0,25,50,100\epsilon_{3}=-7,0,25,50,100; for the row a=0.5a=0.5, ϵ3=6,0,10,25,50\epsilon_{3}=-6,0,10,25,50; for the row a=0.9a=0.9, ϵ3=2,0,10,25,50\epsilon_{3}=-2,0,10,25,50; and for the row a=0.99a=0.99, ϵ3=1,0,10,25,50\epsilon_{3}=-1,0,10,25,50. The second column on each panel corresponds to Kerr BH with ϵ3=0\epsilon_{3}=0. We have set M=1M=1.

.

Figs.6-9 exhibit the accretion disk images with different values of spin parameter aa, inclination angle θ0\theta_{0} and deviation parameters α13,α22,α52,ϵ3\alpha_{13},\alpha_{22},\alpha_{52},\epsilon_{3}. To facilitate the comparison between images with different BH parameters, we put the images of Kerr BH in the second column in each figure. From Fig.6 one sees that, α13\alpha_{13} affects the sizes of the central dark regions of accretion disk images, the sizes of central dark regions increase with α13\alpha_{13}. For α13>0\alpha_{13}>0, the sizes of central dark region are larger than that of Kerr BH. The effects of α22\alpha_{22} on accretion disk images are not so evident as shown in Fig.7. From Fig.8 one sees that, α52\alpha_{52} affects both the brightness of the images and the sizes of central dark regions. The images become darker and darker as α52\alpha_{52} increases, the positions of the regions with high luminosity vary with α52\alpha_{52} as well. The central dark regions shrink with α52\alpha_{52} for some values of aa, but extend with α52\alpha_{52} for other values of aa. The shapes of central dark regions vary with α52\alpha_{52} as well. Fig.9 shows that, as ϵ3\epsilon_{3} increases the images become darker and the central dark regions shrink. All the Figs.6-9 exhibit that the central dark regions shrink with the spin parameter aa. As we can see, the accretion disk images vary with BH parameters evidently, therefore the effects of BH parameters on BH shadow can be tested by future observations.

5 Discussion and Conclusion

Now let’s discuss the results we obtained. Our calculations show that the shadow boundaries are affected by the deviation parameters α13\alpha_{13} and α22\alpha_{22}. Though calculating the radii of shadows and comparing with the observations of M87* and SgrA*, we constrained the deviation parameters with BH image data motivated to test no-hair theorem. Our results showed that current observations cannot constrain the deviation parameters to be pure positive or pure negative, but constrain the deviation parameters to lie in a neighborhood of zero, which implies the current observations on BH image can’t confirm the existence of BH hair. Our constraints on deviation parameters can be used to cross check with other constraints by other observations[89, 90].

We plot accretion disk image with ray-tracing method. The figures in our paper show that accretion disk images of Johannsen BH may differ from that of Kerr BH significantly. For α13<0\alpha_{13}<0 the size of central dark region is smaller than that of Kerr BH. As α13\alpha_{13} increases, the central dark region extends. For α13>0\alpha_{13}>0 the size of central dark region is larger than that of Kerr BH. The effects of α22\alpha_{22} are not very evident. The deviation parameter α52\alpha_{52} affects the brightness of the images. As α52\alpha_{52} increases, the images become darker compared to that of Kerr BH, the central dark regions extend for some values of aa and shrink for other values of aa. Moreover, the position of the shadow’s bright region changes with α52\alpha_{52}. Likewise, the deviation parameter ϵ3\epsilon_{3} affects the brightness of the images as well. As ϵ3\epsilon_{3} increases, the images become darker and the sizes of the central dark region become smaller compared to that of Kerr BH. The larger the value of the spin parameter aa, the more evident the phenomenon is. When aa approaches 1, the images become very dark for large values of ϵ3\epsilon_{3} and α52\alpha_{52}. These properties can be used to test no-hair theorem with future observations.

In this paper, we investigate the shadow of a general rotating BH, that is, the Johannsen BH. Johannsen BH is a parameterized generalization of Kerr BH which admits three constants of motion, i.e., mass, angular momentum and Carter constant. The BH is characterized by mass, rotating parameters and four deviation parameters α13\alpha_{13}, α22\alpha_{22}, α52\alpha_{52} and ϵ3\epsilon_{3} to leading truncation order. Johannsen BH is not solution of any particular gravity theory, but it can be mapped to several known BH solutions of familiar theories of modified gravity. The universal features of Johannsen BH make it is very suitable for testing no-hair theorem in a model-independent way.

We plot the shadow boundaries on the plane of celestial coordinates. Among the four deviation parameters, only α13\alpha_{13} and α22\alpha_{22} get into the expression of R(r)R(r), α52\alpha_{52} and ϵ3\epsilon_{3} don’t appear in the expression of R(r)R(r), thus the shadow boundaries are only affected by α13\alpha_{13} and α22\alpha_{22}. The sizes of shadow boundaries grow with both α13\alpha_{13} and α22\alpha_{22} as can be seen from Fig.1 and Fig.2. The shadow boundaries are more deformed by larger rotating parameter aa and larger inclination angle θ0\theta_{0}. The shadow boundaries are deformed more greatly by larger α22\alpha_{22}, but are deformed more greatly by smaller α13\alpha_{13}.

Since the two deviation parameters, α13\alpha_{13} and α22\alpha_{22}, affect the shadow boundary, they will also affect the shadow radius. Thus the observations on shadow radii can be used to constrain the deviation parameters. We use the observation results of M87* and SgrA* by EHT collaboration to constrain these parameters. Within the theoretically allowed ranges of the deviation parameters, the theoretical shadow radii varying with α13\alpha_{13} may exceed both the upper and the lower observation bounds, which means too large and too small values of α13\alpha_{13} are all disfavored. The shadow radii varying with α22\alpha_{22} may exceed the upper observation bound while the lower observation bound are not exceeded, thus large values of α22\alpha_{22} are disfavored. Therefore the deviation parameters are constrained in finite ranges with the observation results. There exist intersection points for the theoretical curves, on the left of the intersection points shadow radii are smaller for larger rotating parameters, while on the right of the intersection points the shadow radii are larger for larger rotating parameters.

Accretion disk is an important light source of the shadow, the accreted matter surrounding the central compact object emit photons which may reach the image plane of remote observer. Supposing the accretion disk to be a Page-Thorne thin disk, we draw the images of accretion disk through ray-tracing method. We study the effects of the BH parameters and learn that, the sizes of the central dark regions of the accretion disk images grow with α13\alpha_{13}, while the central dark regions are not affected evidently by α22\alpha_{22}. The shadows become darker as α52\alpha_{52} or ϵ3\epsilon_{3} increases, and the position of the region with high luminosity changes with α52\alpha_{52}. Moreover, for fixed deviation parameters, the central dark regions shrink with increase of the rotating parameter aa. Since the BH shadows are affected evidently by BH parameters, the upcoming observations of BH shadow can be used to constrain modified gravities in a unified way.

Acknowledgment

We would like to thank N.Yang for useful discussions. This work is supported by the National Natural Science Foundation of China Grants Nos.11922303, 12275106, 12235019 and 12173071, and by Natural Science Foundation of Shandong Province No.ZR2023MA014, and The National Key R&\&D Program of China (no. 2021YFC2203002). X. Fan is also supported by the Fundamental Research Funds for the Central Universitiesv(2042022kf1182). W. Han is also supported by CAS Project for Young Scientists in Basic Research YSBR-006.

References

  • [1] K. Akiyama et al. [Event Horizon Telescope], First M87 Event Horizon Telescope Results. VI. The Shadow and Mass of the Central Black Hole, Astrophys. J. Lett. 875, no.1, (2019) L6 [arXiv:1906.11243 [astro-ph.GA]].
  • [2] K. Akiyama et al. [Event Horizon Telescope], First Sagittarius A* Event Horizon Telescope Results. VI. Testing the Black Hole Metric, Astrophys. J. Lett. 930, no.2, L17 (2022).
  • [3] B. P. Abbott et al. [LIGO Scientific and Virgo], Observation of Gravitational Waves from a Binary Black Hole Merger, Phys. Rev. Lett. 116, no.6, 061102 (2016) [arXiv:1602.03837 [gr-qc]].
  • [4] B. P. Abbott et al. [LIGO Scientific and Virgo], GW151226: Observation of Gravitational Waves from a 22-Solar-Mass Binary Black Hole Coalescence, Phys. Rev. Lett. 116, no.24, 241103 (2016) [arXiv:1606.04855 [gr-qc]].
  • [5] W. Israel, Event horizons in static vacuum space-times, Phys. Rev. 164, 1776-1779 (1967).
  • [6] B. Carter, Axisymmetric Black Hole Has Only Two Degrees of Freedom, Phys. Rev. Lett. 26, 331-333 (1971).
  • [7] S. W. Hawking, Black holes in general relativity, Commun. Math. Phys. 25, 152-166 (1972).
  • [8] C. A. R. Herdeiro and E. Radu, Kerr black holes with scalar hair, Phys. Rev. Lett. 112, 221101 (2014) [arXiv:1403.2757 [gr-qc]].
  • [9] C. Herdeiro, E. Radu and H. Rúnarsson, Kerr black holes with Proca hair, Class. Quant. Grav. 33, no.15, 154001 (2016) [arXiv:1603.02687 [gr-qc]].
  • [10] C. Bambi and L. Modesto, Rotating regular black holes, Phys. Lett. B 721, 329-334 (2013) [arXiv:1302.6075 [gr-qc]].
  • [11] M. Zubair, M. A. Raza and G. Abbas, Optical features of rotating black hole with nonlinear electrodynamics, Eur. Phys. J. C 82, no.10, 948 (2022) [arXiv:2210.13750 [gr-qc]].
  • [12] J. R. Mureika and G. U. Varieschi, Black hole shadows in fourth-order conformal Weyl gravity, Can. J. Phys. 95, no.12, 1299-1306 (2017) [arXiv:1611.00399 [gr-qc]].
  • [13] T. Anson, E. Babichev, C. Charmousis and M. Hassaine, Disforming the Kerr metric, JHEP 01, 018 (2021) [arXiv:2006.06461 [gr-qc]].
  • [14] D. D. K. Chow, Single-charge rotating black holes in four-dimensional gauged supergravity, Class. Quant. Grav. 28, 032001 (2011) [arXiv:1011.2202 [hep-th]].
  • [15] H. S. Liu and H. Lu, Charged Rotating AdS Black Hole and Its Thermodynamics in Conformal Gravity, JHEP 02, 139 (2013) [arXiv:1212.6264 [hep-th]].
  • [16] Y. Ling and M. H. Wu, Regular black holes with sub-Planckian curvature, Class. Quant. Grav. 40, no.7, 075009 (2023) [arXiv:2109.05974 [gr-qc]].
  • [17] J. Ben Achour, H. Liu, H. Motohashi, S. Mukohyama and K. Noui, On rotating black holes in DHOST theories, JCAP 11, 001 (2020) [arXiv:2006.07245 [gr-qc]].
  • [18] G. Z. Babar, F. Atamurotov, S. Ul Islam and S. G. Ghosh, Particle acceleration around rotating Einstein-Born-Infeld black hole and plasma effect on gravitational lensing, Phys. Rev. D 103, no.8, 084057 (2021) [arXiv:2104.00714 [gr-qc]].
  • [19] J. Naji and S. Heshmatian, Quark–gluon plasma corresponding to a charged rotating hairy black hole, Can. J. Phys. 94, no.6, 574-577 (2016).
  • [20] A. Das, A. Saha and S. Gangopadhyay, Study of circular geodesics and shadow of rotating charged black hole surrounded by perfect fluid dark matter immersed in plasma, Class. Quant. Grav. 39, no.7, 075005 (2022) [arXiv:2110.11704 [gr-qc]].
  • [21] K. Jusufi, M. Azreg-Aïnou, M. Jamil, S. W. Wei, Q. Wu and A. Wang, Quasinormal modes, quasiperiodic oscillations, and the shadow of rotating regular black holes in nonminimally coupled Einstein-Yang-Mills theory, Phys. Rev. D 103, no.2, 024013 (2021) [arXiv:2008.08450 [gr-qc]].
  • [22] B. Toshmatov, Z. Stuchlík and B. Ahmedov, Rotating black hole solutions with quintessential energy, Eur. Phys. J. Plus 132, no.2, 98 (2017) [arXiv:1512.01498 [gr-qc]].
  • [23] Z. Xu, J. Wang and X. Hou, Kerr–anti-de Sitter/de Sitter black hole in perfect fluid dark matter background, Class. Quant. Grav. 35, no.11, 115003 (2018) [arXiv:1711.04538 [gr-qc]].
  • [24] T. Johannsen, Regular Black Hole Metric with Three Constants of Motion, Phys. Rev. D 88, no.4, 044002 (2013) [arXiv:1501.02809 [gr-qc]].
  • [25] J. L. Synge, The Escape of Photons from Gravitationally Intense Stars, Mon. Not. Roy. Astron. Soc. 131, no.3, 463-466 (1966).
  • [26] J. M. Bardeen, in Black Holes (Les Astres Occlus), edited by C. Dewitt and B. S. Dewitt (Gordon and Breach, 1973), pp. 215–239.
  • [27] V. Perlick, O. Y. Tsupko and G. S. Bisnovatyi-Kogan, Influence of a plasma on the shadow of a spherically symmetric black hole, Phys. Rev. D 92, no.10, 104031 (2015) [arXiv:1507.04217 [gr-qc]].
  • [28] A. Abdujabbarov, B. Toshmatov, Z. Stuchlík and B. Ahmedov, Shadow of the rotating black hole with quintessential energy in the presence of plasma, Int. J. Mod. Phys. D 26, no.06, 1750051 (2016) [arXiv:1512.05206 [gr-qc]].
  • [29] R. A. Konoplya, Shadow of a black hole surrounded by dark matter Phys. Lett. B 795, 1-6 (2019) [arXiv:1905.00064 [gr-qc]].
  • [30] Z. Chang and Q. H. Zhu, Black hole shadow in the view of freely falling observers, JCAP 06, 055 (2020) [arXiv:1911.02190 [gr-qc]].
  • [31] P. C. Li, M. Guo and B. Chen, Shadow of a Spinning Black Hole in an Expanding Universe Phys. Rev. D 101, no.8, 084041 (2020) [arXiv:2001.04231 [gr-qc]].
  • [32] D. Psaltis et al. [Event Horizon Telescope], Gravitational Test Beyond the First Post-Newtonian Order with the Shadow of the M87 Black Hole, Phys. Rev. Lett. 125, no.14, 141104 (2020) [arXiv:2010.01055 [gr-qc]].
  • [33] F. Atamurotov, U. Papnoi and K. Jusufi, Shadow and deflection angle of charged rotating black hole surrounded by perfect fluid dark matter, Class. Quant. Grav. 39, no.2, 025014 (2022) [arXiv:2104.14898 [gr-qc]].
  • [34] A. Chael, M. D. Johnson and A. Lupsasca, Observing the Inner Shadow of a Black Hole: A Direct View of the Event Horizon, Astrophys. J. 918, no.1, 6 (2021) [arXiv:2106.00683 [astro-ph.HE]].
  • [35] J. Badía and E. F. Eiroa, Shadow of a charged black hole surrounded by an anisotropic matter field, [arXiv:2111.05732 [gr-qc]].
  • [36] G. Briozzo, E. Gallo and T. Mädler, Shadows of rotating black holes in plasma environments with aberration effects, Phys. Rev. D 107, no.12, 124004 (2023) [arXiv:2211.05620 [gr-qc]].
  • [37] P. Kocherlakota et al. [Event Horizon Telescope], Constraints on black-hole charges with the 2017 EHT observations of M87*, Phys. Rev. D 103, no.10, 104047 (2021) [arXiv:2105.09343 [gr-qc]].
  • [38] L. Chakhchi, H. El Moumni and K. Masmar, Shadows and optical appearance of a power-Yang-Mills black hole surrounded by different accretion disk profiles, Phys. Rev. D 105, no.6, 064031 (2022).
  • [39] P. V. P. Cunha, J. Grover, C. Herdeiro, E. Radu, H. Runarsson and A. Wittig, Chaotic lensing around boson stars and Kerr black holes with scalar hair Phys. Rev. D 94, no.10, 104023 (2016) [arXiv:1609.01340 [gr-qc]].
  • [40] R. Kumar, S. G. Ghosh and A. Wang, Gravitational deflection of light and shadow cast by rotating Kalb-Ramond black holes, Phys. Rev. D 101, no.10, 104001 (2020) [arXiv:2001.00460 [gr-qc]].
  • [41] S. W. Wei, Y. C. Zou, Y. X. Liu and R. B. Mann, Curvature radius and Kerr black hole shadow JCAP 08, 030 (2019) [arXiv:1904.07710 [gr-qc]].
  • [42] A. Belhaj, M. Benali, A. El Balali, H. El Moumni and S. E. Ennadifi, Deflection angle and shadow behaviors of quintessential black holes in arbitrary dimensions Class. Quant. Grav. 37, no.21, 215004 (2020) [arXiv:2006.01078 [gr-qc]].
  • [43] A. Belhaj, H. Belmahi, M. Benali, W. El Hadri, H. El Moumni and E. Torrente-Lujan, Shadows of 5D black holes from string theory, Phys. Lett. B 812, 136025 (2021) [arXiv:2008.13478 [hep-th]].
  • [44] S. E. Gralla, D. E. Holz and R. M. Wald, Black Hole Shadows, Photon Rings, and Lensing Rings, Phys. Rev. D 100, no.2, 024018 (2019) [arXiv:1906.00873 [astro-ph.HE]].
  • [45] A. B. Abdikamalov, A. A. Abdujabbarov, D. Ayzenberg, D. Malafarina, C. Bambi and B. Ahmedov, Black hole mimicker hiding in the shadow: Optical properties of the γ\gamma metric, Phys. Rev. D 100, no.2, 024014 (2019) [arXiv:1904.06207 [gr-qc]].
  • [46] M. Okyay and A. Övgün, Nonlinear electrodynamics effects on the black hole shadow, deflection angle, quasinormal modes and greybody factors, JCAP 01, no.01, 009 (2022) [arXiv:2108.07766 [gr-qc]].
  • [47] İ. Çimdiker, D. Demir and A. Övgün, Black hole shadow in symmergent gravity, Phys. Dark Univ. 34, 100900 (2021) [arXiv:2110.11904 [gr-qc]].
  • [48] Y. Kumaran and A. Övgün, Deflection Angle and Shadow of the Reissner–Nordström Black Hole with Higher-Order Magnetic Correction in Einstein-Nonlinear-Maxwell Fields, Symmetry 14, no.10, 2054 (2022) [arXiv:2210.00468 [gr-qc]].
  • [49] A. Övgün and İ. Sakallı, Testing generalized Einstein–Cartan–Kibble–Sciama gravity using weak deflection angle and shadow cast, Class. Quant. Grav. 37, no.22, 225003 (2020) [arXiv:2005.00982 [gr-qc]].
  • [50] A. Övgün, İ. Sakallı and J. Saavedra, Shadow cast and Deflection angle of Kerr-Newman-Kasuya spacetime, JCAP 10, 041 (2018) [arXiv:1807.00388 [gr-qc]].
  • [51] C. A. R. Herdeiro, Black holes: on the universality of the Kerr hypothesis, [arXiv:2204.05640 [gr-qc]].
  • [52] P. V. P. Cunha, C. A. R. Herdeiro, E. Radu and H. F. Runarsson, Shadows of Kerr black holes with scalar hair, Phys. Rev. Lett. 115, no.21, 211102 (2015) [arXiv:1509.00021 [gr-qc]].
  • [53] P. V. P. Cunha, C. A. R. Herdeiro, E. Radu and H. F. Runarsson, Shadows of Kerr black holes with and without scalar hair, Int. J. Mod. Phys. D 25, no.09, 1641021 (2016) [arXiv:1605.08293 [gr-qc]].
  • [54] F. Long, S. Chen, M. Wang and J. Jing, Shadow of a disformal Kerr black hole in quadratic degenerate higher-order scalar–tensor theories, Eur. Phys. J. C 80, no.12, 1180 (2020) [arXiv:2009.07508 [gr-qc]].
  • [55] M. Afrin, R. Kumar and S. G. Ghosh, Parameter estimation of hairy Kerr black holes from its shadow and constraints from M87*, Mon. Not. Roy. Astron. Soc. 504, 5927-5940 (2021) [arXiv:2103.11417 [gr-qc]].
  • [56] A. Abdujabbarov, M. Amir, B. Ahmedov and S. G. Ghosh, Shadow of rotating regular black holes, Phys. Rev. D 93, no.10, 104004 (2016) [arXiv:1604.03809 [gr-qc]].
  • [57] N. Tsukamoto, Black hole shadow in an asymptotically-flat, stationary, and axisymmetric spacetime: The Kerr-Newman and rotating regular black holes, Phys. Rev. D 97, no.6, 064021 (2018) [arXiv:1708.07427 [gr-qc]].
  • [58] S. V. M. C. B. Xavier, H. C. D. Lima, Junior. and L. C. B. Crispino, Shadows of black holes with dark matter halo, Phys. Rev. D 107, no.6, 064040 (2023) [arXiv:2303.17666 [gr-qc]].
  • [59] G. Lambiase, R. C. Pantig, D. J. Gogoi and A. Övgün, Investigating the Connection between Generalized Uncertainty Principle and Asymptotically Safe Gravity in Black Hole Signatures through Shadow and Quasinormal Modes, [arXiv:2304.00183 [gr-qc]].
  • [60] J. Badía and E. F. Eiroa, Influence of an anisotropic matter field on the shadow of a rotating black hole, Phys. Rev. D 102, no.2, 024066 (2020) [arXiv:2005.03690 [gr-qc]].
  • [61] K. Paithankar and S. Kolekar, Black hole shadow and acceleration bounds for spherically symmetric spacetimes, [arXiv:2305.07444 [gr-qc]].
  • [62] Y. Ling and M. H. Wu, The Shadows of Regular Black Holes with Asymptotic Minkowski Cores, Symmetry 14, no.11, 2415 (2022) [arXiv:2205.08919 [gr-qc]].
  • [63] H. C. D. Lima, Junior., L. C. B. Crispino, P. V. P. Cunha and C. A. R. Herdeiro, Can different black holes cast the same shadow?, Phys. Rev. D 103, no.8, 084040 (2021) [arXiv:2102.07034 [gr-qc]].
  • [64] S. W. Wei and Y. X. Liu, Testing the nature of Gauss-Bonnet gravity by four-dimensional rotating black hole shadow, Eur. Phys. J. Plus 136, no.4, 436 (2021) [arXiv:2003.07769 [gr-qc]].
  • [65] H. M. Wang and S. W. Wei, Shadow cast by Kerr-like black hole in the presence of plasma in Einstein-bumblebee gravity, Eur. Phys. J. Plus 137, no.5, 571 (2022) [arXiv:2106.14602 [gr-qc]].
  • [66] H. M. Wang, Y. M. Xu and S. W. Wei, Shadows of Kerr-like black holes in a modified gravity theory, JCAP 03, 046 (2019) [arXiv:1810.12767 [gr-qc]].
  • [67] M. Guo and P. C. Li, Innermost stable circular orbit and shadow of the 4D4D Einstein–Gauss–Bonnet black hole, Eur. Phys. J. C 80, no.6, 588 (2020) [arXiv:2003.02523 [gr-qc]].
  • [68] S. I. Kruglov, The shadow of M87* black hole within rational nonlinear electrodynamics, Mod. Phys. Lett. A 35, no.35, 2050291 (2020) [arXiv:2009.07657 [gr-qc]].
  • [69] M. A. Anacleto, J. A. V. Campos, F. A. Brito and E. Passos, Quasinormal modes and shadow of a Schwarzschild black hole with GUP, Annals Phys. 434, 168662 (2021) [arXiv:2108.04998 [gr-qc]].
  • [70] F. Atamurotov, K. Jusufi, M. Jamil, A. Abdujabbarov and M. Azreg-Aïnou, Axion-plasmon or magnetized plasma effect on an observable shadow and gravitational lensing of a Schwarzschild black hole, Phys. Rev. D 104, no.6, 064053 (2021) [arXiv:2109.08150 [gr-qc]].
  • [71] X. Wu and X. Zhang, Connections between the Shadow Radius and the Quasinormal Modes of Kerr-Sen Black Hole, Universe 8, no.11, 604 (2022) [arXiv:2112.11066 [gr-qc]].
  • [72] W. D. Guo, S. W. Wei and Y. X. Liu, Shadow of a charged black hole with scalar hair, Eur. Phys. J. C 83, no.3, 197 (2023) [arXiv:2203.13477 [gr-qc]].
  • [73] S. Kala, H. Nandan and P. Sharma, Shadow and weak gravitational lensing of a rotating regular black hole in a non-minimally coupled Einstein-Yang-Mills theory in the presence of plasma, Eur. Phys. J. Plus 137, no.4, 457 (2022) [arXiv:2207.10717 [gr-qc]].
  • [74] R. C. Pantig, A. Övgün and D. Demir, Testing symmergent gravity through the shadow image and weak field photon deflection by a rotating black hole using the M87 and Sgr. A\hbox{A}^{*} results, Eur. Phys. J. C 83, no.3, 250 (2023) [arXiv:2208.02969 [gr-qc]].
  • [75] S. K. Jha and A. Rahaman, Superradiance scattering off rotating Simpson-Visser black hole and its shadow in the non-commutative setting, [arXiv:2208.13176 [gr-qc]].
  • [76] X. X. Zeng, M. I. Aslam and R. Saleem, The optical appearance of charged four-dimensional Gauss–Bonnet black hole with strings cloud and non-commutative geometry surrounded by various accretions profiles, Eur. Phys. J. C 83, no.2, 129 (2023) [arXiv:2208.06246 [gr-qc]].
  • [77] M. Gammon and R. Mann, Slowly Rotating Black Holes in 4D Gauss-Bonnet Gravity, [arXiv:2210.01909 [gr-qc]].
  • [78] Z. Chang and Q. H. Zhu, Revisiting a rotating black hole shadow with astrometric observables, Phys. Rev. D 101, no.8, 084029 (2020) [arXiv:2001.05175 [gr-qc]].
  • [79] C. Liu, T. Zhu, Q. Wu, K. Jusufi, M. Jamil, M. Azreg-Aïnou and A. Wang, Shadow and quasinormal modes of a rotating loop quantum black hole, Phys. Rev. D 101, no.8, 084001 (2020) [erratum: Phys. Rev. D 103, no.8, 089902 (2021)] [arXiv:2003.00477 [gr-qc]].
  • [80] M. Heydari-Fard, M. Heydari-Fard and H. R. Sepangi, Null geodesics and shadow of hairy black holes in Einstein-Maxwell-dilaton gravity, Phys. Rev. D 105, no.12, 124009 (2022) [arXiv:2110.02713 [gr-qc]].
  • [81] M. Heydari-Fard, S. G. Honarvar and M. Heydari-Fard, Thin accretion disc luminosity and its image around rotating black holes in perfect fluid dark matter, Mon. Not. Roy. Astron. Soc. 521, no.1, 708-716 (2023) [arXiv:2210.04173 [gr-qc]].
  • [82] K. Nozari and S. Saghafi, Asymptotically locally flat and AdS higher-dimensional black holes of Einstein–Horndeski–Maxwell gravity in the light of EHT observations: shadow behavior and deflection angle Eur. Phys. J. C 83, no.7, 588 (2023) [arXiv:2305.17237 [gr-qc]].
  • [83] S. Saghafi and K. Nozari, Shadow behavior of the quantum-corrected Schwarzschild black hole immersed in holographic quintessence, JHAP 3, no.1, 31-38 (2022) [arXiv:2306.13767 [hep-th]].
  • [84] S. H. Völkel, E. Barausse, N. Franchini and A. E. Broderick, EHT tests of the strong-field regime of general relativity, Class. Quant. Grav. 38, no.21, 21LT01 (2021) [arXiv:2011.06812 [gr-qc]].
  • [85] G. Lara, S. H. Völkel and E. Barausse, Separating astrophysics and geometry in black hole images, Phys. Rev. D 104, no.12, 124041 (2021) [arXiv:2110.00026 [gr-qc]].
  • [86] R. Shaikh, K. Pal, K. Pal and T. Sarkar, Constraining alternatives to the Kerr black hole, Mon. Not. Roy. Astron. Soc. 506, no.1, 1229-1236 (2021) [arXiv:2102.04299 [gr-qc]].
  • [87] R. A. Konoplya and A. Zhidenko, Shadows of parametrized axially symmetric black holes allowing for separation of variables, Phys. Rev. D 103, no.10, 104033 (2021) [arXiv:2103.03855 [gr-qc]].
  • [88] R. Ghosh, S. Sk and S. Sarkar, Hairy Black Holes: Non-existence of Short Hairs and Bound on Light Ring Size [arXiv:2306.14193 [gr-qc]].
  • [89] A. Tripathi, S. Nampalliwar, A. B. Abdikamalov, D. Ayzenberg, C. Bambi, T. Dauser, J. A. Garcia and A. Marinucci, Toward Precision Tests of General Relativity with Black Hole X-Ray Reflection Spectroscopy, Astrophys. J. 875, no.1, 56 (2019) [arXiv:1811.08148 [gr-qc]].
  • [90] A. Tripathi, A. B. Abdikamalov, D. Ayzenberg, C. Bambi, V. Grinberg and M. Zhou, Testing the Kerr Black Hole Hypothesis with GX 339–4 by a Combined Analysis of Its Thermal Spectrum and Reflection Features, Astrophys. J. 907, no.1, 31 (2021) [arXiv:2010.13474 [astro-ph.HE]].
  • [91] B. H. Lee, W. Lee and Y. S. Myung, Shadow cast by a rotating black hole with anisotropic matter, Phys. Rev. D 103, no.6, 064026 (2021) [arXiv:2101.04862 [gr-qc]].
  • [92] J. L. Rosa and D. Rubiera-Garcia, Shadows of boson and Proca stars with thin accretion disks Phys. Rev. D 106, no.8, 084004 (2022) [arXiv:2204.12949 [gr-qc]].
  • [93] J. L. Rosa, Observational properties of relativistic fluid spheres with thin accretion disks, Phys. Rev. D 107, no.8, 084048 (2023) [arXiv:2302.11915 [gr-qc]].
  • [94] G. J. Olmo, J. L. Rosa, D. Rubiera-Garcia and D. Saez-Chillon Gomez, Shadows and photon rings of regular black holes and geonic horizonless compact objects [arXiv:2302.12064 [gr-qc]].
  • [95] J. L. Rosa, C. F. B. Macedo and D. Rubiera-Garcia, Imaging compact boson stars with hot-spots and thin accretion disks [arXiv:2303.17296 [gr-qc]].
  • [96] Y. Meng, X. M. Kuang and Z. Y. Tang, Photon regions, shadow observables, and constraints from M87* of a charged rotating black hole Phys. Rev. D 106, no.6, 064006 (2022) [arXiv:2204.00897 [gr-qc]].
  • [97] X. M. Kuang, Z. Y. Tang, B. Wang and A. Wang, Constraining a modified gravity theory in strong gravitational lensing and black hole shadow observations Phys. Rev. D 106, no.6, 064012 (2022) [arXiv:2206.05878 [gr-qc]].
  • [98] Z. Y. Tang, X. M. Kuang, B. Wang and W. L. Qian, The length of a compact extra dimension from black hole shadow Sci. Bull. 67, 2272-2275 (2022) [arXiv:2206.08608 [gr-qc]].
  • [99] X. J. Wang, X. M. Kuang, Y. Meng, B. Wang and J. P. Wu, Rings and images of Horndeski hairy black hole illuminated by various thin accretions Phys. Rev. D 107, no.12, 124052 (2023) [arXiv:2304.10015 [gr-qc]].
  • [100] Y. Meng, X. M. Kuang, X. J. Wang, B. Wang and J. P. Wu, Images from disk and spherical accretions of hairy Schwarzschild black holes Phys. Rev. D 108, no.6, 064013 (2023) [arXiv:2306.10459 [gr-qc]].
  • [101] B. Carter, Global structure of the Kerr family of gravitational fields, Phys. Rev. 174, 1559-1571 (1968).
  • [102] B. Carter, “Killing Tensor Quantum Numbers and Conserved Currents in Curved Space,” Phys. Rev. D 16, 3395-3414 (1977).
  • [103] M. Walker and R. Penrose, “On quadratic first integrals of the geodesic equations for type [22] spacetimes,” Commun. Math. Phys. 18, 265-274 (1970).
  • [104] V. P. Frolov, P. Krtous and D. Kubiznak, “Black holes, hidden symmetries, and complete integrability,” Living Rev. Rel. 20, no.1, 6 (2017) [arXiv:1705.05482 [gr-qc]].
  • [105] F. de Felice and G. Preti, “On the Meaning of the separation constant in the Kerr metric,” Class. Quant. Grav. 16, 2929-2935 (1999).
  • [106] S. Vigeland, N. Yunes and L. Stein, Bumpy Black Holes in Alternate Theories of Gravity, Phys. Rev. D 83, 104027 (2011) [arXiv:1102.3706 [gr-qc]].
  • [107] C. M. Will, The Confrontation between general relativity and experiment, Living Rev. Rel. 9, 3 (2006) [arXiv:gr-qc/0510072 [gr-qc]].
  • [108] C. M. Will, Theory and experiment in gravitational physics, (Cambridge University Press, Cambridge, U.K.; New York, U.S.A., 1993), 2nd edition.
  • [109] J. G. Williams, S. G. Turyshev and D. H. Boggs, “Progress in lunar laser ranging tests of relativistic gravity,” Phys. Rev. Lett. 93, 261101 (2004) [arXiv:gr-qc/0411113 [gr-qc]].
  • [110] N. Yunes and L. C. Stein, Non-Spinning Black Holes in Alternative Theories of Gravity, Phys. Rev. D 83, 104002 (2011) [arXiv:1101.2921 [gr-qc]].
  • [111] N. Yunes and F. Pretorius, Dynamical Chern-Simons Modified Gravity. I. Spinning Black Holes in the Slow-Rotation Approximation, Phys. Rev. D 79, 084043 (2009) [arXiv:0902.4669 [gr-qc]].
  • [112] A. N. Aliev and A. E. Gumrukcuoglu, Charged rotating black holes on a 3-brane, Phys. Rev. D 71, 104027 (2005) [arXiv:hep-th/0502223 [hep-th]].
  • [113] K. Hioki and K. i. Maeda, Measurement of the Kerr Spin Parameter by Observation of a Compact Object’s Shadow, Phys. Rev. D 80, 024042 (2009) [arXiv:0904.3575 [astro-ph.HE]].
  • [114] D. N. Page and K. S. Thorne, Disk-Accretion onto a Black Hole. Time-Averaged Structure of Accretion Disk, Astrophys. J. 191, 499-506 (1974).
  • [115] T. Johannsen and D. Psaltis, Testing the No-Hair Theorem with Observations in the Electromagnetic Spectrum: II. Black-Hole Images, Astrophys. J. 718, 446-454 (2010) [arXiv:1005.1931 [astro-ph.HE]].
  • [116] S. Li, T. Mirzaev, A. A. Abdujabbarov, D. Malafarina, B. Ahmedov and W. B. Han, Constraining the deformation of a rotating black hole mimicker from its shadow, Phys. Rev. D 106, no.8, 084041 (2022) [arXiv:2207.10933 [gr-qc]].
  • [117] F. H. Vincent, T. Paumard, E. Gourgoulhon and G. Perrin, GYOTO: a new general relativistic ray-tracing code, Class. Quant. Grav. 28, 225011 (2011) [arXiv:1109.4769 [gr-qc]].