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

\UseRawInputEncoding

Black hole images: A Review

Songbai Chen 1,2 111Corresponding author: [email protected], Jiliang Jing 1,2, Wei-Liang Qian 2,3,4, Bin Wang 2,5 1 Department of Physics, Synergetic Innovation Center for Quantum Effects and Applications, Hunan Normal University, Changsha, Hunan 410081, People’s Republic of China
2 Center for Gravitation and Cosmology, College of Physical Science and Technology, Yangzhou University, Yangzhou 225009, People’s Republic of China
3 Escola de Engenharia de Lorena, Universidade de São Paulo, 12602-810, Lorena, SP, Brazil
4 Faculdade de Engenharia de Guaratinguetá, Universidade Estadual Paulista, 12516-410, Guaratinguetá, SP, Brazil
5 School of Aeronautics and Astronautics, Shanghai Jiao Tong University, Shanghai 200240, People’s Republic of China
Abstract

In recent years, unprecedented progress has been achieved regarding black holes’ observation through the electromagnetic channel. The images of the supermassive black holes M87 and Sgr A released by the Event Horizon Telescope (EHT) Collaboration provided direct visual evidence for their existence, which has stimulated further studies on various aspects of the compact celestial objects. Moreover, the information stored in these images provides a new way to understand the pertinent physical processes that occurred near the black holes, to test alternative theories of gravity, and to furnish insight into fundamental physics. In this review, we briefly summarize the recent developments on the topic. In particular, we elaborate on the features and formation mechanism of black hole shadows, the properties of black hole images illuminated by the surrounding thin accretion disk, and the corresponding polarization patterns. The potential applications of the relevant studies are also addressed.

Key words: black hole shadow, accretion disk, polarization image

pacs:
04.70. Cs, 98.62.Mw, 97.60.Lf

I Introduction

The releasing of the first image of the supermassive black hole M87 by the EHT Collaboration in 2019 Event Horizon Telescope Collaboration et al. (2019a, b, c, d, e, f) is a milestone event in physics. It provided direct visual evidence of black hole in our universe, which means that black hole is no longer just a theoretical model. Combining with the recently published black hole image of Sgr A Event Horizon Telescope Collaboration and Akiyama (2022), it is widely believed that the observational astronomy of black holes has entered a new era of rapid progress.

One of the most important ingredients in the images is the black hole shadow Falcke et al. (2000). It is a two dimensional dark region in the observer’s sky, which is caused by light rays falling into an event horizon of a black hole Synge (1966); Bardeen (1973); Cunningham (1975); Chandrasekhar (1984). The captured light rays are very close to the black hole so that the shape and size of the shadow carry the fingerprint of the celestial object. Therefore, the study of shadows is beneficial to identify black holes, to examine theories of gravity including general relativity, and to further understand some fundamental problems in physics.

From the light propagation in spacetime, black hole shadow depends on the light source, the background spacetime and even the properties of electrodynamics obeyed by photon itself. The light sources in many theoretical investigations are assumed to homogeneously distribute in the total celestial sphere. However, in the real astronomical environment, accretion disk is a kind of actual and feasible light sources around black holes due to its electromagnetic emission. Undoubtedly, the matter configuration and the accretion process in the disks are indelibly imprinted on black hole images. Conversely, analyzing the luminosity distribution and electromagnetic signals stored in the images can extract the information about the matter fields and the physical processes near the black holes. The twisting patterns in the first polarized image of the M87* black hole Event Horizon Telescope Collaboration et al. (2021a, b) revealed the presence of a poloidal magnetic field about 130G1\sim 30G near the black hole. Thus, black hole images with polarized information provide another new way to probe the matter distribution, the electromagnetic interactions and the accretion processes in the strong gravity region of black holes.

The literature related to black hole images is rapidly increasing. Therefore, it is necessary to summarize the existing works at the present time and make prospects for the future. This review focuses on the black hole images, and their rapid development and potential applications. We first introduce the basic concepts of black hole shadows and summarize the main features of the shadows and their formation mechanism, and review how the black hole shadows are determined by the fundamental photon orbits Cunha et al. (2017a) and the corresponding invariant manifolds Grover and Wittig (2017). Next, we introduce the images of the black holes surrounded by a thin disk and their polarization patterns arising from synchrotron radiation. We also discuss the aspects of the potential applications of black hole images.

II Black hole shadows and their formation

It is well known that the shadow of a common object is determined by the light rays passing through the edge of the object. However, for the objects with strong gravity, such as the black hole, the situation is different. According to general relativity, light rays travelling in a black hole spacetime can be deflected due to the gravitational field of the black hole. This phenomenon is known as gravitational lensing, which is analogous to optical lensing Perlick (2004); Blandford and Narayan (1992); Refsdal and Surdej (1994). The deflection angle of the light ray increases with the decreasing of its impact parameter. Therefore, it is easy to infer that the light rays passing very close to the black hole will be captured. Black hole shadow is a dark silhouette observed in the sky originating from these captured light rays. Although the black hole shadow is caused by the photons fallen into the event horizon, its size is not equal to that of this null hypersurface. Actually, for a Schwarzschild black hole, its shadow is about 2.5 times as large as the event horizon in angular size Synge (1966); Bardeen (1973). This can be explained by two reasons. Firstly, it is not only the photons near the event horizon can be captured by a black hole. In fact, there exists a photon sphere outside the event horizon, which is an envelope surface of unstable photon circular orbits in the spacetime Cunha and Herdeiro (2018); Perlick and Tsupko (2022); Wang et al. (2022). The light rays entered the photon sphere will be captured by the black hole as shown in Fig.1. Thus, the boundary of the shadow is determined by the photon sphere rather than the event horizon. Secondly, due to the strong bending of light rays induced by black hole’s gravity, both the size and shape of the observed dark shadow are different from those naively based on Euclidean geometry without gravity.

Refer to caption
Figure 1: The event horizon (the black disk), the photon sphere rps=3Mr_{\rm ps}=3M (the red circle) and the shadow with a radius rsh=33Mr_{\rm sh}=3\sqrt{3}M for a Schwarzschild black hole Wang et al. (2022).

II.1 Features of black hole shadows

The shape and size of black hole shadows depend on the black hole parameters and the observer inclination. For a Schwarzschild black hole, the shadow is a perfect disk for the observer with arbitrary inclination Bardeen (1973); Chandrasekhar (1984). For a rotating Kerr black hole, the shadow also presents a circular silhouette for the observer located on its rotation axis. However, for the observer in the equatorial plane, the shadow gradually becomes a “D”-shaped silhouette with increasing black hole spin Bardeen (1973); Chandrasekhar (1984). For a Konoplya-Zhidenko rotating non-Kerr black hole with an extra deformation parameter described the deviations from the Kerr metric, the “D”-shaped shadow could disappear and the special cuspy shadow could emerge in a certain range of parameter values for the equatorial observer Wang et al. (2017). This is also true for black holes with Proca hair Cunha et al. (2017a); Sengo et al. (2022). Thus, the dependence of shadows on black hole parameters could provide a potential tool to identify black holes in nature. It also triggers the further study of black hole shadows in various theories of gravity Amarilla et al. (2010); Dastan et al. (2022); Long et al. (2020); Stepanian et al. (2021); Prokopov et al. (2022); Younsi et al. (2021).

Black hole shadow also depends on the (non-)integrability of motion equation of photon travelling in the background spacetime. The completely integrable systems are such kind of dynamical systems where the number of the first integrals is equal to its degrees of freedom. Generally, in static spacetimes of black holes with spherical symmetry, such as in the Schwarzschild black hole spacetime, the dynamical system of photons is completely integrable since it possesses three independent first integrals, i.e., the energy EE, the zz-component of the angular momentum LzL_{z} and the Carter constant QQ Carter (1968). This ensures the motion of photons is regular so that black hole shadow has the same shape as the photon sphere

Refer to caption
Figure 2: The eyebrowlike shadows near the primary shadow for a Kerr black hole with scalar hair Cunha et al. (2015).

surface. In these completely integrable systems, the black hole shadow can be calculated by analytical methods Bardeen (1973); Chandrasekhar (1984). However, as the dynamical system of photons is not completely integrable, the null geodesic equations are not be variable separable because there is no existence of a Carter-like constant apart from the usual two integrals of motion EE and LzL_{z}. This implies that the motion of photons could be chaotic and sharply affect the shadow so that its shape is different from that of the photon sphere surface Cunha et al. (2015); Vincent et al. (2016); Cunha et al. (2016); Wang et al. (2018a, b, 2020). In particular, due to chaotic lensing, there are eyebrowlike shadows with the self-similar fractal structure near the primary shadow, as shown in Fig.2. In addition, the black hole shadows in the nonintegrable cases can be only obtained by numerical simulations with the so-called “ray-tracing” codes Bohn et al. (2015); Cunha et al. (2015); Wang et al. (2018a); Riazuelo (2018).

II.2 Formation mechanism of black hole shadows

The photon sphere plays an important role in the formation of black hole shadows. Actually, the photon sphere is composed of unstable photon circular orbits around black holes. The unstable photon circular orbits in the equatorial plane are determined by the effective potential and its derivatives Claudel et al. (2001); Virbhadra and Ellis (2002), i.e., Veff(r)=0V_{\rm eff}(r)=0, Veff(r),r=0V_{{\rm eff}}(r)_{,r}=0 and Veff(r),rr<0V_{{\rm eff}}(r)_{,rr}<0. These unstable orbits can also be obtained by a geometric way with Gauss curvature and geodesic curvature in the optical geometry Qiao (2022). For a static four dimensional spacetime, its optical geometry restricted in the equatorial plane is defined by dt2gijOPdxidxj{\rm d}t^{2}\equiv g^{\rm OP}_{ij}{\rm d}x^{i}{\rm d}x^{j}, i=r,ϕi=r,\phi. The geodesic curvature and Gauss curvature can be expressed as Qiao (2022)

κgeo=12grrOPlngϕϕOPr,\displaystyle\kappa_{\rm geo}=\frac{1}{2\sqrt{g^{\rm OP}_{rr}}}\frac{\partial\ln g^{\rm OP}_{\phi\phi}}{\partial r}, (1)
κGau=1gOP[ϕ(1gϕϕOPgrrOPϕ)\displaystyle\kappa_{\rm Gau}=-\frac{1}{\sqrt{g^{\rm OP}}}\bigg{[}\frac{\partial}{\partial\phi}\bigg{(}\frac{1}{\sqrt{g^{\rm OP}_{\phi\phi}}}\frac{\partial\sqrt{g^{\rm OP}_{rr}}}{\partial\phi}\bigg{)}
+r(1grrOPgϕϕOPr)].\displaystyle+\frac{\partial}{\partial r}\bigg{(}\frac{1}{\sqrt{g^{\rm OP}_{rr}}}\frac{\partial\sqrt{g^{\rm OP}_{\phi\phi}}}{\partial r}\bigg{)}\bigg{]}. (2)

The geodesic curvature κgeo=0\kappa_{\rm geo}=0 gives the radius of the circular photon orbit and the positive (negative) of Gauss curvature κGau\kappa_{\rm Gau} determines that the circular photon orbit is stable (unstable).

Fundamental photon orbits   The unstable photon circular orbits in the equatorial plane are often called light rings. Light rings also affect dynamical properties of ultracompact objects Cunha et al. (2017a). For the ultracompact objects without horizon Cunha et al. (2017b), light rings often come in pairs, one stable and the other unstable. The existence of a stable light ring always implies a spacetime instability Cunha et al. (2022). In the Schwarzschild spacetime, light rings are the only bound photon orbits. In the rotating Kerr spacetime, there are two light rings located in the equatorial plane, one for co-rotating photon and one for counter-rotating photon with respect to the black hole. Moreover, there also exist the non-planar bound photon orbits with constant rr and motion in θ\theta, known as spherical orbits ( see also Fig. 2 in Cunha et al. (2017a)). These spherical orbits are unstable and completely determine the Kerr black hole shadow.

Fundamental photon orbits are the generalization of light rings and spherical orbits in usual stationary and axisymmetric spacetimes Cunha et al. (2017a). The definition of fundamental photon orbits is given in Cunha et al. (2017a). According to the features of orbits, the fundamental photon orbits can be categorized as Xnsnr±X^{n_{r}\pm}_{n_{s}}, where X={O,C}X=\{O,C\}, and nr,ns0n_{r},n_{s}\in\mathbb{N}_{0}.

Refer to caption
Figure 3: Some fundamental photon orbits in the (r,θ)(r,\theta)-plane and their classification. The grey areas represent forbidden regions for the effective potential Cunha et al. (2017a).

The orbit OO is open and it can reach the boundary of the effective potential. The orbit CC is closed and it can not reach the boundary. The sign +()+(-) denotes the even (odd) parity of the orbit under the 2\mathbb{Z}_{2} reflection symmetry around the equatorial plane. nrn_{r} is the number of distinct rr values at where the orbit crosses the equatorial plane. For the light rings lied in the equatorial plane, their nr=0n_{r}=0 because such special kind of orbits never cross the equatorial plane. nsn_{s} is the number of self-intersection points of the orbit. In Fig.3, some fundamental photon orbits and their classification are illustrated in the (r,θ)(r,\theta)-plane.

By making use of the fundamental photon orbits, P. V. P. Cunha et al. Cunha et al. (2017a) explained the formation of the cuspy silhouette of a Kerr black hole with Proca hair shown in Fig.4(a). To clearly demonstrate the formation mechanism of the black hole shadow, ten fundamental photon orbits are selected out and marked by “A1-A4, B1-B3, C1-C3”. Then, the distribution of Δθ|θmaxπ2|\Delta\theta\equiv|\theta_{\text{max}}-\frac{\pi}{2}| and rperir_{\text{peri}} with the impact parameter η\eta are presented for each fundamental photon orbit, where θmax\theta_{\text{max}} is the maximal/minimal angular coordinate at the spherical orbit and rperir_{\text{peri}} is the perimetral radius as a spherical orbit crosses the equatorial plane. The right panel of Fig.4(b) shows the spatial trajectories of the ten fundamental photon orbits in Cartesian coordinates, which move around the black hole. The orbits A1 and C3 are the unstable prograde and retrograde light rings respectively shown as two black circles on the equatorial plane. Other

Refer to caption
Refer to caption
Figure 4: The cuspy shadow (a) and the fundamental photon orbits (b) for the Kerr black hole with Proca hairCunha et al. (2017a).

fundamental photon orbits are non-planar bound photon orbits crossing the equatorial plane. The continuum of fundamental photon orbits can be split into one stable branch (the red dotted line) and two unstable branches (the green and blue lines). The swallow-tail shape pattern related to the fundamental photon orbits in the ηΔθ\eta-\Delta\theta plane yields a jump occurred at A4 and C1 in the ηrperi\eta-r_{\text{peri}} plane. The discontinuity originating from this jump, i.e., rperi(C1)>rperi(A4)r_{\text{peri}(\text{C1})}>r_{\text{peri}(\text{A4})}, induces the emergence of the cuspy shadow. The unstable fundamental photon orbits (C1-B3) non-related to the shadow could be associated to a set of lensing patterns attached to the shadow edge, called “eyelashes” in Fig.4(a). In the black hole spacetimes where there exists a second pair of light rings, the fundamental photon orbits can be classified into two fully disconnected branches: the shadow related branch and the non-shadow related one. If the shadow related branch is connected, the shadow edge will be smooth with no cusp. But the eyelashes, caused by the non-shadow related unstable branch, appear to be disconnected from the shadow, forming a pixelated banana-shaped strip in the lensing image as shown in Fig.2. This feature has been dubbed “ghost shadow” in Sengo et al. (2022). This mechanism is also applied to explain the formation of the cuspy shadow in the Konoplya-Zhidenko rotating non-Kerr black hole spacetime Wang et al. (2017). It is further confirmed that the unstable fundamental photon orbits play an important role in determining the boundary of shadow and the patterns of the shadow shape. In terms of a toy model Qian et al. (2022), the feature of cuspy shadow can be derived by employing the Maxwell construction for phase transition in a two-component system. In addition, the shadows for the black holes with general parameterized metrics have been studied by using the fundamental photon orbits Li et al. (2021); Bambi (2017); Ghasemi-Nodehi et al. (2015); Kumar and Ghosh (2020); Medeiros et al. (2020); Carson and Yagi (2020). These studies could also be beneficial to test the Kerr hypothesis through black hole shadows.

Invariant phase space structures The invariant phase space structures are very important for dynamical systems because they remain invariant under the dynamics. There are several types of invariant structures including fixed points, periodic orbits and invariant manifolds. The simplest is the fixed points. These phase space structures are applied extensively to design space trajectory for various of spacecrafts Mingotti et al. (2009); Koon et al. (2001); Henon (1997); Richardson (1980). Since black hole shadow depends on the dynamics of photons in the background spacetime, the invariant phase space structures should also play an important role in the formation mechanism of the shadow. For a dynamical system of photons travelling in a curved spacetime, its fixed points can be determined by the conditions

x˙μ=Hpμ=0,p˙μ=Hxμ=0,\displaystyle\dot{x}^{\mu}=\frac{\partial H}{\partial p_{\mu}}=0,\;\;\;\;\;\;\;\;\;\;\;\;\;\;\dot{p}_{\mu}=-\frac{\partial H}{\partial x^{\mu}}=0, (3)

where qμ=(t,r,θ,φ)q^{\mu}=(t,r,\theta,\varphi) and pν=(pt,pr,pθ,pφ)p_{\nu}=(p_{t},p_{r},p_{\theta},p_{\varphi}), HH is the Hamiltonian of the system. Actually, the light rings on the equatorial plane are the fixed points for the photon motion Grover and Wittig (2017); Cunha et al. (2017a). In the vicinity of the fixed points, one can linearize the equations (3) and obtain a matrix equation

𝐗˙=J𝐗,\displaystyle\mathbf{\dot{X}}=J\mathbf{X}, (4)

where 𝐗=(qμ,pν)\mathbf{X}=(q^{\mu},p_{\nu}) and JJ is the Jacobian matrix. The eigenvalues μj\mu_{j} of the Jacobian matrix JJ determine the local dynamical properties of the system near the fixed points (see also in Fig.LABEL:lx0). The stable and unstable invariant manifolds correspond to the cases of Re(μj)<0{\rm Re}(\mu_{j})<0 and Re(μj)>0{\rm Re}(\mu_{j})>0, respectively; while the center manifold corresponds to the case with Re(μj)=0{\rm Re}(\mu_{j})=0 where the eigenvalues μj\mu_{j} are pure imaginary numbers. Due to the special properties of the invariant manifolds, there is no trajectory crossing the invariant manifolds. Points in the unstable (stable) invariant manifold move to the fixed points exponentially in backward (forward) time. In terms of Lyapunov’s central limit theorem, the eigenvalue with Re(μj)=0{\rm Re}(\mu_{j})=0 leads to the so-called Lyapunov orbits, which is a one-parameter family γϵ\gamma_{\epsilon} of periodic orbits Mingotti et al. (2009); Koon et al. (2001); Henon (1997); Richardson (1980); Grover and Wittig (2017). These orbits γϵ\gamma_{\epsilon} in the center manifold collapse into a fixed point as ϵ0\epsilon\rightarrow 0. Similarly, the periodic orbits also have their own stable and unstable manifolds. These invariant phase space structures are also shown in Fig. 1 in Grover and Wittig (2017). Obviously, the photon spheres and other periodic orbits can be generalized to the Lyapunov orbits related to the fixed points Grover and Wittig (2017).

These concepts in dynamical systems provide a powerful theoretical foundation for understanding the formation of shadows cast by black holes. The unstable invariant manifold builds a bridge between the photon sphere and the observer because such manifold can approach the fixed points exponentially in backward time Grover and Wittig (2017); Wang et al. (2018a). Only the Lyapunov family of the spherical orbits near the unstable fixed points are responsible for generating the black hole shadow. For a Kerr black hole with scalar hair, there are three unstable light rings 𝔏1\mathfrak{L}_{1}, 𝔏2\mathfrak{L}_{2} and 𝔏3\mathfrak{L}_{3}. However, the Lyapunov orbits associated with 𝔏1\mathfrak{L}_{1} is non-spherical and only the orbits emanating from 𝔏2\mathfrak{L}_{2} and 𝔏3\mathfrak{L}_{3} are spherical Grover and Wittig (2017). The numerical simulated shadow in Fig.5

Refer to caption
Figure 5: Intersections of the unstable manifolds of 𝔏1\mathfrak{L}_{1}, 𝔏2\mathfrak{L}_{2}, and 𝔏3\mathfrak{L}_{3} as well as their Lyapunov orbits with the image plane. Lyapunov orbits related to 𝔏1\mathfrak{L}_{1}, 𝔏2\mathfrak{L}_{2}, and 𝔏3\mathfrak{L}_{3} are marked by red, green, and blue dots, respectively Grover and Wittig (2017).

shows that the complicated and disconnected boundaries of the shadow are completely determined by the Lyapunov spherical orbits. Thus, the invariant manifolds of certain Lyapunov orbits are directly related to black hole shadows even in the case of complicated non-convex, disconnected shadows. Moreover, the curved streamlines in the unstable invariant manifolds could lead to that the shape of the black hole shadow detected by the observer at spatial infinity differs from that of the photon sphere surface Grover and Wittig (2017); Wang et al. (2018a, 2020).

III Image of a black hole with a thin accretion disk

In the real astrophysical conditions, a black hole is surrounded by a hot accretion disk within which it emits a characteristic spectrum of electromagnetic radiation. The electromagnetic radiation emitted by the disk illuminates the background around the black hole and makes the black hole shadow visible. Thus, the accretion disk is the actual light source in the formation of black hole shadows. Clearly, the black hole image cast by the light source with the disk-like structure differs from that by the former homogeneous light source in the previous analyses. Simultaneously, due to the strong gravitational lensing near the central black hole, the shape of the accretion disk is heavily distorted.

Luminet first simulated a photograph of a Schwarzschild black hole with a rotating thin accretion disk Luminet (1979). Here, the proper luminosity of the disk is calculated according to the model described by Page and Thorne Page and Thorne (1974) in its relativistic version, where the intensity of radiation emitted at arbitrary given point of the disk only depends on the radial distance to the black hole. As shown in Fig.11 in Luminet (1979), the flying-saucer-shaped bright region is the primary image of disk, which is formed by the light emitted directly from the upper side of the disk Luminet (1979). Due to the considerable distortion caused by the strong gravitational lensing near the central black hole, the primary image related to the back part of the disk is completely visible rather than hidden by the black hole.

Moreover, one also sees a highly deformed image associated with the bottom of the gaseous disk. This is because the light rays emitted from the bottom side can climb back to the top and reach to the observer at the spatial distance Luminet (1979). Actually, the gravitational lensing gives rise to an infinity of images of the disk, which are caused by the light rays traveling around the black hole any number of times before reaching a distant astronomer Luminet (1979); Bisnovatyi-Kogan and Tsupko (2022). The number of times of light ray crossing the disk determines the order of the image. The higher order images are closer to the central black spot and become thinner and fainter. The inner infinite order image is related to the photon sphere, which represents the actual shadow boundary of the black hole. Generally, it is difficult to distinguish the higher order images optically because they are standing quite closely to each other. The central black area is the black hole shadow formed by the gravitational lensing and capture of light rays.

The existence of a dark gap between the primary image and the higher order images is not surprising because the accretion disk is forbidden to touch the surface of the black hole and then there is not any radiation from the region between the black hole’s event horizon and the inner edge of the disk. Moreover, below the inner stable circular orbit, the disk is unstable so that the gas particles plunge directly towards the black hole without having enough time to emit electromagnetic radiation Luminet (1979). Unlike the shadow itself, the darkness in these patches is of a fundamentally different nature, which may be filled with the emission from lensed images of distant sources in the entire universe although it will also be extremely faint Bronzwaer and Falcke (2021); Bronzwaer et al. (2021).

For a disk around the black hole, the region closer to the horizon is generally brighter because the gas is hotter there. However, the apparent luminosity of the disk’s image for the distant observer is very different from the intrinsic luminosity in the disk. The main reason is that the electromagnetic radiation detected at a great distance undergoes shifts in frequency and intensity with respect to the original radiation emitted directly by the disk Luminet (1979); James et al. (2015). There are two kinds of shift effects. One of them is the so-called gravitational redshift caused by the gravity of the central black hole, which lowers the frequency and decreases the intensity of the electromagnetic radiation. The other is the well-known Doppler effect originating from the displacement of the source with respect to the observer. Doppler effect gives rise to amplification for the approaching source and attenuation for the retreating source. Therefore, for a disk rotating counterclockwise around the black hole, the apparent luminosity of the disk in the left side is brighter than that of in the right side Luminet (1979). The strong gravity of the black hole can give a speed of gas rotation close to the speed of light in the internal regions of an accretion disk, which yields a very strong difference of Doppler shift effects on two sides of the black hole. This strong asymmetry of apparent luminosity is the main signature of the black hole image with a thin accretion disk. In short, the effects from Doppler shift and gravitational redshift drastically modify the luminosity distribution for the observed disk image at large distance.

The black hole spin hardly affects the shape of the primary image. The principal effect of black hole spin is to change the radius of the marginally stable orbit and hence to modify the location of the inner edge of the accretion disk. Unlike in the case of a Schwarzschild black hole, a rapid rotation of Kerr black hole could lead to that the inner edge of the direct image coincides with the higher order images, so the dark gap between them may no longer exist Viergutz (1993a, b). Due to the inner edge of the accretion disk being located far deeper in the gravitational potential, the range of accessible redshift in the disk for the rapidly rotating Kerr black hole is far broader than for the Schwarzschild case. Thus, the higher order images round a rapidly spinning black hole carry less flux than in the Schwarzschild case, which means that they are much more difficult to spatially resolve from the direct image of the disk in the rapidly rotating black hole case.

Moreover, the gravitational field of the accretion disk also affects the propagation of photon and further modifies the shape of black hole shadow. Recently, a static axially symmetric solution, which describes the superposition of a Schwarzschild black hole with a relativistic thin and heavy accretion disk ( Lemos-Letelier disk Lemos and Letelier (1994)), is applied to study black hole shadow Cunha et al. (2020). This static disk with an inner edge is assumed to be made of two streams of counter-rotating particles Lemos and Letelier (1994), which leads to a total vanishing angular momentum and ensures the existence of a static disk in equilibrium with the black hole. A heavy accretion disk yields some new features for the black hole image Cunha et al. (2020). There is a progressive optical enlargement of the disk image covering part of the shadow, despite the fact that the disk is infinitesimally thin. This is a consequence of the increasing light rays’ bending towards the disk due to the increase of disk’s “weight”. The heavy disk also stretches the black hole shadow so that there is an extra deformation of the shadow shape, which becomes more prolate as the disk contributes to a higher fraction of the total mass. Furthermore, the nonintegrability of the photon motion arising from a heavy accretion disk also leads to some chaotic patterns both in the black hole shadow and the disk image. These features also appear in the gravity system of a Schwarzschild black hole surrounded by a massive Bach-Weyl ring Wang et al. (2020). The chaotic lensing also leads to some distinct differences in the shape of photon sphere and the black hole shadow. This is because the chaotic orbits sharply modify the locally measured four-momentum of the photons reaching a distant observer and further influence the celestial coordinates of the images associated with these photons in the observer’s sky, and the latter directly determines the shape of the black hole shadow and the disk image.

IV Polarized image of a black hole

Electromagnetic wave is a kind of transverse waves so the optical image of a black hole must carry the polarization information about the light emitted from the accretion disk around the black hole. Recently, the EHT Collaboration has published the polarimetric image of the black hole M87 Event Horizon Telescope Collaboration et al. (2021a, b). The twisting polarization patterns revealed the existence of magnetic field near the black hole. It is the first time to measure the polarization information characterized by the magnetic field near the black hole, which is helpful to understand the formation of the black hole jet far from 55 million light years.

Actually, in order to extract the information carried in the polarized image of a black hole, one must compare the observed polarimetry data with the theoretical one. Thus, it is very vital to make theoretical analyses and numerical simulations on the polarized images for various black holes. In general, the polarization structures in the black hole images depend on the details of the emitting plasma, principally the magnetic field geometry, and are also affected by the strongly curved spacetime near the black hole. For the origin of the polarized emission around a black hole, there is a typical scenario where the light with high polarization degrees, especially the linearly polarized light, is produced by synchrotron emission in a compact and energetic region of the inner hot disk Tsunetoe et al. (2020, 2021). It is because the relativistic Doppler beaming effect yields that the propagation directions of the photons emitted by a charged relativistic particle are beamed almost along the tangent direction of the particle’s motion so that the light rays in the particle’s orbital plane are linearly polarized. In the cold disk model Page and Thorne (1974), the situation is different, the dominant thermal radiation leads to that the polarized directions of light waves are disorder so that the disk becomes a source of natural light without the total polarization. Thus, in the simulations of the polarized image of a black hole, only the hot disk model is considered. Moreover, as the linearly polarized light passes through the outer magnetized regions in plasma, it further undergoes the Faraday depolarization effects Agol (2000); Marrone et al. (2006); Kuo et al. (2014).

Along the path of each light ray from plasma to observer, the polarization components expressed by the Stokes parameters (I,Q,U,V)(I,Q,U,V) Hamaker and Bregman (1996); Smirnov (2011) satisfy the polarized radiative transfer equations Krolik et al. (2005); Shcherbakov (2008); Shcherbakov and Huang (2011); Dexter (2016); Kulkarni et al. (2011)

ddλ=𝒥𝒦,\displaystyle\frac{\rm d\mathcal{I}}{\rm d\lambda}=\mathcal{J}-\mathcal{K}\mathcal{I}, (5)

where λ\lambda is an affine parameter. The Stokes vector =g3(I,Q,U,V)\mathcal{I}=g^{3}(I,Q,U,V), the propagation matrix 𝒦\mathcal{K}, and the emission vector 𝒥\mathcal{J} describe synchrotron emission and absorption coefficients in all Stokes parameters, as well as Faraday rotation and conversion. Thus, the propagations of the polarized light rays depend heavily on the plasma properties.

In the general relativistic magnetohydrodynamic (GRMHD) simulations, the plasma in the hot disk around the supermassive black hole can be simplified by a model, where the plasma is assumed to be collisionless with electrons and ions so that the electron temperature TeT_{e} deviates from the ion temperature TiT_{i}. The ratio between the temperatures TiT_{i} and TeT_{e} can be expressed as Moscibrodzka et al. (2016); Tsunetoe et al. (2020, 2021)

R=TiTe=Rhighβ21+β2+Rlow11+β2,\displaystyle R=\frac{T_{i}}{T_{e}}=R_{\rm high}\frac{\beta^{2}}{1+\beta^{2}}+R_{\rm low}\frac{1}{1+\beta^{2}}, (6)

where β\beta is the ratio of gas pressure to magnetic pressure. RhighR_{\rm high} and RlowR_{\rm low} are numerical constants, which correspond to the ratio of ion to electron temperatures in the inner disk and in the jet region, respectively.

Through quantitatively evaluating a large library of images based on GRMHD models and comparing with the resolved EHT 2017 linear polarization map of M87 Event Horizon Telescope Collaboration et al. (2021a, b), the viable GRMHD models revealed that the characteristic parameters for average intensity-weighted plasma in the emission region are the electron number density ne1045cm3n_{e}\sim 10^{4-5}cm^{-3}, the magnetic field strength B730GB\simeq 7-30G, and the dimensionless electron temperature θe860\theta_{e}\sim 8-60.

Moreover, recent theoretical investigation shows that the polarization images of M87 jets are very sensitive to the black hole spin Tsunetoe et al. (2020), which could provide a new possibility for measuring the spin parameter of a black hole. In the low-spin case, there are much more symmetric ring shape patterns. This is because the beaming and de-beaming effects are not so large and the jet acceleration is not so significant as the spin is small. In the high-spin case as a=0.99MBHa=0.99M_{\rm BH}, the polarized image of the approaching jet disappeared in the low-spin case is clear Tsunetoe et al. (2020). This is because the high black hole spin gives arise to that the particle motion in the plasma can be accelerated up to the Lorentz factor of ΓL3\Gamma_{L}\sim 3 and further yields that the approaching jet is more bright than the counter one Tsunetoe et al. (2020). Furthermore, there is the crescent-like image produced by the toroidal motion of gas blobs, which demonstrates that the jet acceleration process strongly depends on the black hole spin Nakamura et al. (2018).

One can also extract information about circular polarization through analyzing the Stokes quantity VV in the black hole images. The circular polarization can be amplified by the Faraday conversion in the well-ordered magnetic field. This is different from the case of the linear polarization where the polarization vectors are disordered by the strong Faraday rotation near the black hole. Generally, in a model with hot disk, the circular polarization light images are faint and turbulent because the hot region occupied with chaotic magnetic fields is Faraday thick so that the Faraday conversion cannot be efficient. However, the study of circular polarization images is helpful to understand the polarized information in black hole images more completely. The combination of linear and circular polarizations in future observations could provide a higher-precision detection on the magnetic structure, the temperature distribution and the coupling between proton and electron near black holes. It is shown that the circular polarization images are sensitive to the inclination angle Tsunetoe et al. (2021). Moreover, there is a “separatrix” in the circular polarization images and across which the sign of the circular polarization is reversed. This can be attributed to the helical magnetic field structure in the disk Tsunetoe et al. (2021). It implies that future full polarization EHT images are quite useful tracers of the magnetic field structures near black holes.

The numerical simulations for the polarization image of the black hole are generally computationally expensive due to the broad parameter surveys and the complicated couplings among astrophysical and relativistic effects. Recently, a simple model of an equatorial ring of magnetized fluid has been developed to investigate the polarized images of synchrotron emission around the Schwarzschild black hole Akiyama et al. (2021) and the Kerr black hole Gelles et al. (2021). Although only the emission from a single radius is considered, this model can clearly reveal the dependence of the polarization signatures on the magnetic field configuration, the black hole spin and the observer inclination. Moreover, with this model, the image of a finite thin disk can be produced by simply summing contributions from individual radii. The studies Akiyama et al. (2021); Gelles et al. (2021) also indicates that the ring model image is broadly consistent with the polarization morphology of the EHT image. However, one must note that this simple ring model produces a high fractional polarization (60%\geq 60\%) even after blurring, which is much larger than that in the M87 image where the resolved fractional polarization is about 20%\leq 20\% Akiyama et al. (2021). This suggests that the significant depolarization from the internal Faraday effects is essential when modeling and interpreting the M87 image Moscibrodzka et al. (2017). Nevertheless, the success of the ring model in reproducing the structure of some GRMHD images that have significant Faraday effects is encouraging for the prospects of physical inference from this simple model. Moreover, this simple model can be used to study the loops in the Stokes QUQ-U plane, which describes the continuous variability in the polarization around a black hole Eckart et al. (2006); Trippe et al. (2007); Zamaninasab et al. (2010); Fish et al. (2009); Broderick and Loeb (2006, 2005). It is beneficial to understand some time-varying features of emission from a localized orbiting hotspot near black hole in the real astronomical environment. Thus, this model has been recently applied to study the polarized images of black holes in various spacetimes Qin et al. (2022a); Zhu and Guo (2022); Delijski et al. (2022); Qin et al. (2022b); Liu et al. (2022).

In this simple ring model, the calculation of the polarization vector usually resorts to a so-called Walker-Penrose quantity Walker and Penrose (1970); Himwich et al. (2020). It is conserved along the null geodesic in the spacetimes where the dynamical system of photon motion is integrable and the equation of motion is full variable separable Walker and Penrose (1970). The conserved Walker-Penrose quantity builds a direct connection between the polarization vectors of photon starting from the emitting source and reaching the observer. So in such spacetimes, the propagation of polarization vectors can be calculated by analytical methods, which greatly simplifies the calculation of polarization vectors along null geodesics. However, in the spacetimes where the system of photon motion is nonintegrable, such as, in the Bonnor black dihole spacetime Bonnor (1966), the Walker-Penrose quantity is no longer conserved along null geodesic. Without the help of the Walker-Penrose constant, the calculation of the polarization vectors in this ring model may still rely on the numerical methods. In the Bonnor black dihole spacetime, there exist some fine fractal structures in the distribution of Stokes parameters QQ and UU in the polarized images Zhang et al. (2022a). The signs of QQ and UU are opposite for two adjacent indirect images. It could be caused by that the photons forming two adjacent indirect images are emitted from the upper and lower surfaces of accretion disk, respectively, resulting in a large difference in the corresponding polarization vectors.

V Application prospects of black hole images

The significance of studying black hole images lies in the following aspects. Firstly, such detections can identify black holes and further verify and test the theories of gravity including general relativity, and deepen our understandings on the nature of gravity. Secondly, analyzing information carried in black hole images enables us to understand matter distribution and physical processes around the black holes, and to give further insight into some fundamental problems in physics. In the following, we present some potential application prospects of black hole images.

V.1 Probe the matter distribution around black holes

To probe the matter distribution around black holes, one must simulate images of black hole models by considering different choices and select a model that could accurately represent the main features of the observed images. For the black hole M8787^{*}, it is well known that it belongs to the class of low luminosity active galactic nuclei, and its spectral energy distribution presents features associated with emission from an optically thin and geometrically thick accretion disk ascribed to the synchrotron radiation with an observed brightness temperature in radio wavelengths in the range of 1091010K10^{9}-10^{10}K Event Horizon Telescope Collaboration et al. (2019a). Recently, the most salient features appearing in the EHT Collaboration images of M8787^{*} were reproduced with impressive fidelity and the corresponding configuration model revealed that there may exist an asymmetric bar-like structure attached to a two-temperature thin disk in the equatorial plane of the black hole Boero and Moreschi (2021). Moreover, the asymmetry in brightness is a robust indicator of the orientation of the spin axis. The simulations using different orientations of the black hole spin show that the spin direction opposite to the observed jet is favored by the asymmetric shape of the observed crescent sector.

As mentioned in the previous part, the comparison between the polarization patterns of the M87 image and the viable GRMHD models reveals the existence of magnetic field near the black hole. Actually, the magnetic field can generate some features of black hole images Wang et al. (2021); Junior et al. (2021). For a rotating black hole immersed in a Melvin magnetic field Wang et al. (2021), the shadow becomes oblate for the weak magnetic field. However, in the case with the strong magnetic field, the multiple disconnected shadows emerge, including a middle oblate shadow and many striped shadows. Moreover, the novel feature in the Melvin-Kerr black hole shadow is the gray regions on both sides of the middle main shadow Wang et al. (2021), which are caused by the stable photon orbits around the stable light rings. In fact, the photons moving along the stable photon orbits are trapped and they can’t enter the black hole. Strictly, the gray regions don’t belong to the black hole shadow, but if there are no light sources in the stable photon orbit regions, the observer also see dark shadows in the gray regions Wang et al. (2021); Junior et al. (2021). The chaotic lensing arising from the magnetic field gives rise to the self-similar fractal structures in the black hole shadows. The chaotic image also occurs for the case illuminated by an accretion disk in the Kerr-Melvin black hole spacetime with a strong enough magnetic field Hou et al. (2022). These new effects in shadows could provide a new way to probe the magnetic field near black holes.

The images of black holes indicate that the supermassive black holes in the centers of galaxies are actually surrounded by plasma. Besides as a light source to illuminate black holes, plasma is a dispersive medium where the index of refraction depends on the spacetime point, the plasma frequency and the photon frequency, so the plasma changes the path of the light traveling through it and further affects the geometrical features of black hole shadows Perlick et al. (2015a); Atamurotov and Ahmedov (2015); Abdujabbarov et al. (2016); Crisnejo and Gallo (2018); Huang et al. (2018); Yan (2019); Babar et al. (2020); Matsuno (2021); Tsupko (2021); Li et al. (2022); Atamurotov et al. (2021a); Wang and Wei (2022); Badía and Eiroa (2021); Zhang et al. (2022b); Atamurotov et al. (2021b); Perlick and Tsupko (2017); Perlick et al. (2015b); Bezdekova et al. (2022). The influence of plasma on the shadows depends mainly on the ratio between the plasma frequency and the photon frequency. If the plasma frequency is smaller than the photon frequency, the shadow is not very much different from the vacuum case. However, if the plasma frequency tends to the photon frequency, the significant changes in the photon regions will lead to a drastic modification of the properties of the shadow. In the realistic case where the plasma frequency is much smaller than the photon frequency, the plasma has a decreasing effect on the size of the shadows if the plasma density is higher at the photon sphere than at the observer position. The above analyses are based on an assumption of plasma with radial power-law density. Recent study of angular Gaussian distributed plasma Zhang et al. (2022b), where the plasma is non-spherically symmetric, shows that the effect of plasma can be qualitatively explained by taking the plasma as a convex lens with the refractive index being less than 1. For the supermassive black holes at the centers of the Milky Way and the galaxy M87, which are the main targets of the current observations by the EHT, it is shown that the plasma effects start to become relevant at radio wavelengths of a few centimeters or more. However, the present and planned instruments focus on the submillimeter range, where the scattering and self-absorption have no significant effect on the emitted radiation around the black holes and the plasma effects are very small Perlick and Tsupko (2017); Perlick et al. (2015b), so a realistic observation of the plasma influence on the shadows seems unfeasible at present.

V.2 Constrain black hole parameters and test theories of gravity

It is natural to expect to constrain black hole parameters by the using of shadows because the shape and size of shadows depend on the black hole parameters themselves. In general, since black hole shadows have complex shapes in the observer’s sky, the precise description of the shadow boundaries is crucial for measuring black hole parameters.

Refer to caption
Figure 6: The observables for the apparent shape of the Kerr black hole RsR_{s} and δs=Dcs/Rs\delta_{s}=D_{cs}/R_{s} Hioki and Maeda (2009).

To fit astronomical observations, several observables were constructed by using special points on the shadow boundaries in the celestial coordinates. For the Kerr black hole, the two observables RsR_{s} and δs=Dcs/Rs\delta_{s}=D_{cs}/R_{s} ( as shown in Fig.6) are introduced to measure the approximate size of the shadow and its deformation with respect to the reference circle Hioki and Maeda (2009), respectively. If the inclination angle is given, the values of the mass and spin of the black hole can be obtained by the precise enough measurements of RsR_{s} and δs\delta_{s}. Recently, the length of the shadow boundary and the local curvature radius are introduced to describe the shadow boundary Wei et al. (2019a). The black hole spin and the observer inclination can be constrained by simply measuring the maximum and minimum of the curvature radius. Moreover, a topological covariant quantity is analyzed to measure and distinguish different topological structures of the shadows Wei et al. (2019b); Cunha and Herdeiro (2020). To further describe the general characterization of the shadow boundaries, a coordinate-independent formalism Abdujabbarov et al. (2015) is proposed where the shadow curves Rψ(ψ)R_{\psi}(\psi) are expressed in terms of Legendre polynomials Rψ=l=0clPl(cosψ)R_{\psi}=\sum\limits_{l=0}^{\infty}c_{l}P_{l}(\cos\psi) with the expansion coefficients clc_{l}. The dimensionless deformation parameters δn\delta_{n} are defined to measure the relative difference between the shadow at ψ=0\psi=0 and at other angles ψ=π/n\psi=\pi/n, n=1,2,kn=1,2,...k, and kk is an arbitrarily positive integer. These distortions are both accurate and robust so they can also be implemented to analyse the noisy data.

Above analyses are based on an assumption that the black hole shadows are cast by a bundle of photons in parallel trajectories that originating at infinity. For a realistic black hole surrounded by an accretion disk, the shadow is imprinted on the image of the accretion flow. In principle, comparing a detailed model of the accretion disk around the black hole with astronomical observations will yield a measurement of the size and shape of the shadow. However, it is not feasible to predict the details of the brightness profile of the accretion flow image. The first reason is the incompleteness of accretion disk models, and all theoretical models are simplified by introducing some assumptions so they are impossible to be completely consistent with the real disks. The other reason is the observed variability of the emission in the disk, since the inner accretion flow is highly turbulent and variable in the real astronomical environment. Thus, it is necessary to build a procedure to analyze the observation data that focuses on directly measuring the properties of the shadow in a manner that is not seriously affected by our inability to predict the brightness profile of the rest of the image Psaltis et al. (2015). The gradient method Marr and Hildreth (1980); Canny (1983) is such kind of model-independent algorithms in image processing, which has already been applied successfully to interferometric images to quantify the properties of the turbulent structure of the interstellar magnetic field. The basic concept in this algorithm is that the magnitude of the gradient of the accretion flow image has local maxima at the locations of the steepest gradients, such as, in the case of the expected EHT images, which coincide with the edge of the back hole shadow Psaltis et al. (2015). With the obtained gradient image where the rim of the black hole shadow appears as the most discernible feature, a shadow pattern algorithm matching with the Hough/Radon transform is employed to determine the shape and size of the shadow. This algorithm not only measures the properties of the black hole shadow, but also assesses the statistical significance of the results.

The distinct features of black holes originating from deviation parameters in the alternative theory can help test the general relativity. It is shown that the shadow becomes prolate for the negative deviation parameter and becomes oblate for the positive one Broderick et al. (2014). The large deformation parameter in the Konoplya-Zhidenko rotating non-Kerr black hole yields the special cusp-shaped shadow for the equatorial observer Wang et al. (2017). The large deviation arising from the quadrupole mass moment leads to chaotic shadow and the eyeball-like shadows with the self-similar fractal structures Wang et al. (2018b). The similar features of shadows also appear in other non-Einstein theories of gravity including the quadratic degenerate higher-order scalar-tensor theories Long et al. (2020). Moreover, using the priori known estimates for the mass and distance of M87 based on stellar dynamics Event Horizon Telescope Collaboration et al. (2019a, b, c, d, e, f); Gebhardt et al. (2011); Walsh et al. (2013); Kormendy and Ho (2013), the inferred size of the shadow from the horizon-scale images of the object M87 Event Horizon Telescope Collaboration et al. (2019a) is found to be consistent with that predicted from general relativity for a Schwarzschild black hole within 17%17\% for a 68%68\% confidence interval. However, this measurement still admits other possibilities. The size of the black hole shadow M87 can be used as a proxy to measure the deviations from Kerr metric satisfied weak-field tests Psaltis et al. (2020). For the parameterized Johannsen-Psaltis black hole, it has four lowest-order parameters and the shadow depends primarily on the parameter α13\alpha_{13} and only weakly on spin Johannsen (2013). The 2017 EHT measurement for M87 places a bound on the deviation parameter 3.6<α13<5.9-3.6<\alpha_{13}<5.9 Psaltis et al. (2020). For the modified gravity bumpy Kerr metric Gair and Yunes (2011), the size of the shadow depends primarily on the parameter γ1,2\gamma_{1,2} and the requirement that the shadow size is consistent with the measurement of M87 within 17%17\% gives a constraint on the deviation parameter 5.0<γ1,2<4.9-5.0<\gamma_{1,2}<4.9 Psaltis et al. (2020). For the Konoplya-Rezzolla-Zhidenko metric Konoplya et al. (2016), the EHT measurements results in the constraint 1.2<α1<1.3-1.2<\alpha_{1}<1.3 Psaltis et al. (2020). For these parametric deviation metrics, the measurements of the shadow size lead to significant constraints on the deviation parameters that control the second post-Newtonian orders. This means that the EHT measurement of the size of a black hole leads to metric tests that are inaccessible in the weak-field tests. In general, such parametric tests cannot be connected directly to an underlying property of the alternative theory. Recently, the EHT measurements have been applied to set bounds on the physical parameters, such as, the electric charge Kocherlakota et al. (2021) and the MOG parameter in the Scalar-Tensor-Vector-Gravity Theory Kuang et al. (2022). The quality of the measurements Kocherlakota et al. (2021) is already sufficient to rule out that M87 is a highly charged dilaton black hole, a Reissner-Nordström naked singularity or a Janis-Newman-Winicour naked singularity with large scalar charge. Similarly, it also excludes considerable regions of the space of parameters for the doubly-charged dilaton and the Sen black holes. Such tests are very instructive Amarilla et al. (2010); Bambi and Freese (2009); Mizuno et al. (2018); Cunha et al. (2019a) because they can shed light on which underlying theories are promising candidates and which must be discarded or modified. The constraints and tests from shadows are complementary to those imposed by observations of gravitational waves from stellar-mass sources.

Black hole shadow may also provide a way to test binary black hole. Nowadays, the gravitational-wave events detected by the LIGO-Virgo-KAGRA Collaborations Abbott et al. (2016a, b, 2017a, 2017b, 2017c) confirm the existence of binary black hole system in the universe, and the systems of binary black hole are expected to be common astrophysical systems. The shadows of the colliding between two black holes were simulated by adopting the Kastor-Traschen cosmological multiblack hole solution, which describes the collision of maximally charged black holes with a positive cosmological constant Nitta et al. (2011); Yumoto et al. (2012). Fig.7 shows the change of the shadows with time tt during the collision of the two black holes with equal mass. At t=0t=0, the two black holes are mutually away enough and their shadows are separated. However, each shadow is a little bit elongated in the α\alpha direction because of the interaction between the two black holes. At t=1.6t=1.6, the eyebrowlike shadows appear around the main shadows. The eyebrowlike shadows can be explained by a fact that light rays bypass one black hole of binary system and enter the other one. With the further increase of time, the eyebrowlike structures grow and the main shadows approach each other. Although not discernible in the figure, in fact there appear the fractal structures of the eyebrows, i.e., infinitely many thinner eyebrows at the outer region of these eyebrows as well as at the inner region of the main shadows Yumoto et al. (2012).

Refer to caption
Figure 7: The change of black hole shadows with time tt during the collision of two equal mass black holesYumoto et al. (2012).

As time elapses, the interval between two black hole shadows becomes indefinitely narrower, and it is expected that the black hole shadows eventually merge with each other Yumoto et al. (2012). However, due to the special properties of Kastor-Traschen metric, the recent investigation also implies that there is no observer who will see the merge of black hole shadows even if the black holes coalesce into one Okabayashi et al. (2020). Another important solution of binary black hole with analytical metric form is Majumdar-Papapetrou solution, which describes the geometry of two extremally charged black holes in static equilibrium where gravitational attraction is in balance with electrostatic repulsion. The similar eyebrowlike shadows are found in the Majumdar-Papapetrou binary black hole system Shipley and Dolan (2016).

Actually, these eyebrowlike shadows with fractal structures also appear in other binary black hole systems, such as, in the double-Schwarzschild and double-Kerr black hole systems Cunha et al. (2018) in which two black holes are separated by a conical singularity. These common key features imprinted in the shadows of binary systems, such as disconnected shadows with characteristic eyebrows, open up a new analytic avenue for exploring four dimensional black hole binaries Hertog et al. (2019).

V.3 Fundamental problems in physics

Dark matter The nature of dark matter is one of the most important open fundamental questions of physics. Dark matter is assumed to be an invisible matter, which constitutes the dominant form of matter in the universe and has feeble couplings with the common visible matter at most. Despite extensive observational data supporting its presence on a large scale, dark matter has not been directly detected by any scientific instrument. Dark matter should influence black hole shadow due to its gravitational effects. A simple spherical model consisting of a Schwarzschild black hole with mass MM and a homocentric spherical shell of dark matter halo with mass ΔM\Delta M is applied to tentatively study the effects of dark matter on the black hole shadow Konoplya (2019). It is found that the mass of dark matter and its distance over mass distribution lead to larger radius of shadows. However, it must be pointed out that in this simple model the dark matter is unlikely to manifest itself in the shadows of galactic black holes, unless its concentration near black holes is abnormally high Konoplya (2019).

The effect of dark matter halo on black hole shadows has been studied in the spacetimes of a spherically symmetric black hole and of a rotating black hole Hou et al. (2018); Xu et al. (2018); Jusufi et al. (2020, 2019); Cunha et al. (2019b). It is shown that the structures of the black hole shadows in the cold dark matter (CDM) and scalar field dark matter (SFDM) halos are very similar to the cases of the Schwarzschild and Kerr black holes, respectively. Both dark matter models influence the shadows in a similar way and the sizes of the shadows increase with the dark matter parameter kρcR3k\equiv\rho_{c}R^{3}, where the characteristic density ρc\rho_{c} and the radius RR are related to the distribution of dark matter halo in two models. In general, the influence of the dark matter on the black hole shadows is minor and only becomes significant when kk increases to order of magnitude of 10710^{7} for both CDM and SFDM models Hou et al. (2018). The calculation of the angular radii of the shadows shows that the dark matter halo could influence the shadow of Sgr A at a level of order of magnitude of 103μas10^{-3}\mu as and 105μas10^{-5}\mu as, for CDM and SFDM, respectively. However, it is out of the reach of the current astronomical instruments Hou et al. (2018). The current EHT resolution is 60μas\sim 60\mu as at 230 GHz and will achieve 15μas15\mu as by observing at a higher frequency of 345 GHz and adding more very long baseline interferometry (VLBI) telescopes. The space-based VLBI RadioAstron Kardashev and Khartov (2013) will be able to obtain a resolution of 110μas1-10\mu as. This is still at least three orders of magnitude lower than the resolution required by the CDM model. The black hole shadow has been studied for a rotating black hole solution surrounded by superfluid dark matter and baryonic matter. Using the current values for the parameters of the superfluid dark matter and baryonic density profiles for the Sgr A black hole, it is shown that the effects of the superfluid dark matter and baryonic matter on the sizes of shadows are almost negligible compared to the Kerr vacuum black hole Jusufi et al. (2020). Moreover, comparing with the dark matter, the shadow size increases considerably with the baryonic mass. This can be understood by the fact that the baryonic matter is mostly located in the galactic center. Similarly, the baryonic matter in this model yields an increase of the angular diameter of the shadow of the magnitude 105μas10^{-5}\mu as for the Sgr A black hole Jusufi et al. (2020).

The axion is a hypothetical particle beyond the standard model, which is initially proposed to solve the strong CP (charge-conjugation and parity) problem Peccei and Quinn (1977); Weinberg (1978); Wilczek (1978); Di Luzio et al. (2020). Nowadays, axionlike particles are also introduced in fundamental theories and served as an excellent dark matter candidate so there are many search experiments designed to prob axions Marsh (2016); Sikivie (2021); Raffelt (2008, 1990); Anastassopoulos et al. (2017); Payez et al. (2015). Axion cloud around a rotating black hole may be formed through the superradiance mechanism if the Compton wavelength of axion particle is at the same order of the black hole size Strafuss and Khanna (2005); Brito et al. (2015). Due to the existence of the axion cloud, the axion-electromagnetic-field coupling gives rise to that the position angles of linearly polarized photons emitted near the horizon oscillate periodically Carroll et al. (1990); Plascencia and Urbano (2018); Fujita et al. (2019); Fedderke et al. (2019). Along this line, a novel strategy of detecting axion clouds around supermassive black holes is recently proposed by using the high spatial resolution and polarimetric measurements of the EHT Chen et al. (2020).

Refer to caption
Figure 8: The expected axion parameter space probed by polarimetric observations of M87 (green) and Sgr A (red) for different position angle precisions Chen et al. (2020). The bounds from CAST Anastassopoulos et al. (2017) (gray) and Supernova 1987A (pale yellow) are shown to make a comparison.

Fig.8 presents the axion parameter space which is potentially probed by M87 and Sgr A for different position angle precisions Chen et al. (2020). This method is complementary to the constraints from the black hole spin measurements through gravitational wave detections Arvanitaki et al. (2015). Since the position angle oscillation induced by the axion background does not depend on photon frequency, it is expected that polarimetric measurements at different frequencies in the future can be used to distinguish astrophysical background and to improve the sensitivity of tests of the axion superradiance scenario. Moreover, the possibility of probing ultralight axions by the circular polarization light is also studied in Shakeri and Hajkarim (2022).

Extra dimension The possible existence of extra dimensions is one of the most remarkable predictions of the string theory. The extra spatial dimension could play an important role in fundamental theories within the context of the unification of the physical forces and also in black hole physics. For the high-dimensional black holes, it is shown that the extra dimension influences the shape and size of the shadows Hertog et al. (2019); Singh and Ghosh (2018); Papnoi et al. (2014); Amir et al. (2018). Using the size and deviation from circularity of the shadow of the black hole M87 observed by the EHT collaboration, the curvature radius of AdS5 in the Randall-Sundrum brane-world scenario is bounded by an upper limit l170l\lesssim 170AU Vagnozzi and Visinelli (2019). This upper limit is far from being competitive with current 𝒪\mathcal{O} (mm) scale constraints from precision tests of gravity, but greatly improves the limit l0.535l\lesssim 0.535 Mpc obtained from GW170817 Visinelli et al. (2018). More importantly, it is an independent limit from imaging the dark shadow of M87. Using a rotating black hole solution with a cosmological in the vacuum brane, the black hole shadow together with the observed data of M87 also provides a upper bound for the normalized tidal charge q<0.004q<0.004 Neves (2020a), which is the second best result for the tidal charge to date and is a little higher than the best one q<0.003q<0.003 from a solar system test Boehmer et al. (2008). Moreover, the negative values of the tidal charge are reported to be favored with the M87 and Sgr A data in the brane contexts by the using of Reissner-Nordström-type geometry Bin-Nun (2010); Zakharov (2018); Horvath and Gergely (2013) and a rotating black hole without a cosmological constant Banerjee et al. (2020).

For the case of the compactified extra dimension, the shadow of a rotating uniform black string has been studied where the extra spatial dimension is treated as a compacted circle with the circumference ll Tang et al. (2022). The momentum of photon arising from the fifth dimension enlarges the photon regions and the shadow of the rotating 5D black string while it has slight impact on the distortion. The angular diameter in the EHT observations of M87 leads to the constraint on the length of the compact extra dimension 2.03125 mm l\lesssim l\lesssim 2.6 mm Tang et al. (2022). Similarly, from the observations of Sgr A, the constraints 2.28070 mm l\lesssim l\lesssim 2.6 mm and 2.13115 mm l\lesssim l\lesssim 2.6 mm can be given by the upper bounds of the emission ring and the angular shadow diameter respectively Tang et al. (2022). In particular, within these bounds, the rotating 5D black string spacetime is free from the Gregory-Laflamme instability Tang et al. (2022).

Effects of the specific angular momentum ξψ\xi_{\psi} of photon from the fifth dimension on black hole shadow have also been studied for a rotating squashed Kaluza-Klein black hole Long et al. (2019), which is a kind of interesting Kaluza-Klein type metrics with the special topology and asymptotical structure Wang (2006). It has squashed S3S^{3} horizons so the black hole has a structure similar to a five-dimensional black hole in the vicinity of horizon, but behaves as the four-dimensional black holes with a constant twisted S1S^{1} fiber in the far region. For this special black hole, the radius RsR_{s} of the black hole image in the observer’s sky has different values for the photons with different angular momentum ξψ\xi_{\psi}. The real radius of the black shadow is equal to the minimum value of RsminR_{s_{\rm min}}. Especially, as the black hole parameters lie in a certain special range, it is found that there is no shadow for a black hole since the minimum value Rsmin=0R_{s_{\rm min}}=0 in these special cases Long et al. (2019), which is novel since it does not appear in the usual black hole spacetimes. It must be pointed out that the emergence of black hole without shadow does not mean that light rays can penetrate through the black hole. Actually, it is just because the photons near the black hole with certain range of ξψ\xi_{\psi} change their propagation directions and then become far away from the black hole. The phenomenon of black hole without black shadow will vanish if there exists the further constraint on the specific angular momentum ξψ\xi_{\psi} of photon from the fifth dimension. In the case where black hole shadow exists, the radius of the black hole shadow increases monotonically with the increase of extra dimension parameter in the non-rotating case. With the increasing of rotation parameter, the radius of the black hole shadow gradually becomes a monotonously decreasing function of the extra dimension parameter. With the latest observation data, the angular radii of the shadows for the supermassive black hole Sgr A at the centre of the Milky Way Galaxy and the supermassive black hole in M87 are estimated Long et al. (2019), which implies that there is a room for the theoretical model of such a rotating squashed Kaluza-Klein black hole.

Coupling between the photon and background field Analogous to the motion of charged particles in an electromagnetic field, the propagation of light rays in a spacetime is also influenced by the coupling between the photon and background field, which could leave observable effects on the black hole shadow. In the standard Einstein-Maxwell theory, there is only a quadratic term of Maxwell tensor directly related to electromagnetic field, which can be seen as an interaction between Maxwell field and metric tensor. Actually, the interactions between electromagnetic field and curvature tensor could appear naturally in quantum electrodynamics with the photon effective action originating from one-loop vacuum polarization Drummond and Hathrell (1980). Although these curvature tensor corrections appear firstly as an effective description of quantum effects, the extended theoretical models without the small coupling constant limit have been investigated for some physical motivations Hehl and Obukhov (1999); Balakin et al. (2008); Clarke et al. (2001); Bamba and Odintsov (2008).

The coupling between the photon and Weyl tensor leads to birefringence phenomenon so that the paths of light ray propagations are different for the coupled photons with different polarizations. Thus, it is natural to give rise to double shadows for a single black hole because the natural lights near the black hole can be separated into two kinds of linearly polarized light beams with mutually perpendicular polarizations Huang et al. (2016). With the increase of the coupling strength, the umbra of the black hole decreases and the penumbra increases. In the case of an equatorial thin accretion disk around the Schwarzschild black hole, the black hole image and its polarization distribution are also affected by the coupling strength Zhang et al. (2021). The observed polarized intensity in the bright region is stronger than that in the darker region. It is also noted that the effect of the coupling on the observed polarized vector is weak in general and the stronger effect appears in the bright region close to the black hole in the image plane. Moreover, for the different coupling strengths, the observed polarized patterns have a counterclockwise vortex-like distribution with a rotational symmetry as the observed inclination angle θ0=0\theta_{0}=0^{\circ}. The rotational symmetry in polarized patterns gradually vanishes with the increase of the inclination angle. Quantum electrodynamic effects from the Euler-Heisenberg effective Lagrangian on the shadow have been studied in the black hole background Hu et al. (2021). Similarly, in this case, the birefringence effect also yields that observer sees different shadow sizes of a single black hole for different polarization lights.

The coupling between a photon and a generic vector field is also introduced to study black hole shadow Yan et al. (2020). The generic vector field is assumed to obey the symmetries possessed by the black hole and the boundary condition that the vector field vanishes at infinity. It is found that the black hole shadow in edge-on view also has different appearances for different frequencies of the observed light. This is because the coupling form alters the way that the system depends on the initial conditions. These new phenomena about the black hole shadow originating from the coupling between the photon and background vector field are not simply caused by modifications of the metric, which could help give insight into new physics Yan et al. (2020). In particular, such a kind of coupling can affect the motion of photons and phenomenologically depict a violation of equivalence principle Yan et al. (2020). Thus, it is proposed as a mechanism to test the equivalence principle by analyzing black hole shadows. Although the current observation conditions might not allow us to directly detect these novel phenomena, it is expected that the future project of the next generation EHT with other future multi-band observations Blackburn et al. (2019) as well as the related data-processing techniques could allow for tests of these new physics imprinted in the black hole shadows. Moreover, the shadow images of M87 and Sgr A are recently used to constrain the parameters in the generalized uncertainty principle (GUP) Neves (2020b) and the Lorentz symmetry violation Khodadi and Lambiase (2022), respectively. Although these best upper limits are weaker than those obtained in most other physical frameworks, they are valuable for further understanding black hole images and fundamental problems in physics Vagnozzi et al. (2022); Ozel et al. (2021); Psaltis (2019).

VI Summary

The near-horizon images of the shadows of the supermassive compact objects M87 and Sgr A delivered by the EHT have opened an amazing window for the strong-field test of gravity theories as well as fundamental physics. These images are composed of black hole shadow and the image of accretion disk around the central black hole. Black hole shadow is essentially formed by the light rays entering the black hole’s event horizon, in spite that its shape and size also depend on the position of observer and the types of light sources. The fundamental photon orbits and the invariant phase space structures determine the intrinsic features of the black hole shadow. However, the visualization of the shadow must resort to the emission in the accretion disk around the black hole in the real astronomical environment. This means that the visible images of the black hole also depend on the properties of the accretion disk and the physical processes in the disk, which yields that the black hole images could have a highly model-dependent appearance Psaltis et al. (2015). For example, some models show a partially obscured shadow and others present an apparently exaggerated shadow. Especially, if the disk is optically thick, there may be no visible shadow at all, which means that the geometrical thickness is a key ingredient for observing the shadow. On the other hand, the information on luminance and polarization stored in the image of accretion disk can be helpful to understand the matter distribution and structures in the strong field region near the black hole.

Although black hole shadow and image carry the characteristic information of a black hole, it must be pointed out that the black hole shadows and images in some spacetimes may be not sensitive enough to certain parameters so that the effects of these parameters on the black hole images can not be discriminated in terms of the resolution of the current observation devices. With the increasing accuracy and resolution of the future astronomical observations and the technological development, as well as the more theoretical investigations, it is expected that these mint markings of black holes can be more clearly detected in the next generation EHT, the BlackHoleCam and the space-based experiments. The future detections of the fractural fine structures in black hole shadows arising from the chaotic lensing and the competitive constraints on fundamental physics principles from black hole shadows will help better test theories of gravity and to deeply understand the fundamental problems in modern physics. In a word, the study of black hole images is still in its infancy, and the detection of images for M87 and Sgr A black holes is only a starting point.

VII Acknowledgments

We would like to thank Profs. Carlos Herdeiro and Jieci Wang for their useful comments and suggestions. This work was supported by the National Natural Science Foundation of China under Grant Nos. 12035005, 12275078 and 11875026.

References

  • Event Horizon Telescope Collaboration et al. (2019a) Event Horizon Telescope Collaboration, K. Akiyama, A. Alberdi, et al., Astrophys. J. Lett. 875, L1 (2019a), eprint arXiv:1906.11238.
  • Event Horizon Telescope Collaboration et al. (2019b) Event Horizon Telescope Collaboration, K. Akiyama, A. Alberdi, et al., Astrophys. J. Lett. 875, L2 (2019b), eprint arXiv:1906.11239.
  • Event Horizon Telescope Collaboration et al. (2019c) Event Horizon Telescope Collaboration, K. Akiyama, A. Alberdi, et al., Astrophys. J. Lett. 875, L3 (2019c), eprint arXiv:1906.11240.
  • Event Horizon Telescope Collaboration et al. (2019d) Event Horizon Telescope Collaboration, K. Akiyama, A. Alberdi, et al., Astrophys. J. Lett. 875, L4 (2019d), eprint arXiv:1906.11241.
  • Event Horizon Telescope Collaboration et al. (2019e) Event Horizon Telescope Collaboration, K. Akiyama, A. Alberdi, et al., Astrophys. J. Lett. 875, L5 (2019e), eprint arXiv:1906.11242.
  • Event Horizon Telescope Collaboration et al. (2019f) Event Horizon Telescope Collaboration, K. Akiyama, A. Alberdi, et al., Astrophys. J. Lett. 875, L6 (2019f), eprint arXiv:1906.11243.
  • Event Horizon Telescope Collaboration and Akiyama (2022) Event Horizon Telescope Collaboration and K. Akiyama, Astrophys. J. Lett. 930, L17 (2022).
  • Falcke et al. (2000) H. Falcke, F. Melia, and E. Agol, Astrophys. J. Lett. 528, L13 (2000), eprint astro-ph/9912263.
  • Synge (1966) J. L. Synge, Monthly Notices of the Royal Astronomical Society 131, 463 (1966), ISSN 0035-8711, eprint https://academic.oup.com/mnras/article-pdf/131/3/463/8076763/mnras131-0463.pdf, URL https://doi.org/10.1093/mnras/131.3.463.
  • Bardeen (1973) J. M. Bardeen, in Les Houches Summer School of Theoretical Physics: Black Holes (1973), pp. 215–240.
  • Cunningham (1975) C. T. Cunningham, Astrophys. J. 202, 788 (1975).
  • Chandrasekhar (1984) S. Chandrasekhar, Fundam. Theor. Phys. 9, 5 (1984).
  • Event Horizon Telescope Collaboration et al. (2021a) Event Horizon Telescope Collaboration, K. Akiyama, J. C. Algaba, et al., Astrophys. J. Lett. 910, L12 (2021a), eprint arXiv:2105.01169.
  • Event Horizon Telescope Collaboration et al. (2021b) Event Horizon Telescope Collaboration, K. Akiyama, J. C. Algaba, et al., Astrophys. J. Lett. 910, L13 (2021b), eprint arXiv:2105.01173.
  • Cunha et al. (2017a) P. V. P. Cunha, C. A. R. Herdeiro, and E. Radu, Phys. Rev. D 96, 024039 (2017a), eprint 1705.05461.
  • Grover and Wittig (2017) J. Grover and A. Wittig, Phys. Rev. D 96, 024045 (2017), eprint 1705.07061.
  • Perlick (2004) V. Perlick, Living Rev. Rel. 7, 9 (2004).
  • Blandford and Narayan (1992) R. D. Blandford and R. Narayan, Ann. Rev. Astron. Astrophys. 30, 311 (1992).
  • Refsdal and Surdej (1994) S. Refsdal and J. Surdej, Rept. Prog. Phys. 57, 117 (1994).
  • Cunha and Herdeiro (2018) P. V. P. Cunha and C. A. R. Herdeiro, Gen. Rel. Grav. 50, 42 (2018), eprint 1801.00860.
  • Perlick and Tsupko (2022) V. Perlick and O. Y. Tsupko, Phys. Rept. 947, 1 (2022), eprint 2105.07101.
  • Wang et al. (2022) M. Wang, S. Chen, and J. Jing, Commun. Theor. Phys. 74, 097401 (2022), eprint 2205.05855.
  • Wang et al. (2017) M. Wang, S. Chen, and J. Jing, JCAP 10, 051 (2017), eprint 1707.09451.
  • Sengo et al. (2022) I. Sengo, P. V. P. Cunha, C. A. R. Herdeiro, and E. Radu (2022), eprint 2209.06237.
  • Amarilla et al. (2010) L. Amarilla, E. F. Eiroa, and G. Giribet, Phys. Rev. D 81, 124045 (2010), eprint 1005.0607.
  • Dastan et al. (2022) S. Dastan, R. Saffari, and S. Soroushfar, Eur. Phys. J. Plus 137, 1002 (2022), eprint 1606.06994.
  • Long et al. (2020) F. Long, S. Chen, M. Wang, and J. Jing, Eur. Phys. J. C 80, 1180 (2020), eprint 2009.07508.
  • Stepanian et al. (2021) A. Stepanian, S. Khlghatyan, and V. G. Gurzadyan, Eur. Phys. J. Plus 136, 127 (2021), eprint 2101.08261.
  • Prokopov et al. (2022) V. A. Prokopov, S. O. Alexeyev, and O. I. Zenin, J. Exp. Theor. Phys. 135, 91 (2022), eprint 2107.01115.
  • Younsi et al. (2021) Z. Younsi, D. Psaltis, and F. Özel (2021), eprint 2111.01752.
  • Carter (1968) B. Carter, Phys. Rev. 174, 1559 (1968).
  • Cunha et al. (2015) P. V. P. Cunha, C. A. R. Herdeiro, E. Radu, and H. F. Runarsson, Phys. Rev. Lett. 115, 211102 (2015), eprint 1509.00021.
  • Vincent et al. (2016) F. H. Vincent, E. Gourgoulhon, C. Herdeiro, and E. Radu, Phys. Rev. D 94, 084045 (2016), eprint 1606.04246.
  • Cunha et al. (2016) P. V. P. Cunha, J. Grover, C. Herdeiro, E. Radu, H. Runarsson, and A. Wittig, Phys. Rev. D 94, 104023 (2016), eprint 1609.01340.
  • Wang et al. (2018a) M. Wang, S. Chen, and J. Jing, Phys. Rev. D 97, 064029 (2018a), eprint 1710.07172.
  • Wang et al. (2018b) M. Wang, S. Chen, and J. Jing, Phys. Rev. D 98, 104040 (2018b), eprint 1801.02118.
  • Wang et al. (2020) M. Wang, S. Chen, J. Wang, and J. Jing, Eur. Phys. J. C 80, 110 (2020), eprint 1904.12423.
  • Bohn et al. (2015) A. Bohn, W. Throwe, F. Hébert, K. Henriksson, D. Bunandar, M. A. Scheel, and N. W. Taylor, Class. Quant. Grav. 32, 065002 (2015), eprint 1410.7775.
  • Riazuelo (2018) A. Riazuelo, Int. J. Mod. Phys. D 28, 1950042 (2018), eprint 1511.06025.
  • Claudel et al. (2001) C.-M. Claudel, K. S. Virbhadra, and G. F. R. Ellis, J. Math. Phys. 42, 818 (2001), eprint gr-qc/0005050.
  • Virbhadra and Ellis (2002) K. S. Virbhadra and G. F. R. Ellis, Phys. Rev. D 65, 103004 (2002).
  • Qiao (2022) C.-K. Qiao (2022), eprint 2208.01771.
  • Cunha et al. (2017b) P. V. P. Cunha, E. Berti, and C. A. R. Herdeiro, Phys. Rev. Lett. 119, 251102 (2017b), eprint 1708.04211.
  • Cunha et al. (2022) P. V. P. Cunha, C. Herdeiro, E. Radu, and N. Sanchis-Gual (2022), eprint 2207.13713.
  • Qian et al. (2022) W.-L. Qian, S. Chen, C.-G. Shao, B. Wang, and R.-H. Yue, Eur. Phys. J. C 82, 91 (2022), eprint 2102.03820.
  • Li et al. (2021) S. Li, A. A. Abdujabbarov, and W.-B. Han, Eur. Phys. J. C 81, 649 (2021), eprint 2103.08104.
  • Bambi (2017) C. Bambi, in 14th Marcel Grossmann Meeting on Recent Developments in Theoretical and Experimental General Relativity, Astrophysics, and Relativistic Field Theories (2017), vol. 4, pp. 3494–3499, eprint 1507.05257.
  • Ghasemi-Nodehi et al. (2015) M. Ghasemi-Nodehi, Z. Li, and C. Bambi, Eur. Phys. J. C 75, 315 (2015), eprint 1506.02627.
  • Kumar and Ghosh (2020) R. Kumar and S. G. Ghosh, Astrophys. J. 892, 78 (2020), eprint 1811.01260.
  • Medeiros et al. (2020) L. Medeiros, D. Psaltis, and F. Özel, Astrophys. J. 896, 7 (2020), eprint 1907.12575.
  • Carson and Yagi (2020) Z. Carson and K. Yagi, Phys. Rev. D 101, 084030 (2020), eprint 2002.01028.
  • Mingotti et al. (2009) G. Mingotti, F. Topputo, and F. Bernelli-Zazzera, Celestial Mechanics and Dynamical Astronomy 105, 61 (2009).
  • Koon et al. (2001) W. S. Koon, M. W. Lo, J. E. Marsden, and S. D. Ross, Celestial Mechanics and Dynamical Astronomy 81, 63 (2001).
  • Henon (1997) M. Henon, Generating Families in the Restricted Three-Body Problem (1997).
  • Richardson (1980) D. L. Richardson, Celestial Mechanics 22, 241 (1980).
  • Luminet (1979) J. P. Luminet, A&A 75, 228 (1979).
  • Page and Thorne (1974) D. N. Page and K. S. Thorne, Astrophys. J. 191, 499 (1974).
  • Bisnovatyi-Kogan and Tsupko (2022) G. S. Bisnovatyi-Kogan and O. Y. Tsupko, Phys. Rev. D 105, 064040 (2022), eprint 2201.01716.
  • Bronzwaer and Falcke (2021) T. Bronzwaer and H. Falcke, Astrophys. J. 920, 155 (2021), eprint 2108.03966.
  • Bronzwaer et al. (2021) T. Bronzwaer, J. Davelaar, Z. Younsi, M. Mościbrodzka, H. Olivares, Y. Mizuno, J. Vos, and H. Falcke, Mon. Not. Roy. Astron. Soc. 501, 4722 (2021), eprint 2011.00069.
  • James et al. (2015) O. James, E. von Tunzelmann, P. Franklin, and K. S. Thorne, Class. Quant. Grav. 32, 065001 (2015), eprint 1502.03808.
  • Viergutz (1993a) S. U. Viergutz, A&A 272, 355 (1993a).
  • Viergutz (1993b) S. U. Viergutz, Ap&SS 205, 155 (1993b).
  • Lemos and Letelier (1994) J. P. S. Lemos and P. S. Letelier, Phys. Rev. D 49, 5135 (1994).
  • Cunha et al. (2020) P. V. P. Cunha, N. A. Eiró, C. A. R. Herdeiro, and J. P. S. Lemos, J. Cosmology Astropart. Phys. 2020, 035 (2020), eprint 1912.08833.
  • Tsunetoe et al. (2020) Y. Tsunetoe, S. Mineshige, K. Ohsuga, T. Kawashima, and K. Akiyama, Publ. Astron. Soc. Jap. 72, 32 (2020), eprint 2002.00954.
  • Tsunetoe et al. (2021) Y. Tsunetoe, S. Mineshige, K. Ohsuga, T. Kawashima, and K. Akiyama, Publ. Astron. Soc. Jap. 73, 912 (2021), eprint 2012.05243.
  • Agol (2000) E. Agol, ApJ 538, L121 (2000), eprint astro-ph/0005051.
  • Marrone et al. (2006) D. P. Marrone, J. M. Moran, J.-H. Zhao, and R. Rao, Astrophys. J. 640, 308 (2006), eprint astro-ph/0511653.
  • Kuo et al. (2014) C. Y. Kuo et al., Astrophys. J. Lett. 783, L33 (2014), eprint 1402.5238.
  • Hamaker and Bregman (1996) J. P. Hamaker and J. D. Bregman, A&AS 117, 161 (1996).
  • Smirnov (2011) O. M. Smirnov, A&A 527, A106 (2011), eprint 1101.1764.
  • Krolik et al. (2005) J. H. Krolik, J. F. Hawley, and S. Hirose, Astrophys. J. 622, 1008 (2005), eprint astro-ph/0409231.
  • Shcherbakov (2008) R. Shcherbakov, Astrophys. J. 688, 695 (2008), eprint 0809.0012.
  • Shcherbakov and Huang (2011) R. V. Shcherbakov and L. Huang, Mon. Not. Roy. Astron. Soc. 410, 1052 (2011), eprint 1007.4831.
  • Dexter (2016) J. Dexter, Mon. Not. Roy. Astron. Soc. 462, 115 (2016), eprint 1602.03184.
  • Kulkarni et al. (2011) A. K. Kulkarni, R. F. Penna, R. V. Shcherbakov, J. F. Steiner, R. Narayan, A. Sadowski, Y. Zhu, J. E. McClintock, S. W. Davis, and J. C. McKinney, Mon. Not. Roy. Astron. Soc. 414, 1183 (2011), eprint 1102.0010.
  • Moscibrodzka et al. (2016) M. Moscibrodzka, H. Falcke, and H. Shiokawa, Astron. Astrophys. 586, A38 (2016), eprint 1510.07243.
  • Nakamura et al. (2018) M. Nakamura et al., Astrophys. J. 868, 146 (2018), eprint 1810.09963.
  • Akiyama et al. (2021) K. Akiyama et al. (Event Horizon Telescope), Astrophys. J. 912, 35 (2021), eprint 2105.01804.
  • Gelles et al. (2021) Z. Gelles, E. Himwich, D. C. M. Palumbo, and M. D. Johnson, Phys. Rev. D 104, 044060 (2021), eprint 2105.09440.
  • Moscibrodzka et al. (2017) M. Moscibrodzka, J. Dexter, J. Davelaar, and H. Falcke, Mon. Not. Roy. Astron. Soc. 468, 2214 (2017), eprint 1703.02390.
  • Eckart et al. (2006) A. Eckart, R. Schodel, L. Meyer, S. Trippe, T. Ott, and R. Genzel, Astron. Astrophys. 455, 1 (2006), eprint astro-ph/0610103.
  • Trippe et al. (2007) S. Trippe, T. Paumard, T. Ott, S. Gillessen, F. Eisenhauer, F. Martins, and R. Genzel, Mon. Not. Roy. Astron. Soc. 375, 764 (2007), eprint astro-ph/0611737.
  • Zamaninasab et al. (2010) M. Zamaninasab et al., Astron. Astrophys. 510, A3 (2010), eprint 0911.4659.
  • Fish et al. (2009) V. L. Fish, S. S. Doeleman, A. E. Broderick, A. Loeb, and A. E. E. Rogers, Astrophys. J. 706, 1353 (2009), eprint 0910.3893.
  • Broderick and Loeb (2006) A. E. Broderick and A. Loeb, Mon. Not. Roy. Astron. Soc. 367, 905 (2006), eprint astro-ph/0509237.
  • Broderick and Loeb (2005) A. E. Broderick and A. Loeb, Mon. Not. Roy. Astron. Soc. 363, 353 (2005), eprint astro-ph/0506433.
  • Qin et al. (2022a) X. Qin, S. Chen, and J. Jing, Eur. Phys. J. C 82, 784 (2022a), eprint 2111.10138.
  • Zhu and Guo (2022) H. Zhu and M. Guo (2022), eprint 2205.04777.
  • Delijski et al. (2022) V. Delijski, G. Gyulchev, P. Nedkova, and S. Yazadjiev (2022), eprint 2206.09455.
  • Qin et al. (2022b) X. Qin, S. Chen, Z. Zhang, and J. Jing, Astrophys. J. 938, 2 (2022b), eprint 2207.12034.
  • Liu et al. (2022) X. Liu, S. Chen, and J. Jing (2022), eprint 2205.00391.
  • Walker and Penrose (1970) M. Walker and R. Penrose, Commun. Math. Phys. 18, 265 (1970).
  • Himwich et al. (2020) E. Himwich, M. D. Johnson, A. Lupsasca, and A. Strominger, Phys. Rev. D 101, 084020 (2020), eprint 2001.08750.
  • Bonnor (1966) W. B. Bonnor, Zeitschrift fur Physik 190, 444 (1966).
  • Zhang et al. (2022a) Z. Zhang, S. Chen, and J. Jing, Eur. Phys. J. C 82, 835 (2022a), eprint 2205.13696.
  • Boero and Moreschi (2021) E. F. Boero and O. M. Moreschi, Mon. Not. Roy. Astron. Soc. 507, 5974 (2021), eprint 2105.07075.
  • Wang et al. (2021) M. Wang, S. Chen, and J. Jing, Phys. Rev. D 104, 084021 (2021), eprint 2104.12304.
  • Junior et al. (2021) H. C. D. L. Junior, P. V. P. Cunha, C. A. R. Herdeiro, and L. C. B. Crispino, Phys. Rev. D 104, 044018 (2021), eprint 2104.09577.
  • Hou et al. (2022) Y. Hou, Z. Zhang, H. Yan, M. Guo, and B. Chen, Phys. Rev. D 106, 064058 (2022), eprint 2206.13744.
  • Perlick et al. (2015a) V. Perlick, O. Y. Tsupko, and G. S. Bisnovatyi-Kogan, Phys. Rev. D 92, 104031 (2015a), eprint 1507.04217.
  • Atamurotov and Ahmedov (2015) F. Atamurotov and B. Ahmedov, Phys. Rev. D 92, 084005 (2015), eprint 1507.08131.
  • Abdujabbarov et al. (2016) A. Abdujabbarov, B. Juraev, B. Ahmedov, and Z. Stuchlík, Astrophys. Space Sci. 361, 226 (2016).
  • Crisnejo and Gallo (2018) G. Crisnejo and E. Gallo, Phys. Rev. D 97, 124016 (2018), eprint 1804.05473.
  • Huang et al. (2018) Y. Huang, Y.-P. Dong, and D.-J. Liu, Int. J. Mod. Phys. D 27, 1850114 (2018), eprint 1807.06268.
  • Yan (2019) H. Yan, Phys. Rev. D 99, 084050 (2019), eprint 1903.04382.
  • Babar et al. (2020) G. Z. Babar, A. Z. Babar, and F. Atamurotov, Eur. Phys. J. C 80, 761 (2020), eprint 2008.05845.
  • Matsuno (2021) K. Matsuno, Phys. Rev. D 103, 044008 (2021), eprint 2011.07742.
  • Tsupko (2021) O. Y. Tsupko, Phys. Rev. D 103, 104019 (2021), eprint 2102.00553.
  • Li et al. (2022) Q. Li, Y. Zhu, and T. Wang, Eur. Phys. J. C 82, 2 (2022), eprint 2102.00957.
  • Atamurotov et al. (2021a) F. Atamurotov, K. Jusufi, M. Jamil, A. Abdujabbarov, and M. Azreg-Aïnou, Phys. Rev. D 104, 064053 (2021a), eprint 2109.08150.
  • Wang and Wei (2022) H.-M. Wang and S.-W. Wei, Eur. Phys. J. Plus 137, 571 (2022), eprint 2106.14602.
  • Badía and Eiroa (2021) J. Badía and E. F. Eiroa, Phys. Rev. D 104, 084055 (2021), URL https://link.aps.org/doi/10.1103/PhysRevD.104.084055.
  • Zhang et al. (2022b) Z. Zhang, H. Yan, M. Guo, and B. Chen (2022b), eprint 2206.04430.
  • Atamurotov et al. (2021b) F. Atamurotov, A. Abdujabbarov, and W.-B. Han, Phys. Rev. D 104, 084015 (2021b).
  • Perlick and Tsupko (2017) V. Perlick and O. Y. Tsupko, Phys. Rev. D 95, 104003 (2017), eprint 1702.08768.
  • Perlick et al. (2015b) V. Perlick, O. Y. Tsupko, and G. S. Bisnovatyi-Kogan, Phys. Rev. D 92, 104031 (2015b), URL https://link.aps.org/doi/10.1103/PhysRevD.92.104031.
  • Bezdekova et al. (2022) B. Bezdekova, V. Perlick, and J. Bicak (2022), eprint 2204.05593.
  • Hioki and Maeda (2009) K. Hioki and K.-I. Maeda, Phys. Rev. D 80, 024042 (2009), eprint 0904.3575.
  • Wei et al. (2019a) S.-W. Wei, Y.-C. Zou, Y.-X. Liu, and R. B. Mann, JCAP 08, 030 (2019a), eprint 1904.07710.
  • Wei et al. (2019b) S.-W. Wei, Y.-X. Liu, and R. B. Mann, Phys. Rev. D 99, 041303 (2019b), eprint 1811.00047.
  • Cunha and Herdeiro (2020) P. V. P. Cunha and C. A. R. Herdeiro, Phys. Rev. Lett. 124, 181101 (2020), eprint 2003.06445.
  • Abdujabbarov et al. (2015) A. A. Abdujabbarov, L. Rezzolla, and B. J. Ahmedov, MNRAS 454, 2423 (2015), eprint 1503.09054.
  • Psaltis et al. (2015) D. Psaltis, F. Ozel, C.-K. Chan, and D. P. Marrone, Astrophys. J. 814, 115 (2015), eprint 1411.1454.
  • Marr and Hildreth (1980) D. Marr and E. Hildreth, Proceedings of the Royal Society of London Series B 207, 187 (1980).
  • Canny (1983) J. F. Canny, Theory of Computing Systems Mathematical Systems Theory p. 16 (1983).
  • Broderick et al. (2014) A. E. Broderick, T. Johannsen, A. Loeb, and D. Psaltis, Astrophys. J. 784, 7 (2014), eprint 1311.5564.
  • Gebhardt et al. (2011) K. Gebhardt, J. Adams, D. Richstone, T. R. Lauer, S. M. Faber, K. Gultekin, J. Murphy, and S. Tremaine, Astrophys. J. 729, 119 (2011), eprint 1101.1954.
  • Walsh et al. (2013) J. L. Walsh, A. J. Barth, L. C. Ho, and M. Sarzi, Astrophys. J. 770, 86 (2013), eprint 1304.7273.
  • Kormendy and Ho (2013) J. Kormendy and L. C. Ho, Ann. Rev. Astron. Astrophys. 51, 511 (2013), eprint 1304.7762.
  • Psaltis et al. (2020) D. Psaltis et al. (Event Horizon Telescope), Phys. Rev. Lett. 125, 141104 (2020), eprint 2010.01055.
  • Johannsen (2013) T. Johannsen, Phys. Rev. D 87, 124017 (2013), URL https://link.aps.org/doi/10.1103/PhysRevD.87.124017.
  • Gair and Yunes (2011) J. Gair and N. Yunes, Phys. Rev. D 84, 064016 (2011), eprint 1106.6313.
  • Konoplya et al. (2016) R. Konoplya, L. Rezzolla, and A. Zhidenko, Phys. Rev. D 93, 064015 (2016), eprint 1602.02378.
  • Kocherlakota et al. (2021) P. Kocherlakota et al. (Event Horizon Telescope), Phys. Rev. D 103, 104047 (2021), eprint 2105.09343.
  • Kuang et al. (2022) X.-M. Kuang, Z.-Y. Tang, B. Wang, and A. Wang, Phys. Rev. D 106, 064012 (2022), eprint 2206.05878.
  • Bambi and Freese (2009) C. Bambi and K. Freese, Phys. Rev. D 79, 043002 (2009), eprint 0812.1328.
  • Mizuno et al. (2018) Y. Mizuno, Z. Younsi, C. M. Fromm, O. Porth, M. De Laurentis, H. Olivares, H. Falcke, M. Kramer, and L. Rezzolla, Nature Astron. 2, 585 (2018), eprint 1804.05812.
  • Cunha et al. (2019a) P. V. P. Cunha, C. A. R. Herdeiro, and E. Radu, Phys. Rev. Lett. 123, 011101 (2019a), eprint 1904.09997.
  • Abbott et al. (2016a) B. P. Abbott, R. Abbott, T. D. Abbott, et al., Phys. Rev. Lett. 116, 061102 (2016a), eprint 1602.03837.
  • Abbott et al. (2016b) B. P. Abbott, R. Abbott, T. D. Abbott, et al., Phys. Rev. Lett. 116, 241103 (2016b), eprint 1606.04855.
  • Abbott et al. (2017a) B. P. Abbott, R. Abbott, T. D. Abbott, et al., Phys. Rev. Lett. 118, 221101 (2017a), eprint 1706.01812.
  • Abbott et al. (2017b) B. P. Abbott, R. Abbott, T. D. Abbott, et al., Phys. Rev. Lett. 119, 141101 (2017b), eprint 1709.09660.
  • Abbott et al. (2017c) B. P. Abbott, R. Abbott, T. D. Abbott, et al., ApJ 851, L35 (2017c), eprint 1711.05578.
  • Nitta et al. (2011) D. Nitta, T. Chiba, and N. Sugiyama, Phys. Rev. D 84, 063008 (2011), eprint 1106.2425.
  • Yumoto et al. (2012) A. Yumoto, D. Nitta, T. Chiba, and N. Sugiyama, Phys. Rev. D 86, 103001 (2012), eprint 1208.0635.
  • Okabayashi et al. (2020) K. Okabayashi, N. Asaka, and K.-i. Nakao, Phys. Rev. D 102, 044011 (2020), eprint 2003.07519.
  • Shipley and Dolan (2016) J. Shipley and S. R. Dolan, Class. Quant. Grav. 33, 175001 (2016), eprint 1603.04469.
  • Cunha et al. (2018) P. V. P. Cunha, C. A. R. Herdeiro, and M. J. Rodriguez, Phys. Rev. D 98, 044053 (2018), eprint 1805.03798.
  • Hertog et al. (2019) T. Hertog, T. Lemmens, and B. Vercnocke, Phys. Rev. D 100, 046011 (2019), eprint 1903.05125.
  • Konoplya (2019) R. A. Konoplya, Phys. Lett. B 795, 1 (2019), eprint 1905.00064.
  • Hou et al. (2018) X. Hou, Z. Xu, M. Zhou, and J. Wang, JCAP 07, 015 (2018), eprint 1804.08110.
  • Xu et al. (2018) Z. Xu, X. Hou, X. Gong, and J. Wang, JCAP 09, 038 (2018), eprint 1803.00767.
  • Jusufi et al. (2020) K. Jusufi, M. Jamil, and T. Zhu, Eur. Phys. J. C 80, 354 (2020), eprint 2005.05299.
  • Jusufi et al. (2019) K. Jusufi, M. Jamil, P. Salucci, T. Zhu, and S. Haroon, Phys. Rev. D 100, 044012 (2019), eprint 1905.11803.
  • Cunha et al. (2019b) P. V. P. Cunha, C. A. R. Herdeiro, and E. Radu, Universe 5, 220 (2019b), eprint 1909.08039.
  • Kardashev and Khartov (2013) N. S. Kardashev and V. V. Khartov (RadioAstron), Astronomy Reports 57, 153 (2013), eprint 1303.5013.
  • Peccei and Quinn (1977) R. D. Peccei and H. R. Quinn, Phys. Rev. Lett. 38, 1440 (1977), URL https://link.aps.org/doi/10.1103/PhysRevLett.38.1440.
  • Weinberg (1978) S. Weinberg, Phys. Rev. Lett. 40, 223 (1978), URL https://link.aps.org/doi/10.1103/PhysRevLett.40.223.
  • Wilczek (1978) F. Wilczek, Phys. Rev. Lett. 40, 279 (1978), URL https://link.aps.org/doi/10.1103/PhysRevLett.40.279.
  • Di Luzio et al. (2020) L. Di Luzio, M. Giannotti, E. Nardi, and L. Visinelli, Phys. Rept. 870, 1 (2020), eprint 2003.01100.
  • Marsh (2016) D. J. E. Marsh, Phys. Rept. 643, 1 (2016), eprint 1510.07633.
  • Sikivie (2021) P. Sikivie, Rev. Mod. Phys. 93, 015004 (2021), eprint 2003.02206.
  • Raffelt (2008) G. G. Raffelt, Lect. Notes Phys. 741, 51 (2008), eprint hep-ph/0611350.
  • Raffelt (1990) G. G. Raffelt, Phys. Rept. 198, 1 (1990).
  • Anastassopoulos et al. (2017) V. Anastassopoulos et al. (CAST), Nature Phys. 13, 584 (2017), eprint 1705.02290.
  • Payez et al. (2015) A. Payez, C. Evoli, T. Fischer, M. Giannotti, A. Mirizzi, and A. Ringwald, JCAP 02, 006 (2015), eprint 1410.3747.
  • Strafuss and Khanna (2005) M. J. Strafuss and G. Khanna, Phys. Rev. D 71, 024034 (2005), eprint gr-qc/0412023.
  • Brito et al. (2015) R. Brito, V. Cardoso, and P. Pani, Lect. Notes Phys. 906, pp.1 (2015), eprint 1501.06570.
  • Carroll et al. (1990) S. M. Carroll, G. B. Field, and R. Jackiw, Phys. Rev. D 41, 1231 (1990), URL https://link.aps.org/doi/10.1103/PhysRevD.41.1231.
  • Plascencia and Urbano (2018) A. D. Plascencia and A. Urbano, JCAP 04, 059 (2018), eprint 1711.08298.
  • Fujita et al. (2019) T. Fujita, R. Tazaki, and K. Toma, Phys. Rev. Lett. 122, 191101 (2019), eprint 1811.03525.
  • Fedderke et al. (2019) M. A. Fedderke, P. W. Graham, and S. Rajendran, Phys. Rev. D 100, 015040 (2019), eprint 1903.02666.
  • Chen et al. (2020) Y. Chen, J. Shu, X. Xue, Q. Yuan, and Y. Zhao, Phys. Rev. Lett. 124, 061102 (2020), eprint 1905.02213.
  • Arvanitaki et al. (2015) A. Arvanitaki, M. Baryakhtar, and X. Huang, Phys. Rev. D 91, 084011 (2015), eprint 1411.2263.
  • Shakeri and Hajkarim (2022) S. Shakeri and F. Hajkarim (2022), eprint 2209.13572.
  • Singh and Ghosh (2018) B. P. Singh and S. G. Ghosh, Annals Phys. 395, 127 (2018), eprint 1707.07125.
  • Papnoi et al. (2014) U. Papnoi, F. Atamurotov, S. G. Ghosh, and B. Ahmedov, Phys. Rev. D 90, 024073 (2014), eprint 1407.0834.
  • Amir et al. (2018) M. Amir, B. P. Singh, and S. G. Ghosh, Eur. Phys. J. C 78, 399 (2018), eprint 1707.09521.
  • Vagnozzi and Visinelli (2019) S. Vagnozzi and L. Visinelli, Phys. Rev. D 100, 024020 (2019), eprint 1905.12421.
  • Visinelli et al. (2018) L. Visinelli, N. Bolis, and S. Vagnozzi, Phys. Rev. D 97, 064039 (2018), eprint 1711.06628.
  • Neves (2020a) J. C. S. Neves, Eur. Phys. J. C 80, 717 (2020a), eprint 2005.00483.
  • Boehmer et al. (2008) C. G. Boehmer, T. Harko, and F. S. N. Lobo, Class. Quant. Grav. 25, 045015 (2008), eprint 0801.1375.
  • Bin-Nun (2010) A. Y. Bin-Nun, Phys. Rev. D 82, 064009 (2010), URL https://link.aps.org/doi/10.1103/PhysRevD.82.064009.
  • Zakharov (2018) A. F. Zakharov, Eur. Phys. J. C 78, 689 (2018), eprint 1804.10374.
  • Horvath and Gergely (2013) Z. Horvath and L. A. Gergely, Astron. Nachr. 334, 1047 (2013), eprint 1203.6576.
  • Banerjee et al. (2020) I. Banerjee, S. Chakraborty, and S. SenGupta, Phys. Rev. D 101, 041301 (2020), eprint 1909.09385.
  • Tang et al. (2022) Z.-Y. Tang, X.-M. Kuang, B. Wang, and W.-L. Qian (2022), eprint 2206.08608.
  • Long et al. (2019) F. Long, J. Wang, S. Chen, and J. Jing, JHEP 10, 269 (2019), eprint 1906.04456.
  • Wang (2006) T. Wang, Nucl. Phys. B 756, 86 (2006), eprint hep-th/0605048.
  • Drummond and Hathrell (1980) I. T. Drummond and S. J. Hathrell, Physical Review D 22, 343 (1980).
  • Hehl and Obukhov (1999) F. W. Hehl and Y. N. Obukhov, Lecture Notes in Physics 562, 479 (1999).
  • Balakin et al. (2008) A. B. Balakin, V. V. Bochkarev, and J. P. S. Lemos, Phys. Rev. D 77, 084013 (2008), eprint 0712.4066.
  • Clarke et al. (2001) T. E. Clarke, P. P. Kronberg, and H. Boehringer, Astrophys. J. Lett. 547, L111 (2001), eprint astro-ph/0011281.
  • Bamba and Odintsov (2008) K. Bamba and S. D. Odintsov, JCAP 04, 024 (2008), eprint 0801.0954.
  • Huang et al. (2016) Y. Huang, S. Chen, and J. Jing, European Physical Journal C 76, 594 (2016), eprint 1606.04634.
  • Zhang et al. (2021) Z. Zhang, S. Chen, X. Qin, and J. Jing, Eur. Phys. J. C 81, 991 (2021), eprint 2106.07981.
  • Hu et al. (2021) Z. Hu, Z. Zhong, P.-C. Li, M. Guo, and B. Chen, Phys. Rev. D 103, 044057 (2021), eprint 2012.07022.
  • Yan et al. (2020) S.-F. Yan, C. Li, L. Xue, X. Ren, Y.-F. Cai, D. A. Easson, Y.-F. Yuan, and H. Zhao, Phys. Rev. Res. 2, 023164 (2020), eprint 1912.12629.
  • Blackburn et al. (2019) L. Blackburn et al. (2019), eprint 1909.01411.
  • Neves (2020b) J. C. S. Neves, Eur. Phys. J. C 80, 343 (2020b), eprint 1906.11735.
  • Khodadi and Lambiase (2022) M. Khodadi and G. Lambiase (2022), eprint 2206.08601.
  • Vagnozzi et al. (2022) S. Vagnozzi et al. (2022), eprint 2205.07787.
  • Ozel et al. (2021) F. Ozel, D. Psaltis, and Z. Younsi (2021), eprint 2111.01123.
  • Psaltis (2019) D. Psaltis, Gen. Rel. Grav. 51, 137 (2019), eprint 1806.09740.